0% found this document useful (0 votes)
315 views369 pages

New Directions For Chemical Engineering 2022

Uploaded by

Yohan Parmanand
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
315 views369 pages

New Directions For Chemical Engineering 2022

Uploaded by

Yohan Parmanand
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 369

THE NATIONAL ACADEMIES PRESS

This PDF is available at http://nap.edu/26342 SHARE


   

New Directions for Chemical Engineering (2022)

DETAILS

368 pages | 7 x 10 | PAPERBACK


ISBN 978-0-309-26842-4 | DOI 10.17226/26342

CONTRIBUTORS

GET THIS BOOK Committee on Chemical Engineering in the 21st Century: Challenges and
Opportunities; Board on Chemical Sciences and Technology; Division on Earth and
Life Studies; National Academy of Engineering; National Academies of Sciences,
Engineering, and Medicine
FIND RELATED TITLES

SUGGESTED CITATION

National Academies of Sciences, Engineering, and Medicine 2022. New Directions


for Chemical Engineering. Washington, DC: The National Academies Press.
https://doi.org/10.17226/26342.


Visit the National Academies Press at NAP.edu and login or register to get:

– Access to free PDF downloads of thousands of scientific reports


– 10% off the price of print titles
– Email or social media notifications of new titles related to your interests
– Special offers and discounts

Distribution, posting, or copying of this PDF is strictly prohibited without written permission of the National Academies Press.
(Request Permission) Unless otherwise indicated, all materials in this PDF are copyrighted by the National Academy of Sciences.

Copyright © National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Committee on Chemical Engineering in the 21st Century:


Challenges and Opportunities

Board on Chemical Sciences and Technology

Division on Earth and Life Studies

National Academy of Engineering

A Consensus Study Report of

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

THE NATIONAL ACADEMIES PRESS 500 Fifth Street, NW Washington, DC 20001

This material is based upon work supported by the U.S. Department of Energy, Office of Science,
Biological and Environmental Research Program under Award Number DE-SC0019159; the U.S.
Department of Energy, Office of Energy Efficiency & Renewable Energy, Advanced
Manufacturing Office under Award Number DE-EP0000026/89243420FEE400139; and the U.S.
Department of Energy, Office of Fossil Energy and Carbon Management under Award Number
DE–EP0000026/89303018 FFE400005. This report was prepared as an account of work sponsored
by an agency of the United States Government. Neither the United States Government nor any
agency thereof, nor any of their employees, makes any warranty, express or implied; or assumes
any legal liability or responsibility for the accuracy, completeness, or usefulness of any
information, apparatus, product, or process disclosed; or represents that its use would not infringe
privately owned rights. Reference herein to any specific commercial product, process, or service
by trade name, trademark, manufacturer, or otherwise does not necessarily constitute or imply its
endorsement, recommendation, or favoring by the United States Government or any agency
thereof. The views and opinions of authors expressed herein do not necessarily state or reflect those
of the United States Government or any agency thereof. The activity was supported by the National
Science Foundation under Award Number CHE - 1926880, as well as private contributions from
universities, industry, and professional organizations (Appendix D). Any opinions, findings,
conclusions, or recommendations expressed in this publication do not necessarily reflect the views
of the National Science Foundation or any organization or agency that provided support for the
project.

International Standard Book Number-13: 978-0-309-26842-4


International Standard Book Number-10: 0-309-26842-7
Library of Congress Control Number: 2022937743
Digital Object Identifier: https://doi.org/10.17226/26342

Additional copies of this publication are available from the National Academies Press, 500 Fifth
Street, NW, Keck 360, Washington, DC 20001; (800) 624-6242 or (202) 334-3313;
http://www.nap.edu.

Copyright 2022 by the National Academy of Sciences. All rights reserved.

Printed in the United States of America

Suggested citation: National Academies of Sciences, Engineering, and Medicine. 2022. New
Directions for Chemical Engineering. Washington, DC: The National Academies Press.
https://doi.org/10.17226/26342.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

The National Academy of Sciences was established in 1863 by an Act of Congress,


signed by President Lincoln, as a private, nongovernmental institution to advise the
nation on issues related to science and technology. Members are elected by their peers
for outstanding contributions to research. Dr. Marcia McNutt is president.

The National Academy of Engineering was established in 1964 under the charter of the
National Academy of Sciences to bring the practices of engineering to advising the
nation. Members are elected by their peers for extraordinary contributions to
engineering. Dr. John L. Anderson is president.

The National Academy of Medicine (formerly the Institute of Medicine) was established
in 1970 under the charter of the National Academy of Sciences to advise the nation on
medical and health issues. Members are elected by their peers for distinguished
contributions to medicine and health. Dr. Victor J. Dzau is president.

The three Academies work together as the National Academies of Sciences,


Engineering, and Medicine to provide independent, objective analysis and advice to
the nation and conduct other activities to solve complex problems and inform public
policy decisions. The National Academies also encourage education and research,
recognize outstanding contributions to knowledge, and increase public understanding
in matters of science, engineering, and medicine.

Learn more about the National Academies of Sciences, Engineering, and Medicine at
www.nationalacademies.org.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Consensus Study Reports published by the National Academies of Sciences,


Engineering, and Medicine document the evidence-based consensus on the study’s
statement of task by an authoring committee of experts. Reports typically include
findings, conclusions, and recommendations based on information gathered by the
committee and the committee’s deliberations. Each report has been subjected to a
rigorous and independent peer-review process and it represents the position of the
National Academies on the statement of task.

Proceedings published by the National Academies of Sciences, Engineering, and


Medicine chronicle the presentations and discussions at a workshop, symposium, or
other event convened by the National Academies. The statements and opinions
contained in proceedings are those of the participants and are not endorsed by other
participants, the planning committee, or the National Academies.

For information about other products and activities of the National Academies, please
visit www.nationalacademies.org/about/whatwedo.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

COMMITTEE ON CHEMICAL ENGINEERING IN THE 21st CENTURY:


CHALLENGES AND OPPORTUNITIES

Members

ERIC W. KALER, NAE (Chair), Case Western Reserve University


MONTY M. ALGER, NAE, The Pennsylvania State University
GILDA A. BARABINO, NAE, NAM, Olin College of Engineering
GREGG T. BECKHAM, National Renewable Energy Laboratory
DIMITRIS I. COLLIAS, The Procter & Gamble Co.
JUAN J. DE PABLO, NAE, University of Chicago
SHARON C. GLOTZER, NAS, NAE, University of Michigan
PAULA T. HAMMOND, NAS, NAE, NAM, Massachusetts Institute of Technology
ENRIQUE IGLESIA, NAE, University of California, Berkeley
SANGTAE KIM, NAE, Purdue University
SAMIR MITRAGOTRI, NAE, NAM, Harvard University
BABATUNDE A. OGUNNAIKE,1 NAE, University of Delaware
ANNE S. ROBINSON, Carnegie Mellon University
JOSÉ G. SANTIESTEBAN, NAE, ExxonMobil Research and Engineering Company,
retired
RACHEL A. SEGALMAN, NAE, University of California, Santa Barbara
DAVID S. SHOLL, Oak Ridge National Laboratory
KATHLEEN J. STEBE, NAE, University of Pennsylvania
CHERYL TEICH, Teich Process Development, LLC (until September 2020)

Consultants

PHILIP B. HENDERSON, EMD Electronics


REINALDO M. MACHADO, EMD Electronics
LAURA MATZ, EMD Electronics

Staff

MAGGIE L. WALSER, Study Director


BRENNA ALBIN, Program Assistant
BRITTANY BISHOP, Christine Mirzayan Science Policy Fellow
KESIAH CLEMENT, Research Assistant
ANNE MARIE HOUPPERT, Senior Librarian
GURU MADHAVAN, NAE Senior Director of Programs
REBECCA MORGAN, Senior Librarian
NICHOLAS ROGERS, Deputy Director, Program Finance
LIANA VACCARI, Program Officer
JESSICA WOLFMAN, Research Associate
ELISE ZAIDI, Communications Associate

1
Deceased, February 20, 2022

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

BOARD ON CHEMICAL SCIENCES AND TECHNOLOGY

Members

SCOTT COLLICK (Co-Chair), DuPont


JENNIFER SINCLAIR CURTIS (Co-Chair), University of California, Davis
GERARD BAILLELY, The Procter & Gamble Co.
RUBEN G. CARBONELL, NAE, North Carolina State University
JOHN FORTNER, Yale University
KAREN I. GOLDBERG, NAS, University of Pennsylvania
JENNIFER M. HEEMSTRA, Emory University
JODIE L. LUTKENHAUS, Texas A&M University
SHELLEY D. MINTEER, University of Utah
AMY PRIETO, Colorado State University
MEGAN L. ROBERTSON, University of Houston
SALY ROMERO-TORRES, Thermo Fisher Scientific
REBECCA T. RUCK, Merck Research Laboratories
ANUP K. SINGH, Lawrence Livermore National Laboratory
VIJAY SWARUP, ExxonMobil Research and Engineering Corporation

Staff

MAGGIE L. WALSER, Interim Board Director


BRENNA ALBIN, Program Assistant
MEGAN HARRIES, Program Officer
AYANNA LYNCH, Program Assistant
THANH NGUYEN, Finance Business Partner
LINDA NHON, Associate Program Officer
EMMA SCHULMAN, Program Assistant
ABIGAIL ULMAN, Research Assistant
BENJAMIN ULRICH, Senior Program Assistant
LIANA VACCARI, Program Officer
JESSICA WOLFMAN, Research Associate

vi

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Reviewers

This Consensus Study Report was reviewed in draft form by individuals chosen for
their diverse perspectives and technical expertise. The purpose of this independent review is
to provide candid and critical comments that will assist the National Academies of Sciences,
Engineering, and Medicine in making each published report as sound as possible and to ensure
that it meets the institutional standards for quality, objectivity, evidence, and responsiveness
to the study charge. The review comments and draft manuscript remain confidential to protect
the integrity of the deliberative process.
We thank the following individuals for their review of this report:

Nicholas Abbott, NAE, Cornell University


Aristos Aristidou, NAE, Cargill, Inc.
Gretchen Baier, The Dow Chemical Company
Donna Blackmond, NAS/NAE, The Scripps Research Institute
Joan Brennecke, NAE, The University of Texas at Austin
Prodromos Daoutidis, University of Minnesota
Alice Gast, NAE, Imperial College London
Julia Kornfield, NAE, California Institute of Technology
Cato Laurencin, NAS/NAE/NAM, University of Connecticut Health Center
Jodie Lutkenhaus, Texas A&M University
Phillip Westmoreland, North Carolina State University

Although the reviewers listed above provided many constructive comments and
suggestions, they were not asked to endorse the conclusions or recommendations of this report
nor did they see the final draft before its release. The review of this report was overseen by
Thomas Connolly Jr., American Chemical Society, and Elsa Reichmanis, Lehigh
University. They were responsible for making certain that an independent examination of this
report was carried out in accordance with the standards of the National Academies and that
all review comments were carefully considered. Responsibility for the final content rests
entirely with the authoring committee and the National Academies.

vii

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Acknowledgments

This study would not have been completed successfully without the contributions of
many individuals and organizations. The committee would especially like to thank the indi-
viduals who participated in our town hall at the American Institute of Chemical Engineers
(AIChE) 2019 Annual Meeting and the AIChE Virtual Local Section meeting in spring 2020.
We are grateful as well for the insights provided by respondents to our community question-
naire in spring 2021 (Appendix C), as well as by the numerous individuals who spoke to the
committee during an open information-gathering session or otherwise provided input, and we
thank the organizations that contributed financial support for this study (Appendix D). We
also are grateful to Elsevier for providing access to its SciVal tool.

ix

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

In Memory of Babatunde A. Ogunnaike

Professor Babatunde (Tunde) Ogunnaike was a valuable member of the report com-
mittee who passed away just after the report was released in 2022. Tunde’s contributions can
be seen in every part of the report. He had a broad and deep knowledge of our field, and his
perspective and clear thinking both empowered forward thinking and constrained the growth
of bad ideas. His kind spirit and easy style of collaboration made him a true joy to work with;
his fluid and precise writing style, tireless energy, and ability to meet a deadline made him an
ideal committee member. Tunde was a warm and engaged scholar with valuable insights and
a broad vision that spanned many areas of chemical engineering. Beyond his engineering con-
tributions, Tunde was incredibly generous with his time, teaching and mentoring countless
early career scientists and engineers and leading his College. He was a true friend to many,
and those of us who were lucky to know him carry with us a bit of Tunde in the example he
leaves for us. Our community has lost a giant.

Eric W. Kaler, Chair


On Behalf of the Committee and Staff

xi

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Preface

“It is hard to make predictions, especially about the future.”

Attributed to many. And true.

Yet we as a group accepted the challenge of developing a report designed to articulate


the status of and challenges and promising opportunities for the field of chemical engineering
in the United States, as well as benchmark its international stature, for the next 10 to 30 years.
A committee comprising 17 chemical engineers with diverse backgrounds, expertise, and life
experiences explored a question not investigated by the National Academies since the 1980s:
What is the future of chemical engineering?
As the only engineering field with molecules and molecular transformations at its
core, chemical engineering represents an area of intellectual inquiry and commercial applica-
tions that is profoundly important for society’s future advances in such vital areas as energy,
food, water, medicine, and manufacturing. Chemical engineering is also the natural door
through which the implications and applications of molecular biology—writ large in its cur-
rent incarnations, including genetic engineering, personalized medicine, organs-on-chips, and
even artificial intelligence—enter the realm of practice and application. The future of this field
has crucial implications for what the future looks like for everyone.
Despite remarkable advances and contributions, the legacy of chemical engineering
is complicated. As a profession and a discipline, chemical engineering has enabled the cost-
effective production of materials and chemicals. On the other hand, the durability of some of
these products, such as plastics and fluorinated chemicals, continues to have unintended con-
sequences for the environment. At the same time, energy transformations have generated
greenhouse gases that threaten Earth’s climate. It is essential, therefore, that any future ad-
vances in the field address the history of the advances of the past—an emphasis throughout
this report.
The report describes how chemical engineering is well positioned to serve as the en-
abling discipline in advancing the decarbonization of energy systems and materials without
compromising reliability and cost, and while remaining cognizant of the existential threat of
global climate change. In the foreseeable future, no single energy carrier will be able to meet
the energy demands of all sectors, and the work of chemical engineers will play a vital role in
informing the selection of options for the scale-up, delivery, systems integration, and optimi-
zation of the mix of energy carriers that will address the world’s energy needs with lower
carbon emissions and costs across all regions and sectors of society. At the same time, global
pressures associated with climate change, energy demand, and population growth will change,
in unprecedented ways, the ways in which humans meet their needs for food and water. As in
the past, chemical engineers will confront these challenges through such enabling technolo-
gies as precision agriculture, the development of protein alternatives, and the reduction or
elimination of food waste.
Chemical engineers also will be leaders in the engineering of targeted and accessible
solutions for human health. Their domain of influence will range from personalized medicine

xiii

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

xiv Preface

to the application of systems engineering to biology and health. This work will include strate-
gic modification of the molecular pathways and genomic networks involved in the regulation
of both normal physiology and disease states. It will also include the application of systems-
level thinking to the production of and end-of-life considerations for useful materials, includ-
ing polymers and a variety of other hard and soft materials, in a circular economy. Chemical
engineers will lead the way as well in the application of new tools—such as machine learning
and artificial intelligence—to solve complex problems.
As for the U.S. position in chemical engineering, it is critical to note that China is
making large investments in technologies that are either central or highly relevant to chemical
engineering. These investments, combined with China’s accelerating productivity and schol-
arly output, makes investment in the U.S. research enterprise imperative. Failure to make these
investments will cede global leadership not only of chemical engineering, but of technology
more broadly.
I commend the committee members for their enthusiastic engagement and hard work.
We all found our ways to collaborate and communicate while constrained by the COVID-19
pandemic, but I know we also all missed the synergies and spontaneous insights that in-person
conversations would have generated. While we engaged in virtual meetings and chats instead
of face-to-face meetings, at the end of the day, the creative engagement and critical thinking
of the group made it possible to crystallize important ideas. Finally, but of crucial importance,
the expert guidance, gifted diplomacy, and detailed engagement of the National Academies
staff, led by Dr. Maggie Walser and including Kesiah Clement, Dr. Liana Vaccari, and Jessica
Wolfman, made this report possible.

Eric W. Kaler, Chair


Committee on Chemical Engineering in the 21st Century:
Challenges and Opportunities

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Contents

SUMMARY ................................................................................................................................... 1

1 INTRODUCTION ............................................................................................................. 13
Purpose of This Report, 13
Study Scope and Approach, 13
Audiences for This Report, 15
Report Organization, 15

2 CHEMICAL ENGINEERING TODAY ......................................................................... 17


The Discipline, 20
The Profession, 22
Times of Change, 23
Educational Challenges and Opportunities, 27
Growth of Interdisciplinary Work, 31

3 DECARBONIZATION OF ENERGY SYSTEMS......................................................... 32


The Need for Decarbonization, 33
Energy Sources, 35
Energy Carrier Production, 59
Energy Storage, 65
Energy Conversion and Efficiency, 67
Carbon Capture, Use, and Storage, 79
Challenges and Opportunities, 84

4 SUSTAINABLE ENGINEERING SOLUTIONS FOR ENVIRONMENTAL


SYSTEMS .......................................................................................................................... 87
The Water–Energy–Food Nexus, 88
Molecular Science and Engineering of Water Solutions, 93
Feeding a Growing Population, 104
Understanding and Improving Air Quality, 110
Challenges and Opportunities, 115

5 ENGINEERING TARGETED AND ACCESSIBLE MEDICINE ............................. 117


The Role of Biomolecular Engineering in Health and Medicine, 118
Personalized Medicine, 121
Engineering Approaches to Improving Therapeutics, 125
Modeling and Understanding the Microbiome, 131
Design of Materials, Devices, and Delivery Mechanisms, 136
Hygiene and the Role of Chemical Engineering, 143
Engineering Solutions for Accessibility and Equity in Healthcare, 145
Challenges and Opportunities, 148

6 FLEXIBLE MANUFACTURING AND THE CIRCULAR ECONOMY.................. 151


Intersection of Manufacturing and Chemical Engineering, 152
Feedstock Flexibility for Manufacturing of Existing and Advantaged Products, 155

xv

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

xvi Contents

Process Intensification and Distributed Manufacturing, 158


The Circular Economy and Design for End of Life, 163
Challenges and Opportunities, 174

7 NOVEL AND IMPROVED MATERIALS FOR THE 21st CENTURY .................... 176
Polymer Science and Engineering, 177
Complex Fluids and Soft Matter, 180
Biomaterials, 186
Electronic Materials, 192
Challenges and Opportunities, 197

8 TOOLS TO ENABLE THE FUTURE OF CHEMICAL ENGINEERING ............... 199


Data Science and Computational Tools, 200
Modeling and Simulation, 211
Novel Instruments, 218
Sensors, 220
Challenges and Opportunities, 225

9 TRAINING AND FOSTERING THE NEXT GENERATION OF


CHEMICAL ENGINEERS ............................................................................................ 226
The Undergraduate Core Curriculum, 229
Becoming a Chemical Engineer: The Importance of Diversity, 235
Making Chemical Engineering Broadly Accessible, 240
Teaching Undergraduate Students Today and Tomorrow, 242
Teaching Graduate Students Today and Tomorrow, 245
New Learning and Innovation Practices to Address Current Challenges, 247
Conclusion, 250

10 INTERNATIONAL LEADERSHIP .............................................................................. 253


Publication Rates and Citation Analysis, 254
Observations, 261
Conclusion, 261

REFERENCES ......................................................................................................................... 265

APPENDIXES

A LIST OF ACRONYMS ................................................................................................... 308

B JOURNALS USED IN INTERNATIONAL BENCHMARKING .............................. 313

C SUMMARY OF RESULTS OF THE CHEMICAL ENGINEERING


COMMUNITY QUESTIONNAIRE.............................................................................. 318

D ACKNOWLEDGMENTS .............................................................................................. 339

E COMMITTEE MEMBER AND STAFF BIOGRAPHICAL SKETCHES ................ 343

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Summary

Chemical engineering is the engineering of systems—at scales ranging from the


molecular to the macroscopic—that integrate chemical, physical, and biological elements
to design processes and produce materials and products for the benefit of society. Chem-
ical transformations are at the heart of the technologies that enable modern society, and
the work of chemical engineers has affected societies and individual lives around the
world. Without synthetic fertilizers made with chemical engineering processes, for exam-
ple, a Green Revolution to feed the world would not have been possible. Without the
invention of Ziegler-Natta catalysts, polyolefins would not exist, and the myriad benefits
of plastics would not have been realized. Without the invention of tough, stable polymers
such as Teflon® and Kevlar®, the commercial and medical devices made from those pol-
ymers would not have emerged. Without the contributions of many chemical engineers,
the silicon chips, glass materials, and plastics that make up today’s ubiquitous electronic
devices would not have been developed. And without an army of chemical engineers,
there would be no oil and gas industry to power the world and no pharmaceutical industry
to discover and produce the medicines, therapeutics, and vaccines needed for a long and
healthy life. More recently, chemical engineers have contributed to the tools of directed
evolution, which has allowed for the engineering of improved function in proteins, meta-
bolic pathways, and genomes.
Unfortunately, the discoveries of chemical engineers have also been responsible
for the production of chemicals that will persist in the environment indefinitely, green-
house gas emissions that contribute to climate change, plastic materials that accumulate
in landfills and the oceans, and the chemicals of war that have inflicted long-term or per-
manent damage on humans and the environment. Thus the field of chemical engineering
today faces opportunities and challenges not only to innovate for the future, but also to
innovate in ways that repair the unintended consequences of the past.
Chemical engineering is a discipline and a profession that evolved from the roots
of industrial and applied chemistry, which in turn emerged from such ancient chemical
processes as fermentation and leather tanning. Its academic legacy traces back to the late
1880s, when steam engines still powered the world, and internal combustion was a nas-
cent idea. The world has of course changed since then, and continues to do so at a rapid
rate, as illustrated by the pace at which technology is disrupting established practices and
organizations. Yet the core chemical engineering curriculum has evolved more slowly
over the preceding decades, even as the challenges facing engineers have expanded and
become more difficult.
At its most fundamental level, engineering is about solving problems, and it is
natural to expect that as one problem is solved, another will emerge or grow in importance.
The last time the National Academies of Sciences, Engineering, and Medicine surveyed
the challenges and opportunities for chemical engineering—in the 1988 report Frontiers
in Chemical Engineering: Research Needs and Opportunities, better known as the
“Amundson Report”—the conclusions of that study suggested an approach to developing

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

2 New Directions for Chemical Engineering

new technologies and maintaining leadership in established ones. Given that the pace of
change has only increased since the 1980s, and in the face of a rapidly evolving landscape
for higher education in general, a fresh look at what new challenges and opportunities lie
ahead for both the discipline and the profession of chemical engineering could not be
more timely.
Challenges faced today include not only addressing climate change and the en-
ergy transition, but also reducing raw material usage and increasing recycling to move
from a linear to a circular economy, generating and distributing food worldwide while
conserving water and other resources, and creating and scaling the manufacture and dis-
tribution of new medicines and therapies. Across all these applications, chemical engi-
neers have opportunities to address today’s most important problems by collaborating
with multiple disciplines and engaging systems-level thinking. To leverage these oppor-
tunities, now and in the coming decades, chemical engineering will need to define and
pursue new directions. To this end, this report details a vision for the future of chemical
engineering research, innovation, and education.
To provide a framework for discussion in this report, the study committee exam-
ined the role of chemical engineering in addressing key challenges that face society. While
several organizations have outlined grand challenges in various areas, this report focuses
on the areas of energy and the energy transition; water, food, and air; health and medicine;
manufacturing and the circular economy; and materials. Also included is a discussion of
tools and techniques with the potential to enable future advances across all of these areas.
In addition, the report examines the current state of chemical engineering education and
the need for innovation to ensure that the next generation of chemical engineers is
equipped to address the challenges that lie ahead. Observations on U.S. international lead-
ership in chemical engineering are provided as well.

DECARBONIZATION OF ENERGY SYSTEMS

Mitigation of climate change is one of, if not the most, pressing problems facing
humankind and the planet today. Addressing this problem will require decarbonization of
current energy systems, a challenge rendered all the more difficult by the complexity and
magnitude of the energy landscape and the resultant inability of any single energy carrier
to meet the energy demands of all sectors in the foreseeable future. The field of chemical
engineering continues to make important contributions to the scalability, delivery, sys-
tems integration, and optimization of the mix of energy carriers that will meet energy
needs across different regions and sectors of society with lower carbon emissions and
costs. Chemical engineers will enable technological advances at every point in the energy
value chain, from sources to end uses, and bring to bear the systems-level thinking nec-
essary to balance the economic and environmental trade-offs that will be necessary to
transition to a low-carbon energy system.
The increasing market penetration of electric vehicles for personal transportation,
for example, calls for a reimagining of petroleum refineries that were designed to produce
gasoline and diesel fuel as their main products. The transition to a low-carbon energy
system will require a bridging strategy that relies on a hybrid system consisting of a mix

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Summary 3

of energy carriers. Chemical engineering is rooted in the transformation of stored energy


carriers into forms that are more convenient and into chemicals and materials, and chem-
ical engineers have an important opportunity to continue applying their skillsets to non–
fossil-based energy sources and carriers.
In the long term, achieving net-zero carbon emissions will require significant ad-
vances in photochemistry, electrochemistry, and engineering to enable efficient use of the
predominant source of energy for Earth—the solar flux. To this end, novel systems will
be required to improve the efficiency of photon capture and conversion to electrons; im-
prove the storage of electrons; and advance the direct and/or sequential conversion of
photons to energy carriers via reactions with water, nitrogen, and CO2 to produce hydro-
gen, ammonia, and liquid fuels, respectively.
Successful mitigation of climate change and the transition to a low-carbon energy
system will also require chemical engineers to collaborate with other disciplines, includ-
ing chemistry, biology, economics, social science, and others. In the energy sector, coor-
dination between academic researchers and industrial practitioners, as well as interna-
tional collaboration, will be crucial to ensuring that solutions are economically
competitive and deployable at scale.

Recommendation 3-1:1 Across the energy value chain, federal research funding
should be directed to advancing technologies that shift the energy mix to lower-car-
bon-intensity sources; developing novel low- or zero-carbon energy technologies; ad-
vancing the field of photochemistry; minimizing water use associated with energy
systems; and developing cost-effective and secure carbon capture, use, and storage
methods.

Recommendation 3-2: Researchers in academic and government laboratories and


industry practitioners should form interdisciplinary, cross-sector collaborations fo-
cused on pilot- and demonstration-scale projects and modeling and analysis for low-
carbon energy technologies.

SUSTAINABLE ENGINEERING SOLUTIONS


FOR ENVIRONMENTAL SYSTEMS

Chemical engineers have historically played a central role in the energy sector,
but their contributions have been more modest in solving problems in the interconnected
space of water, food, and air quality. Yet while water, food, and air have historically been
the focus of other disciplines, chemical engineers bring both molecular- and systems-level
thinking to pioneering efforts in this highly interconnected space. The positive impact of
chemical engineers will be magnified as they adapt to thinking beyond the traditional unit
operation scale to focus at a global scale. A continued increase in the world’s population
will lead to increased resource demands, a challenge that is key to defining the future of
chemical engineering.
1
The committee’s recommendations are numbered according to the chapter of the main report
in which they appear.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

4 New Directions for Chemical Engineering

Chemical engineers can support water conservation by both designing higher-


efficiency processes and developing methods for using alternative fluids to freshwater.
Specific research opportunities range from better understanding the fundamentals of wa-
ter structure and dynamics to developing membranes and other separation methods. In the
domain of water use and purification, U.S. chemical engineers would benefit from col-
laborations with civil engineers and other scientists and engineers in arid regions that have
more experience with desalination.
Global pressures associated with climate change and population growth will re-
quire substantial changes in the world’s food sources, a need that chemical engineers can
help address through enabling technologies. Specific opportunities for chemical engineers
include precision agriculture, non–animal-based food and low-carbon-intensity food pro-
duction, and reduction or elimination of food waste. Advanced agricultural practices de-
signed to improve yield while reducing demand for both energy and water will require
collaboration with other disciplines, as well as systems-level approaches such as life-cycle
assessments. A particularly valuable opportunity for collaboration is with researchers who
are pioneering initial demonstrations of “lab-grown” foods on small scales.
The Earth’s atmosphere, with its large range of spatial and temporal scales, pre-
sents intriguing challenges for chemical engineers. Chemical engineers have contributed
to fundamental understanding in this area, and their work will continue to contribute to
improving air quality, including through the removal of CO2 and other heat-trapping
gases. Chemical engineers have contributed to current understanding of aerosol particles
in particular, and will have an opportunity to aid in improving air quality by advancing
understanding of the nature and physics of aerosol particles and applying separation tech-
nologies, as well as the molecular- and systems-level thinking that will be necessary to
address this global challenge. Atmospheric science is already an interdisciplinary field
that includes chemistry, physics, meteorology, and climatology, making it a promising
area in which chemical engineering can contribute through increased collaboration.

Recommendation 4-1: Federal research funding should be directed to both basic and
applied research to advance fundamental understanding of the structure and dy-
namics of water and develop the advanced separation technologies necessary to re-
move and recover increasingly challenging contaminants.

Recommendation 4-2: To minimize the land, water, and nutrient demands of agri-
culture and food production, researchers in academic and government laboratories
and industry practitioners should form interdisciplinary, cross-sector collaborations
focused on the scale-up of innovations in metabolic engineering, bioprocess develop-
ment, precision agriculture, and lab-grown foods, as well as the development of sus-
tainable technologies for improved food preservation, storage, and packaging.

ENGINEERING TARGETED AND ACCESSIBLE MEDICINE

There are few areas of science and engineering in which the rate of progress has
been, and continues to be, more rapid than advances in biology and biochemistry aimed

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Summary 5

at treatments and cures for human illness. Specific contributions of chemical engineers
include reactor design and separations, and more recently cell engineering, formulations,
and other aspects of drug manufacturing. Since the first attempts to isolate small mole-
cules from biological organisms and control and reengineer cell behavior, the develop-
ment of biologically derived products has increased, with major advances resulting from
recombinant DNA technology, the sequencing of genomes, the development of polymer-
ase chain reaction, the discovery of induced pluripotent stem cells, and the discovery and
implementation of gene editing.
All of these challenges present opportunities for chemical engineers to apply sys-
tems-level approaches at scales ranging from molecules to manufacturing facilities, and
to coordinate and collaborate across disciplines. Opportunities to apply quantitative chem-
ical engineering skills to immunology include cancer immunotherapies, vaccine design,
and therapeutic treatments for infectious diseases and autoimmune disorders. The devel-
opment of completely noninvasive methods for drug delivery represents an exciting fron-
tier of device- and materials-based strategies. Chemical engineers are also well positioned
to advance work with sustained-release depots and targeted delivery of therapeutics.
In addition, the demand for monoclonal antibodies, therapeutic proteins, and mes-
senger RNA (mRNA) therapeutics will continue to grow, in part in response to the aging
U.S. population. At the same time, the cost to produce biologics and the subsequent cost
to the consumer create pressure to improve flexibility and reduce costs so as to increase
health care equity while maintaining reliability and stability during manufacturing and
distribution. This challenge provides an opportunity for chemical engineers to develop
novel bioprocess and cell-based improvements through collaborations with biologists and
biochemists.

Recommendation 5-1: Federal research investments in biomolecular engineering


should be directed to fundamental research to

 advance personalized medicine and the engineering of biological mole-


cules, including proteins, nucleic acids, and other entities such as viruses
and cells;
 bridge the interface between materials and devices and health;
 improve the use of tools from systems and synthetic biology to under-
stand biological networks and the intersections with data science and
computational approaches; and
 develop engineering approaches to reduce costs and improve equity and
access to health care.

Recommendation 5-2: Researchers in academic and government laboratories and


industry practitioners should form interdisciplinary, cross-sector collaborations to
develop pilot- and demonstration-scale projects in advanced pharmaceutical manu-
facturing processes.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

6 New Directions for Chemical Engineering

FLEXIBLE MANUFACTURING AND THE CIRCULAR ECONOMY

Chemical engineering as a discipline was founded in the need to deal with heter-
ogeneous raw materials, especially petroleum, and this need will be amplified in the tran-
sition to more sustainable feedstocks. The production and manufacturing of useful mate-
rials and molecules enabled by chemical engineers are now creating previously
unforeseen problems that must be solved at scale. Chemical engineers play a critical role
in manufacturing and can thus contribute to more sustainable manufacturing through ef-
ficiency, nimbleness, and process intensification.
A sustainable future will require a shift to a circular economy in which the end of
life of products is accounted for, utilizing new developments and advances in green chem-
istry and engineering. This shift represents another opportunity for chemical engineers to
innovate from the molecular to manufacturing scales. The continued drive toward more
efficient, environmentally friendly, and cost-effective manufacturing processes will ben-
efit from a wider range of feedstocks for the production of chemicals and materials. The
challenge of feedstock flexibility offers chemical engineers an opportunity to develop ad-
vances in reductive chemistry and processes that will allow the use of oxygenated feed-
stocks such as lignocellulosic biomass. Chemical engineers also have substantial oppor-
tunities to develop scaled-out, distributed manufacturing systems and innovative, large-
scale processes that can compete with the conversion of fossil resources.
Current challenges in process design include the need for improvements in dis-
tributed manufacturing and process intensification—areas in which the chemical engi-
neering research community can provide intellectual leadership. Collaborations between
academic researchers and industrial practitioners will be important for demonstration at
process scale. In the transition from a linear to a circular economy, specific opportunities
for chemical engineers include redesigning processes and products to reduce or eliminate
pollution, developing new ways to reduce and utilize waste, designing products to be used
longer and to be recyclable, and designing processes and products using sustainable feed-
stocks.

Recommendation 6-1: Federal research funding should be directed to both basic and
applied research to advance distributed manufacturing and process intensification,
as well as the innovative technologies, including improved product designs and re-
cycling processes, necessary to transition to a circular economy.

Recommendation 6-2: Researchers in academic and government laboratories and


industry practitioners should form interdisciplinary, cross-sector collaborations fo-
cused on pilot- and demonstration-scale projects in advanced manufacturing, in-
cluding scaled-down and scaled-out processes; process intensification; and the tran-
sition from fossil-based organic feedstocks and virgin-extracted inorganic feedstocks
to new, more sustainable feedstocks for chemical and materials manufacturing.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Summary 7

NOVEL AND IMPROVED MATERIALS FOR THE 21st CENTURY

Chemical engineers have a critical role to play in the development of new mate-
rials and materials processes from the molecular to macroscopic scales. Their integration
of theory, modeling, simulation, experiment, and machine learning is accelerating the dis-
covery, design, and innovation of new materials and new materials processes.
Chemical engineers can contribute to materials development across a range of
material types and applications. The combination of molecular-level understanding and
thermodynamic and transport concepts yields important insights and enables advances. In
particular, chemical engineers have a unique role to play in the continued development of
polymer science and engineering because of their understanding of chemical synthesis
and catalysis, thermodynamics, transport and rheology, and process and systems design.
Chemical engineering is also the logical home for research and development of complex
fluids and soft matter. The science and application of nanoparticles by chemical engineers
in both industry and biomedicine are rapidly accelerating, offering the opportunity for
breakthroughs. Chemical engineers play an essential role in advancing the development
of biomaterials for both regenerative engineering and organ-on-a-chip technology, and
chemical engineering principles are at the heart of understanding and improving targeted
drug delivery both spatially and temporally. Chemical engineering expertise around reac-
tor design, separations, and process intensification is critical to the success and growth of
the electronic materials industry.

Recommendation 7-1: Federal and industry research investments in materials


should be directed to

 polymer science and engineering, with a focus on life-cycle considera-


tions, multiscale simulation, artificial intelligence, and structure/
property/processing approaches;
 basic research to build new knowledge in complex fluids and soft matter;
 nanoparticle synthesis and assembly, with the goal of creating new
materials by self- or directed assembly, as well as improvements in the
safety and efficacy of nanoparticle therapies; and
 discovery and design of new reaction schemes and purification processes,
with a steady focus on process intensification, especially for applications
in electronic materials.

TOOLS TO ENABLE THE FUTURE OF CHEMICAL ENGINEERING

Current and future chemical engineers will need to navigate the interface between
the natural world and the data that describe it, as well as use the tools that turn data into
useful information, knowledge, and understanding. Some emerging and future tools will
be developed in other fields but will have a significant impact on the work of chemical

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

8 New Directions for Chemical Engineering

engineers; others will be developed directly by chemical engineers and have an impact in
science and engineering more broadly. Some tools and capabilities will be evolutionary,
with gradual and predictable development and applications, while others will be revolu-
tionary and will change chemical engineering research and practice in ways that may be
difficult to predict or anticipate today. While the list of tools and capabilities—many of
which will drive innovation when used in combination—is virtually endless, this report
focuses on data science and computational tools, modeling and simulation, novel instru-
ments, and sensors.
Developing tools that synthesize available data in real time and frameworks or
models that transform data into information and actionable knowledge could become one
of chemical engineers’ key contributions to society over the next decades. It is easy to
imagine a not-too-distant future characterized by data-on-demand—where data on any-
thing, at any level of granularity, will be readily and instantly accessible. Such a future
suggests profound and exciting opportunities for chemical engineers, who are trained in
process integration and systems-level thinking—skills that will be required to synthesize
disparate data streams into information and knowledge.
The systems thinking, analytical approaches, and creative problem-solving skills
of today’s chemical engineering graduates give them a distinct advantage in using artifi-
cial intelligence in real-world contexts. The evolution of artificial intelligence in the next
decade will have enormous implications not only for the types of problems chemical en-
gineers will be able to solve but also for how they will do so. Chemical engineers are
poised to contribute significantly to the development of modeling and simulation tools
that will influence education, research, and industry. They will continue developing and
disseminating methods, algorithms, techniques, and open-source codes, making it easier
for nonexperts to use computing tools for scientific research.
The increasing operational complexities and decreasing capital investments and
economic margins in the petrochemical industry, coupled with stringent environmental
and quality demands on the manufacture of specialty chemicals and polymers, will con-
tinue to drive increased use of modeling and simulation to run scenarios and test hypoth-
eses. While the pharmaceutical industry currently lags behind the chemical industry in its
use of simulation tools, fundamental changes in regulatory requirements are motivating
greater use of mathematical models and simulation, especially in the rapidly growing
biomanufacturing sector.

Recommendation 8-1: Federal and industry research investments should be directed


to advancing the use of artificial intelligence, machine learning, and other data sci-
ence tools; improving modeling and simulation and life-cycle assessment capabili-
ties; and developing novel instruments and sensors. Such investments should focus
on applications in basic chemical engineering research and materials development,
as well as on accelerating the transition to a low-carbon energy system; improving
the sustainability of food production, water management, and manufacturing; and
increasing the accessibility of health care.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Summary 9

TRAINING AND FOSTERING THE


NEXT GENERATION OF CHEMICAL ENGINEERS

Chemical engineers are in high demand across most professions and job levels,
and chemical engineering provides an excellent foundation for many career paths. The
undergraduate chemical engineering curriculum has served the discipline well and has
continued to evolve, slowly, in response to scientific discoveries, technological advances,
and societal needs. The undergraduate curriculum provides a mathematical framework for
designing and describing (electro-/photo-/bio-) chemical and physical processes across
diverse spatial and temporal scales. Data science and statistics may be delivered most
effectively in a separate course embedded within the core curriculum and taught with
specific emphasis on matters of chemistry and engineering. In addition, experiential learn-
ing is important, and the majority of industrial and academic chemical engineers inter-
viewed by the committee discussed the importance of internships and other practical ex-
periences. However, there are far fewer internships available than the number of students
who would benefit from them, and the density of the core undergraduate curriculum
leaves few openings for incorporating an additional hands-on laboratory course earlier in
the curriculum.
The current chemical engineering curriculum is well suited to preparing students
for a wide variety of industrial roles. Graduate research increasingly encompasses a di-
verse range of topics that do not all require the same level of traditionally curated
knowledge currently delivered in graduate chemical engineering curricula, and so gradu-
ate curricula may need to be adjusted. Internships for graduate students are currently rare,
and new models will need to address issues of equity and inclusion, suitable compensa-
tion, intellectual property considerations, and adequate intern mentoring.
Women and members of historically excluded groups are underrepresented in
chemical engineering relative to their numbers in the general population, even by com-
parison with chemical and biological sciences and related fields. Diversifying the profes-
sion will bring valuable new perspectives, and is therefore essential to the field’s survival
and potential for impact. At all points along their academic path, chemical engineering
students need role models and effective, inclusive mentors, including those who reflect
the diversity of backgrounds that the field needs. Leveraging of professional societies and
associated affinity groups could provide valuable support for people of diverse back-
grounds entering the field, and strong university support for student chapters of profes-
sional organizations will improve access and success.
Additionally, the general affordability of community colleges is a major attraction
for a diverse body of students, ranging from budget-minded high school seniors to non-
traditional students. Increased engagement of transfer students therefore represents an un-
tapped opportunity to broaden participation in and access to the chemical engineering
profession.
Both students from 2-year colleges and those who change their major to chemical
engineering would benefit from a redesign of the curriculum that would allow them to
complete the degree in less time. Better academic and social support structures are needed

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

10 New Directions for Chemical Engineering

to enable successful pathways for these students. New methods that would make it possi-
ble to offer portions of the curriculum in a distributed manner, as well as more general
restructuring, may require flexibility in curriculum design and changes in university pol-
icies and graduation and accreditation requirements.

Recommendation 9-1: Chemical engineering departments should consider revisions


to their undergraduate curricula that would

 help students understand how individual core concepts merge into the
practice of chemical engineering,
 include earlier and more frequent experiential learning through physical
laboratories and virtual simulations, and
 bring mathematics and statistics into the core curriculum in a more
structured manner by either complementing or replacing some of the ed-
ucation that currently occurs outside the core curriculum.

Recommendation 9-2: To provide graduate students with experiential learning op-


portunities, universities, industry, funding agencies, and the American Institute of
Chemical Engineers should coordinate to revise graduate training programs and
funding structures to provide opportunities for and remove barriers to systematic
placement of graduate students in internships.

Recommendation 9-3: To increase recruitment and retention of women and Black,


Indigenous, and People of Color (BIPOC) individuals in undergraduate programs,
chemical engineering departments should emphasize opportunities for chemical en-
gineers to make positive societal impacts, and should build effective mentoring and
support structures for students who are members of such historically excluded
groups. To provide more opportunities for BIPOC students, departments should
consider redesigning their undergraduate curricula to allow students from 2-year
colleges and those who change their major to chemical engineering to complete their
degree without extending their time to degree, and provide the support structures
necessary to ensure the retention and success of transfer students.

Recommendation 9-4: To increase the recruitment of students from historically ex-


cluded communities into graduate programs, chemical engineering departments
should consider revising their admissions criteria to remove barriers faced by, for
example, students who attended less prestigious universities or did not participate in
undergraduate research. To provide more opportunities for women and Black, In-
digenous, and People of Color (BIPOC) individuals, departments should welcome
students with degrees in related disciplines and consider additions to their graduate
curricula that present the core components of the undergraduate curriculum tai-
lored for postgraduate scientists and engineers.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Summary 11

Recommendation 9-5: A consortium of universities, together with the American In-


stitute of Chemical Engineers, should create incentives and practices for building
and sharing curated chemical engineering content for use across universities and
industry. Such sharing could reduce costs and advance broad access to high-quality
content intended both for students and for professional engineers intending to fur-
ther their education or change industries later in their careers.

Recommendation 9-6: Universities, industry, federal funding agencies, and profes-


sional societies should jointly develop and convene a summit to bring together exist-
ing practices across the ecosystem of stakeholders in chemical engineering profes-
sional development. Such a summit would explore the needs, barriers, and
opportunities around creating a technology-enabled learning and innovation infra-
structure for chemical engineering, extending from university education through to
the workplace.

INTERNATIONAL LEADERSHIP

America’s scholarly leadership in chemical engineering with respect to both the


quantity of research, as measured by numbers of publications, and the quality of research,
as measured by citation impact, has decreased significantly in the past 15 years, losing
ground to international competitors, particularly China. The United States is in a leader-
ship position in some areas of chemical engineering technology, but lags in many niches
compared with various other countries.
The increase in research output from China is a result of large investments in a
range of technology areas, many of which are either central or highly relevant to chemical
engineering. Similar levels of investment in the U.S. research enterprise are imperative.
This report outlines the numerous opportunities for chemical engineers to contribute in
the areas of energy; water, food, and air; health and medicine; manufacturing; materials
research; and tools development. Without a sustained investment by federal research
agencies across these areas, as recommended above, it will be impossible for the United
States to maintain its leadership position. At the same time, U.S. chemical engineering
will be strengthened through increased coordination and collaboration across disciplines,
sectors, and political boundaries. Almost all of the areas of research discussed in this re-
port are multidisciplinary in nature and will require close collaboration between research-
ers in academic and government laboratories and industry practitioners to develop appli-
cations that are economically viable and scalable. Such cross-sector collaborations, as
recommended throughout the report, will have additional benefits for graduate student
education and faculty member development while also satisfying the need of industry to
achieve rapid results.

Recommendation 10-1: Across all areas of chemical engineering, in addition to ad-


vancing fundamental understanding, research investments should be set aside for
support of interdisciplinary, cross-sector, and international collaborations in the

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

12 New Directions for Chemical Engineering

areas of energy; water, food, and air; health and medicine; manufacturing; materi-
als research; tools development; and beyond, with the goal of connecting U.S. re-
search to points of strength in other countries.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

1
Introduction

In 1988, the National Academies of Sciences, Engineering, and Medicine laid out
an important vision for the field of chemical engineering in the report Frontiers in
Chemical Engineering: Research Needs and Opportunities, also known as the “Amundson
Report” (NRC, 1988). The report outlined a roadmap for making promising research op-
portunities a reality, and highlighted the remarkable potential of the profession to affect
many aspects of American life and promote the scientific and industrial leadership of the
United States. The report is widely recognized as having been a key driver for many ad-
vances in chemical engineering over the past 30 years. At the same time, tremendous
changes have occurred in chemical engineering, in the scientific enterprise more gener-
ally, in technology, and in the relationship between science and the public. These changes
have affected views on research priorities, education, and the practice of chemical engi-
neering and will continue to do so.

PURPOSE OF THIS REPORT

At a 2016 American Institute of Chemical Engineers (AIChE) roundtable, leaders


from the chemical engineering profession reached a major conclusion: the field of chem-
ical engineering needs a new vision for the 21st century. Participants at that meeting, in-
cluding both current and former AIChE presidents and multiple members of the National
Academy of Engineering, underscored the transformative and lasting impact of the
Amundson Report, unanimously supporting the need to update it. Perhaps more important
than identifying this need, the community provided support for such a study: in addition
to federal sponsors, more than 45 academic departments, private companies, and profes-
sional organizations offered financial contributions. This broad convergence of support
culminated in the formation of the National Academies’ Committee on Chemical Engi-
neering in the 21st Century: Challenges and Opportunities. The committee’s primary task
was to outline an ambitious vision for chemical engineering research, innovation, and
education that could guide the profession for the next 30 years (Box 1-1).

STUDY SCOPE AND APPROACH

The Committee was formed in late 2019 and completed its work over the course
of 18 months. The original work plan included six in-person meetings, five of which
would include information gathering sessions open to the public, to be held in locations
across the United States to maximize participation of the chemical engineering commu-
nity. Additional information-gathering webinars were to be held as needed to fulfill the
committee’s task. The committee’s participation in several professional society meetings
also was planned to increase community engagement and input opportunities.

13

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

14 New Directions for Chemical Engineering

BOX 1-1
Statement of Task

An ad hoc Committee will prepare a report that will articulate the status, challenges, and
promising opportunities for chemical engineering in the United States. In particular, the re-
port will:

 Describe major advances and changes in chemical engineering over the past three
decades, including the importance and contributions of the field to society; tech-
nical progress and major achievements; principal changes in the practice of R&D;
and economic and societal factors that have impacted the field.
 Address the future of chemical engineering over the next 10 to 30 years and offer
guidance to the chemical engineering community:
- Identify challenges and opportunities that chemical engineering faces now
and may face in the next 10-30 years, including the broader impacts that
chemical engineering can have on emerging technologies, national needs,
and the wider science and engineering enterprise.
- Identify a set of existing and new chemical engineering areas that offer
promising intellectual and investment opportunities and new directions for
the future, as well as areas that have major scientific gaps.
- Identify aspects of undergraduate and graduate chemical engineering ed-
ucation that will require changes needed to prepare students and workers
for the future landscape and diversity of the profession.
- Consider recent trends in chemical engineering in the United States rela-
tive to similar research that is taking place internationally. Based on those
trends, the report will recommend steps to enhance collaboration and co-
ordination of research and education for identified subfields of chemical
engineering.

The committee met in person in Washington, DC, in late February 2020 to iden-
tify information-gathering needs and begin its work. Shortly after that meeting, the
COVID-19 pandemic necessitated a major adjustment to the committee’s information-
gathering plans. In its early deliberative sessions, the committee identified the societal
and environmental areas in which chemical engineers have, or are likely to have in the
future, the largest impact. The committee members divided into subgroups based on those
areas (energy; water, food, and air; health and medicine; manufacturing and the circular
economy; materials research; tools development; and education) to gather information
and begin drafting what would become the main chapters of this report. Because of the
limitations of meeting virtually, the committee shifted to shorter but more frequent virtual
meetings. Information-gathering sessions were distributed across many meetings of both
the full committee and subgroups from summer 2020 through spring 2021, during which
time the committee met 42 times, with 27 of those meetings including a session that was
open to the public. To receive additional input from the community in lieu of regional in-
person meetings, the committee broadly distributed a questionnaire to chemical engineers
from all sectors at any stage in their career. The committee chair and National Academies

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Introduction 15

study director also led a town hall discussion at the 2019 AIChE fall meeting and partici-
pated in the April 2020 meeting of the AIChE Virtual Local Section to gather input from
the broader chemical engineering community. Finally, the committee conducted an ex-
tensive literature review.

AUDIENCES FOR THIS REPORT

The primary audience for this report is the broad chemical engineering commu-
nity, including researchers in academia and industry, educators, and students, as well as
federal and state decision makers and program leaders. It is anticipated that the report will
be used by

 students and faculty, to determine their research directions and design their
programs;
 industrial scientists and engineers, in creating their research and development
plans;
 universities and colleges, to improve undergraduate and graduate education
and diversify their student populations; and
 government program leaders and other research sponsors, to design and jus-
tify their programs.

Researchers in both academia and industry, depending on their specific area of


focus and expertise, will find the challenges and opportunities outlined in Chapters 3
through 8 of particular interest. Program managers at federal funding agencies will also
find these chapters useful, as well as the final chapter on maintaining international lead-
ership. Finally, faculty and leadership at colleges and universities, as well as professional
organizations, will find strategies for improving undergraduate, graduate, and lifelong
learning programs in Chapter 9. Chemical engineering departments can also look to Chap-
ter 9 for ways to make their programs more accessible while maintaining the rigor that
has made chemical engineering so successful as a discipline and a profession.

REPORT ORGANIZATION

To address the future of chemical engineering in the coming decades, the com-
mittee was tasked with identifying challenges and opportunities, as well as existing and
new areas for intellectual and investment opportunities and scientific gaps. The committee
chose to treat these elements collectively, as challenges and scientific gaps present excit-
ing opportunities for both intellectual investment by academic and industrial researchers
and funding investment by federal funding agencies and industrial research programs.
To provide a framework for the discussion in this report, the committee examined
the role of chemical engineering in addressing the key challenges facing society. Several
organizations have outlined grand challenges, but the committee chose to organize this
report around the areas noted above: energy and the energy transition; water, food, and
air; health and medicine; manufacturing and the circular economy; and materials.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

16 New Directions for Chemical Engineering

Chapter 2 provides a brief history of chemical engineering as both a discipline


and a profession, including key advances in training. This chapter focuses on some of the
most important contributions of chemical engineering, as well as changes in the field in
the practice of research and development (e.g., a shift from multiple–principal investigator
[MPI] projects and reduced corporate investment in research and development) and the
societal factors (e.g., the internet and global climate change) that have affected the field.
Major advances and changes in chemical engineering over the past three decades
in specific societal and environmental challenge areas are detailed in Chapters 3 through
7. The committee envisions a future for the field of chemical engineering that is more
collaborative with other disciplines and across sectors. Specific opportunities for collab-
oration are described throughout these chapters.
Chapter 3 presents opportunities for chemical engineers to contribute to decar-
bonization of current energy systems, describing the key research needs across the energy
value chain, from sources to various end uses. Energy, water, and food are highly inter-
connected, and solutions in this complex system need to be both environmentally sustain-
able and economically viable. Chemical engineers have historically played a central role
in the energy sector; their contributions in the space of water and food, as well as air
quality, have been important but are growing, as described in Chapter 4. The development
of disease treatments is a multidisciplinary enterprise, and Chapter 5 describes how chem-
ical engineers can contribute to many aspects of medicine by applying systems biology to
physiology, the discovery and development of molecules and materials, and process de-
velopment and scale-up. Chapter 6 explores opportunities for chemical engineers to im-
prove the sustainability of manufacturing by advancing the use of flexible feedstocks,
process intensification and distributed manufacturing, and the transition from a linear to
a circular economy. Specific materials applications are discussed throughout the report,
but Chapter 7 describes the role of chemical engineers in discovery science and the de-
velopment of new materials and materials processes, from the molecular to the macro-
scopic scale. Chapter 8 describes tools and techniques with the potential to enable future
advances across all of the previously discussed challenge areas.
Chapter 9 focuses on education and the need to maintain those aspects of the
curriculum that have made chemical engineering successful while allowing for innovation
to ensure that the next generation of chemical engineers is more demographically diverse
and receives training that ensures its ability to address the challenges facing society. The
report concludes with Chapter 10, focused on international leadership. Appendix A lists
acronyms used in this report; Appendix B provides journal titles used for the discussion
of international leadership in Chapter 10; Appendix C summarizes the results of a ques-
tionnaire distributed to the chemical engineering community; and acknowledgments and
biographical sketches of the committee and staff can be found in Appendixes D and E,
respectively.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

2
Chemical Engineering Today

Through discovery, design, creation, and transformation, chemical engineering is the


engineering of systems at scales ranging from the molecular to the macroscopic that
integrate chemical, physical, and biological elements in order to develop processes and
produce materials and products for the benefit of society.

The work of chemical engineers has transformed societies and individual lives
around the world, particularly in the United States. Chemical transformations are at the
heart of the technologies that enable modern society (Box 2-1). Without synthetic ferti-
lizers made with chemical engineering processes, Norman Borlaug could not have led a
Green Revolution to feed the world. Without the invention of catalysts due to the leader-
ship and vision of Karl Ziegler, a chemist, and Giulio Natta, the first chemical engineer
to win a Nobel Prize, polyolefins would not exist, and the myriad benefits of plastics
would not have been realized. Without the invention of tough, stable polymers by chem-
ical engineers at the DuPont Company, including Roy Plunkett and Stephanie Kwolek,
Teflon® and Kevlar® would not have been developed, nor would any of the commercial
and medical devices made from those polymers. Without the contributions of chemical
engineers such as Andy Grove at Intel and numerous others, the silicon chips, glass ma-
terials, and plastics that make up today’s ubiquitous electronic devices would not have
been created. Without the leadership and vision of Nobel Prize winner Frances Arnold,
who built upon the discoveries of numerous biologists, biochemists, and chemical engi-
neers, the tools of directed evolution would not have emerged. Without an army of chem-
ical engineers, there would be no oil and gas industry to power the world. And without
chemical engineers, a robust pharmaceutical industry would not have been able to dis-
cover and produce the therapies and vaccines needed for a long and healthy life. Figure
2-1 highlights some of the areas in which chemical engineers have made major contribu-
tions.
At the same time, however, chemical engineering is also responsible for unin-
tended consequences, such as those resulting from the production of chemicals that will
persist in the environment indefinitely, greenhouse gas emissions that contribute to cli-
mate change, plastic materials that accumulate in landfills and the oceans, and the chem-
icals of war that have inflicted long-term or permanent damage on humans and the envi-
ronment. Thus the field of chemical engineering today faces challenges and opportunities
not only to innovate for the future, but also to innovate in ways that repair the unintended
consequences of the past.

17

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

18 New Directions for Chemical Engineering

BOX 2-1
The Chemical Industry

The U.S. chemical industry supports more than 25 percent of the country’s gross domestic prod-
uct. It is the country’s number one exporter ($136 billion) with a $35 billion surplus, and directly
touches more than 96 percent of manufactured products. The industry is divided roughly into
basic chemicals, specialty chemicals, pharmaceuticals, agricultural chemicals, and consumer
products (DHS, 2019).*

Basic chemicals. These chemicals are manufactured in large volumes to uniform composition
specifications. The largest example is the transformation of hydrocarbon feedstocks or minerals
into commodity products. The price margins for these chemicals are small, but the large volumes
drive profits, and thus the sector is vulnerable to economic cycles. The sector is extremely en-
ergy intensive and has benefited from the shale gas boom in the United States, which provides
cost advantages for U.S. feedstocks.

Specialty chemicals. These chemicals are manufactured in smaller volumes and are generally
designed or engineered for a specific (and often demanding) application. Examples include fla-
vors and fragrances, inks, catalysts, and electronic materials. The price per unit volume of some
of these products is extremely high, and the manufacturing processes can be challenging because
extremely high purities are required.

Pharmaceuticals. Therapeutic molecules for the prevention or treatment of disease are the most
value-added chemicals. This sector requires a long time for drug development, manufacture, and
testing, and products carry a premium price. As in the case of some specialty chemicals, the
success of the products is completely dependent on the purity and function of the active mole-
cules, and large profits can be realized from small amounts of material. Fewer than 10 percent
of drugs that make it to clinical trials are approved by the U.S. Food and Drug Administration—
another contributor to the high costs of these products (BIO, 2011).

Agricultural chemicals. Chemical agents, such as fungicides, insecticides, and chemicals that
influence the growth cycle of plants are commonly used in agriculture along with various kinds
of fertilizers. Other examples of agricultural chemicals are herbicides, rodenticides, and insect
attractants or repellents. In all cases, the animal and human toxicity of these chemicals is a per-
sistent concern. While these products are produced in smaller volumes than basic chemicals,
annual shipments in 2019 were worth $41.1 billion in the United States (DHS, 2019).

Consumer products. Consumer products made from hydrocarbons or minerals are ubiquitous.
The plastics economy has revolutionized the manufacturing of everything from cars to clothing,
and concerns about its environmental impact are now driving the urgency of finding ways to
reuse and recycle these materials. Similarly, concern about manufacturing of the largest-volume
inorganic chemical—cement—is growing because the process accounts for about 8 percent of
global greenhouse gas emissions (Andrew, 2018).
*
The National Academies of Sciences, Engineering, and Medicine’s Committee on Enhancing
the U.S. Chemical Economy through Investments in Fundamental Research in the Chemical
Sciences is currently examining the role of the chemical industry in the U.S economy.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Chemical Engineering Today 19

FIGURE 2-1 Schematic illustrating the broad impact of the chemical sector on several aspects of
American life. Examples of basic chemicals, specialty chemicals, agricultural chemicals, pharma-
ceuticals, and consumer products are shown. SOURCE: DHS (2019).

Chemical engineering as a discipline brings together the three fundamental sci-


ences of chemistry, physics, and biology, as well as mathematics. Chemical engineers are
agnostic to the material used or the particular application; they work with all phases of
matter—vapor, liquid, supercritical, solid, and plasma—and with multiphase mixtures
and at interfaces. They work at and across all length scales, ranging from molecules to
medicines to materials and even machines. Their work creatively transforms matter and
products into higher-value materials and products using the principles and tools of ther-
modynamics, transport, kinetics, process control, and process design. Chemical engineer-
ing is the only field of engineering that takes advantage of chemical transformations, usu-
ally followed by separation and purification, to add value to products.
Chemical reactions and reacting systems are central to many transformations. The
problems tackled by chemical engineers usually involve time dependence, transport phe-
nomena, and nonequilibrium phenomena and feature many variables. Chemical engineers
are trained to take a system-wide approach to solving problems, recognizing that a single
step is rarely adequate to complete the desired transformation. They realize that a system
of connected components needs to be optimized holistically to be most efficient and eco-
nomic. Importantly, chemical engineers appreciate the balance among physical perfor-
mance, fiscal profitability, and safety.
The education and training unique to chemical engineering produce intellectually
versatile engineers who are comfortable with mathematics and computers, skilled in ana-
lytical thinking, and adept at solving open-ended problems. They draw on an education
that has served them well in diverse areas of accomplishment. Accordingly, chemical en-
gineers are found in interdisciplinary teams in the agricultural, biological, biomedical,
chemical, environmental, food, health, energy, materials, petrochemical, nanotechnology,
pharmaceutical, and semiconductor industries, as well as in the consulting, communica-
tions, software, financial, and insurance sectors. They engage at all levels in these indus-

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

20 New Directions for Chemical Engineering

tries and sectors, from research and development, to analysis, administration, and leader-
ship, and they contribute to the engineering enterprise from scientific discovery, to appli-
cation to scale-up, to commercial deployment, to plant operations.

THE DISCIPLINE

Chemical engineering is both a discipline and a profession. Its academic discipli-


nary roots trace back to the 1880s. In 1880, the English entrepreneur George Davis pro-
posed the creation of a Society for Chemical Engineers, which resulted in the creation of
the Society of Chemical Industry; in 1886 and 1887, diploma curricula began at City and
Guilds College (now Imperial College) and at Glasgow and West of Scotland Technical
College, respectively; and in 1888, Lewis Mills Norton proposed the first course in chem-
ical engineering at the Massachusetts Institute of Technology (MIT, 2021). Chemical en-
gineering evolved as a profession from the roots of industrial and applied chemistry,
which in turn emerged from such ancient chemical processes as fermentation and leather
tanning. Traditionally, industrial and economic needs and academic research have been
tightly coupled, with many faculty members being engaged in consulting or joint research
work with industry. To paraphrase L. E. Scriven, the academic tree takes its nourishment
from the soil of practical challenges and periodically returns its many leaves of results
and trained graduates to that soil (University of Minnesota, 2003).
This discipline is the only one in engineering with a central focus on molecules
and their transformations—that is, chemistry. For much of its history, chemistry and phys-
ics were the core sciences of chemical engineering (along with mathematics). Because of
its fundamental engagement with chemistry, it is natural for chemical engineering to be
the principal home for the engineering of biological systems of all scales. The connection
between chemical engineering and biology has a long history, and their interactions grew
exponentially with the advent of methods for genetic engineering and the advancement of
molecular biology from the late 20th century onward (Box 2-2). Indeed, numerous fields
(e.g., biochemical engineering, metabolic engineering, tissue engineering, synthetic biol-
ogy) have developed as outgrowths of chemical engineering or with key contributions
from chemical engineers. The expansion of the influence of biology on the field has also
been so significant that many academic departments have changed their names to incor-
porate some version of “bio” in addition to “chemical” engineering.
Because chemical engineers deal with both molecules and the enormous indus-
trial plants that produce them, their work encompasses a large range of length and time
scales, from the nanometer scales of chemical bonds and reactions to the kilometer scales
of crude oil (petroleum) refineries, and from nanosecond chemical reactions to batch pro-
cesses that take hours. Few fields of science or technology deal with changes of more than
a dozen orders of magnitude in both length and time scales. The core curriculum of chem-
ical engineering has lasted more than a century because it focuses on the analysis of linked
processes regardless of scale.
The early history of chemical engineering has been well described (Colton, 1991;
Furter, 1983). The growth of the core curriculum is marked by several events, the first of

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Chemical Engineering Today 21

BOX 2-2
The Growth of Biologically Related Content in Chemical Engineering

The birth of biochemical engineering occurred in the mid-1940s, following the successful
application of antibiotics to treat infection during World War II and the realization that the
United States could use such techniques more broadly if appropriately scaled. Merck Chemical
Company was instrumental in designing improved aeration facilities, with consultation from
Richard Wilhelm at Princeton and Elmer Gaden, a graduate student at Columbia, who, as part
of his PhD thesis, examined the mass-transfer characteristics of fermentors. Merck, building on
Wilhelm’s work with fluidized beds, also improved the productivity of suspension cultures and
chiral separation by crystallization. Margaret Hutchinson Rousseau designed and scaled up the
first commercial penicillin production plant during World War II at Pfizer.
A major catalyst for growth followed the collaboration of Herbert Boyer and Robert Swan-
son to found Genentech in the early 1970s. Along with cocreators Stanley Cohen and Paul Berg,
Boyer developed the technology used to “cut and paste” DNA from any organism into another
in which it could be expressed. In 1977, Genentech succeeded in isolating the sequence for
human insulin, and by 1982 it had marketed the first recombinant protein therapeutic. This con-
tinues to be a growing industrial area, and it is estimated that by 2024, the global market in
protein therapeutics will be valued at $500–750 million annually (IndustryARC, 2021). Others
have launched major changes in the field. Raymond F. Baddour worked at the interface of bio-
technology and pharmaceuticals and cofounded Amgen in 1980. James E. Bailey was an Amer-
ican chemical engineer who played a major role in the development of metabolic engineering.
Daniel Wang contributed to many aspects of biochemical engineering, notably enzyme technol-
ogy and mammalian cell culture and bioreactors. And Frances Arnold pioneered the use of di-
rected evolution to create enzymes with improved or novel functions, work recognized by the
2018 Nobel Prize in Chemistry.

which featured George E. Davis in England in the 1880s. Davis was an entrepreneur in
chemical manufacturing who called for the creation of “chemical engineering” to address
the rampant pollution generated by chemical processes (Cohen, 1996). The first course
(Course X) developed by Lewis Mills Norton at MIT had its stops and starts, but by 1902,
William Walker had accepted the leadership of Course X while remaining in consulting
partnership with Arthur D. Little until 1905. Arthur D. Little, as chair of the visiting com-
mittee for chemical engineering at MIT, is credited with creating the concept of “unit
operations” in 1915 to describe the basic physical (and later, chemical) operations used to
transform raw materials into products (Flavell-While, 2011).
Throughout the 20th century, further academic advances in the curriculum took
place across the United States. At the University of Wisconsin, Olaf A. Hougen and Ken-
neth M. Watson wrote the three-volume Chemical Process Principles: Material and En-
ergy Balances (1943); Thermodynamics (1947b), and Kinetics and Catalysis (1947a).
Also at the University of Wisconsin, R. Byron Bird, Warren E. Stewart, and Edwin N.
Lightfoot published the paradigm-shifting textbook Transport Phenomena in 1960. That
book transformed chemical engineering by bringing a strong mathematical approach to
the unification of treatments of fluid mechanics and heat and mass transfer, both reinforc-
ing and explaining the connections made by Allan P. Colburn and Thomas H. Chilton at

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

22 New Directions for Chemical Engineering

the University of Delaware and the DuPont Company, respectively, in 1946 (Chilton and
Colburn J-factor analogy). At the University of Minnesota, Neal Amundson and Ruther-
ford Aris worked to develop the analytic and mathematical foundations for reaction engi-
neering. At the University of California, Berkeley, John M. Prausnitz developed methods
of molecular thermodynamics beginning in the 1950s and 1960s. Throughout and after
this period, the role of computing in chemical engineering control and design grew stead-
ily (Box 2-3). Thus by the mid-1960s, a canonical undergraduate program, rooted in a
foundation of fundamentals from chemistry, physics, and mathematics, was formed in a
way that is easily recognized today.

BOX 2-3
Control, Design, and Computer Applications in Chemical Engineering

The development of process control and process design coincided with the development of
the use of computers in chemical engineering, as these topics were then, and remain today,
heavily computational in nature. Much of the pioneering work in control was centered in indus-
try, with Page S. Buckley at DuPont and Joel O. Hougen at Monsanto (and later at The Univer-
sity of Texas at Austin) at the forefront. F. Greg Shinsky and Edgar H. Bristol at Foxboro also
contributed to the industrial practice of control from the perspective of hardware equipment
manufacturing. The academic pioneers included Rutherford Aris at the University of Minnesota
and Leon Lapidus at Princeton University, who contributed to the theory of optimal control in
chemical processes.
The earliest textbooks on the topic were written by Daniel D. Perlmutter at the University
of Pennsylvania (1975), Ernest F. Johnson at Princeton University (1967), and Donald R.
Coughanowr and Lowell Koppel at Purdue University (1965). The first chemical engineering
textbooks that highlighted computer applications were the numerical analysis books by Lapidus
(1962) and Brice Carnahan and colleagues (1969) at the University of Michigan. The University
of Michigan, motivated by the introduction of FORTRAN (Formula Translation), initiated a
project under the direction of Donald L. Katz to study the use of computers in engineering,
which resulted in a seminal 1966 report entitled Computers in Engineering Design Education.
Chemical engineering education in general, but more specifically the computationally heavy
process control and process design courses, owes much to the vision laid out by Katz in this
report. Pioneering work in chemical process simulation and design was carried out by Rodolphe
L. Motard and Ernest J. Henley at the University of Houston, who, along with J. D. “Bob” Seader
at the University of Utah, wrote the dominant textbook on the topic (1998). The preeminent
textbooks on computer applications in chemical engineering in general include those on process
model building, process analysis, and process simulation written by David M. Himmelblau and
Kenneth B. Bischoff (1968); on numerical methods by Brice Carnahan, H. A. Luther, and James
O. Wilkes (1969); on process optimization, by Gordon S. G. Beveridge and Robert S. Schechter
(1975); and on modeling and simulation, by Roger G. E. Franks (1972).

THE PROFESSION

Chemical engineering as an industrial profession has its roots in processes for


materials extraction and transformation. Éleuthère Irénée du Pont de Nemours founded
his eponymous company in 1802 to make gunpowder on the banks of the Brandywine

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Chemical Engineering Today 23

River in Wilmington, Delaware, at a site with a dam, a millrace, and saltpeter at hand.
John D. Rockefeller became the world’s first oil baron when he united several companies
into the Standard Oil Company in 1870, exploiting wells in Ohio, Pennsylvania, and even-
tually elsewhere. Standard Oil quickly came to control about 90 percent of America’s
refining capacity (see Ohio History Central, 2021). Herbert Henry Dow founded his com-
pany in 1897 to extract bromine from underground brine in Midland, Michigan. Over
time, each of these companies came to appreciate and need chemical engineering exper-
tise and scientific advances to grow and diversify.
By the 1970s, the leading chemical engineering research efforts in U.S. industry
were at Esso (later Exxon) and Mobil laboratories, DuPont Central Research and its Ex-
perimental Station, and Dow Central Research. Space does not allow enumeration of all
of the advances in materials and processes that emerged from those laboratories and de-
velopment efforts, or other important advances in research and product development from
the laboratories at 3M, Shell, Universal Oil Products (later Honeywell UOP), Aramco,
Standard Oil Company, Bell Labs, and elsewhere, but they all contributed in a significant
and largely defining way to the quality of life enjoyed in the United States. Parallel de-
velopments also played out in chemical and other companies focused on materials. In
1953, for example, Lexan® polycarbonate was invented at the General Electric (GE) Com-
pany. This invention led to the creation of the GE Plastics Business, which grew to be a
global business based on major contributions of chemical engineers (Plastics Hall of
Fame, 2021).
Chemical engineering has also played an important role in electronic materials, a
role amplified by the nearly ubiquitous role of electronic devices in today’s society. The
expansive proliferation and adoption of electronics globally has driven an explosion of
uses for electronics well beyond the traditional uses of personal computers and cell
phones. Today, the introduction of 5G, the internet of things (IoT), cloud computing, and
autonomous driving is having a direct impact on how electronics are used. Looking for-
ward, the further adoption of machine learning and artificial intelligence to expand com-
putational capability and demands for electronic devices are at an inflection point. While
these trends had been progressing for some time, they have been accelerated by the
COVID-19 pandemic, which has made technical advancement more urgent than ever—
similar to the way it was in early days of the internet and the introduction of personal
computers.

TIMES OF CHANGE

The world has changed and continues to change at an increasingly rapid rate.
Technology is transforming the way everything is done, disrupting established practices
and organizations. The range of challenges facing engineers is evolving, and the chal-
lenges are becoming more difficult. Engineering is about solving problems, so it is natural
to expect that as one problem is solved, another emerges or grows in importance. The goal
of education in chemical engineering is to equip graduates with the intellectual tools
needed to solve problems and the ability to adapt those tools as new problems emerge.
Information used to be valuable in its own right, but the internet has made information

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

24 New Directions for Chemical Engineering

essentially free. What is now more valuable than ever is the knowledge needed to curate
data and synthesize new and novel solutions, and to do so more rapidly and at lower cost
than competitors. In a real sense, an undergraduate chemical engineering degree is a plat-
form upon which its holder can build, with a research degree, a professional degree, and/or
a career filled with formal and informal learning.
There is of course nothing new about change. The last time the National Acade-
mies surveyed the challenges and opportunities for chemical engineering—in the 1980s
(Frontiers in Chemical Engineering: Research Needs and Opportunities, better known as
the “Amundson Report” [NRC, 1988])—the conclusions reached suggested an approach
to developing new technologies and maintaining leadership in established ones. A Cen-
tennial Symposium of Chemical Engineering (Wei, 1991) laid out the challenges facing
the field at that time, and included a rich discussion of the balance of revolution and evo-
lution in chemical engineering. Comments at the time about the tension between teaching
the established core of such topics as transport, thermodynamics, kinetics, and design and
making room for education relevant to (then) new areas such as biotechnology and mate-
rials science resonate today in this report. Nonetheless, the rate of change driven by an
interconnected world is without doubt faster today than it was yesterday, and will be even
faster tomorrow.
As new advances combine to produce even newer advances, the pace accelerates.
At the time of the Amundson Report in the 1980s, personal computing was just becoming
common, under 1 million cell phones were in use in the United States (Tesar, 1996), the
internet was used by a cognizant few, and Microsoft had just gone public. The top ten
public companies in terms of revenue included Exxon (1, at $91 billion), Mobil (3), Tex-
aco (4), DuPont (7), and Amoco (10) (Fortune, 1985). The undergraduate chemical engi-
neering curriculum focused on the core fundamentals of thermodynamics, transport phe-
nomena, reactors and kinetics, process control, and process design. In 2021, as this report
was being written, the internets of information and things were affecting all aspects of
life, 14 billion cell phones were in use worldwide (Statista, 2021), the human genome was
known and could be edited, it was possible to “see” individual atoms, and the use of arti-
ficial intelligence and deep machine learning was growing rapidly. The top ten U.S. public
companies in terms of revenue in 2020 were Walmart (1, $514 billion), ExxonMobil (2),
Apple (3), Berkshire Hathaway (4), Amazon (5), United Health Group (6), McKesson (7),
CVS Health (8), AT&T (9), and AmerisourceBergen (10). After ExxonMobil, the next
largest oil and gas company was Chevron, at number 15. Dow Chemicals, the largest
materials science and chemical company in the United States, was 78th with revenues of
$43 billion, less than 10 percent of Walmart’s (Fortune, 2020).
The undergraduate chemical engineering curriculum is still focused on the core
fundamentals of thermodynamics, transport phenomena, reactors and kinetics, process
control, and process design. The chemical industry, however, is a less important part of
the overall economy than it once was, although the world can no more do without energy,
materials, food, and water now than it could then. The top companies more than 100 years
ago were all recognized industrial brands; today, technology companies have the largest
market capitalizations. According to the Bureau of Labor Statistics (BLS), overall em-
ployment in chemical engineering dropped from 43,270 in 1997 to 30,120 in 2019, from

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Chemical Engineering Today 25

almost 0.04 percent to shy of 0.02 percent of the national employment total (BLS, 2021).
However, many people educated in chemical engineering work in areas not identified as
such by BLS.
Over the last decade in the United States, the number of bachelor’s and master’s
degrees awarded in chemical engineering more than doubled, a rate outpacing the 60–80
percent growth of engineering and STEM (science, technology, engineering, and mathe-
matics) bachelor’s and master’s degrees. The number of doctorates awarded grew more
modestly, with engineering growing most rapidly (Table 2-1).
Technology has also transformed the way people work. The linear industrial
model of changing raw materials to products has shifted to a global interconnected plat-
form model. That change, the increasing time-rate of change in society and business in
general, and the substantial diminution of the economic significance of the chemical and
oil and gas (but not the pharmaceutical) industries in the U.S. economy all have created
substantial challenges and opportunities for chemical engineering.
The pace of change is now accelerating as a result of global connectivity, massive
amounts of available data, artificial intelligence, sensors, robotics, and more. The back-
ground and training of chemical engineers are well suited to today’s rapidly changing
world, and many of them have found their way into companies leading change. For
example, as the new era of artificial intelligence and machine learning advances, chemical
engineers are well equipped to move into these fields because of their strong background
in reaction (network) engineering and control, along with an understanding of how to
deal with complex interconnected processes, in addition to their fluency in math and
computing.
These longer-term and probably irreversible societal and business changes are
augmented by a critical need to mitigate climate change resulting from greenhouse gas
emissions. The ultimate measure of success in addressing climate change will be reducing
greenhouse gas concentrations in the atmosphere while still delivering the energy that
society needs. The necessary system solutions align with the core chemical engineering
practices of designing, building, and extending major manufacturing assets. Chemical
process engineers have a long history of building safe, resource-efficient processes at the
lowest capital and operating cost.
Along with addressing climate change and the energy transition are parallel
needs—to reduce raw material usage and increase recycling to create a more circular
economy; deal with the need to generate and distribute food worldwide while conserving
water and other resources; and create and scale the manufacturing and distribution of new
medicines and therapeutics. Each of these needs is discussed more fully in later chapters.
There are other needs as well. Complicating efforts to address all of these needs are inev-
itable shorter-term geopolitical and environmental issues related to the price and availa-
bility of energy and feedstocks, as well as existing and emerging security threats and the
viability of the installed asset base (Box 2-4). In these times of change, chemical engineers
also play a central role in advances in many new and important areas, including person-
alized medicine, the infrastructure needed to produce vaccines for a pandemic, the rapid
conversion to an economy based on nonfossil fuels, the need for new materials, and many
more.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering
TABLE 2-1 Bachelor’s, Master’s, and PhD Degrees Awarded in the United States between 2008 and 2019 in Chemical Engineering (ChE),

26
All Engineering (Eng), and All STEM (Science, Technology, Engineering, and Mathematics) Fields
Field 2008 2009 2010 2011 2012 2013 2014 2015 2016 2017 2018 2019
ChE 4919 5137 5838 6416 7176 7678 8202 9070 10032 11021 11653 11148
Bachelors

Eng 70232 70991 74778 78502 83636 88201 94386 100316 109373 118379 124794 114818
Copyright National Academy of Sciences. All rights reserved.

STEM 592,801 610,216 639,822 683,865 736,982 781,937 818,434 850,168 877,786 905,509 934,776 *
ChE 937 996 1051 1284 1395 1453 1521 1629 1701 1801 1921 2003
Masters

Eng 33513 36909 38029 41751 43765 44037 46029 49855 55907 57754 56756 62682
STEM 164,007 175,895 184,097 199,386 215,137 225,261 235,368 249,762 274,084 290,037 298,157 *
ChE 873 807 822 822 839 824 972 1002 920 930 981 1092
PhD

Eng 7860 7637 7575 8024 8452 8998 9585 9875 9455 9771 10179 12372
STEM 34,717 35,313 34,997 36,332 37,846 39,031 40,633 41,178 41,234 41,294 42,227 *
Data from ASEE (2020) and NCSES (2021).
New Directions for Chemical Engineering

Chemical Engineering Today 27

BOX 2-4
Current and Future Threats: Cybersecurity

As chemical and biological process plants and other infrastructure have increasingly be-
come connected to and become part of the internet, the threat of cyberattacks that can cripple a
network, cause physical and/or life-threatening damage, or enable collection of a large ransom
has increased dramatically. Most companies take cybersecurity across all business sectors very
seriously. Many have large units devoted to cybersecurity across manufacturing facilities, re-
search and development, and contractor monitoring. Most of these teams are made up of com-
puter scientists, but they work closely with chemical engineers who are designing new processes
or control systems. In this context, chemical engineers will not need to develop new cybersecu-
rity tools, but they will need to be able to communicate effectively with computer scientists and
other cybersecurity professionals. Universities need to make students aware of the importance
of cybersecurity, as well as development of the skills needed to work with a diverse group of
colleagues with deep expertise in various areas.

EDUCATIONAL CHALLENGES AND OPPORTUNITIES

While the core chemical engineering curriculum has apparently changed little
over the preceding decades, the concepts are now taught in more modern ways. Initial
unemployment for chemical engineering graduates is low and salaries remain high, alt-
hough not at the highest levels relative to other fields, as was the case in the past. Thus,
despite a course catalog that appears at first glance to be stuck in the past, the curriculum
has been shown to yield graduates with the skills needed to adapt and succeed in the
workplace. The hallmark of this continuing success appears to be the ability of chemical
engineering graduates to think quantitatively, draw on data to guide the development of
predictive models that can be expressed mathematically, and integrate pieces into a well-
designed coherent system. In other words, “[the curriculum]…has endured not because it
is frozen but because it has adapted dynamically to new ideas, emphases, challenges, and
opportunities” (Luo et al., 2015). The undergraduate curriculum and chemical engineering
education generally are discussed in more detail in Chapter 9.
New models of collaboration and integration are also emerging. Education has
historically been designed as a linear flow from K-12 to college/university and then to a
career. Today, global connectivity has led to new integrated learning and innovation prac-
tices. New education offerings are emerging from such online providers as EdX,
Coursera, and Udacity. Online certificates and skills programs are finding wide applica-
tion as the pace of change demands the constant refreshing of fundamental knowledge.
The opportunity to build new shared content for use across universities is appealing, but
would need careful consideration in light of the existing funding and reward systems
within universities. Any such program needs to be available for lifelong acquisition and
polishing of skills. Finally, assembling best-in-class content online for use as part of a
university’s offerings can make content available to smaller or underresourced institu-
tions, which in turn could open up new paths for engaging different and more diverse
talent pools with the concepts of chemical engineering.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

28 New Directions for Chemical Engineering

For some, the undergraduate curriculum is a preamble to graduate studies. Grad-


uate education is where the striking differences between the subjects undergraduates learn
and the problems on which academics conduct research become clear. For example, un-
dergraduates learn about small-molecule thermodynamics and some separation processes
that have generally been highly specific to the oil and gas industry, although many bio-
separation processes are now included in multiple courses. The oil and gas examples are
obvious for historical reasons, but a vanishingly small amount of academic research is
now occurring on those subjects. Thus, the real test for graduate students is their ability
to adapt their critical thinking and analytic skills to problems in fields for which they
probably lack both understanding of the vocabulary of the field and at least nontrivial
basic science training.
This disconnect between the intellectual content of undergraduate and graduate
work reflects the collective decisions of mainly federal but also other funding agencies
that in large part support the work of graduate students. There is little federal support for
work on basic thermodynamics or transport phenomena that an undergraduate would rec-
ognize. The disconnect is clear from even a casual look at the titles of grants funded most
recently by the National Science Foundation’s (NSF’s) Division of Chemical, Bioengi-
neering, Environmental, and Transport Systems, where many academic chemical engi-
neers find at least some support for their research. Key phrases for funded proposals in-
clude “dynamic covalent junctions on block copolymer and network self-assembly,”
“sustainably derived high-performance nanofiltration membranes,” “ultrahigh-resolution
magnetic resonance spectroscopic imaging for label-free molecular imaging,” and “pho-
tonic resonator hybrids.” Funding of chemical engineers by the National Institutes of
Health is of course even further disconnected from the traditional focus of the undergrad-
uate curriculum.
Federal funding models are also driving a change in the way research is done.
Through the 1990s, most research proposals in chemical engineering were written by in-
dividual principal investigators and generally were based on elements of intellectual cu-
riosity balanced perhaps with a desire to solve a practical problem. Over the past 20 years,
NSF in particular has increased the number of multi-investigator awards relative to
single–principal investigator awards. Multiple–principal investigator proposals are often
submitted in response to a call for proposals in a particular area (bioseparations or cyber-
security, for example). The data (Figure 2-2) show that this shift has been slow but steady.
The impact on academic research has been positive in the sense that there are more inter-
disciplinary teams, and students are more likely to have multiple advisors or mentors, both
of which improve educational outcomes. However, this shift also has led to a decline in
funding for basic research by individuals on problems they find interesting. The new
model supports innovation, frequently done by large teams, but not necessarily invention,
which often comes from individuals (e.g., Wu et al., 2019). In addition to this shift toward
large teams, the committee’s sense is that funding opportunities are more prescriptive, or
targeted, and less open-ended than in the past. In this context, funding agencies will need
to consider the appropriate balance to meet their respective missions.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Chemical Engineering Today 29

FIGURE 2-2 Number of new research projects funded by the National Science Foundation with a
single principal investigator (SPI) (light blue; left) or multiple principal investigators (MPI) (dark
blue; right) over the past 10 years. The ratio of MPI to SPI awards (solid black line with dashed
black trend line) has increased. Data from NSF (2020).

This challenge is not unique to chemical engineering. Tension between basic and
applied research characterizes most fields of science and engineering. Inventions cannot
be planned and do not arrive on a schedule. Basic research is messy, nonlinear, and ex-
pensive. New knowledge relies on serendipity and the preparation of a fertile mind, and
not all individuals have the ability to invent. The outcomes of basic research can funda-
mentally change the course of history, but for that to happen, a period of innovation after
invention is usually required. Such innovation is often achieved through applied research
or engineering.
On the industrial research front, the linear model of research, development, and
commercialization has also been disrupted in the years since the Amundson Report was
released. In the past, major companies deployed research and development (R&D) centers
in support of their long-term growth objectives. Projects would then be transitioned to the
appropriate business areas for commercialization, which often led to major new products
and services. The R&D centers also served as a source of technical talent for the business
areas because the research innovators would often transition to a business unit along with
their technology developments. R&D centers maintained close connections to universities
and academic programs as well. The overall system was a path from basic discovery at
the university out to the market.
Priorities began to shift in the 1980s. Companies need to create new products or
processes to maintain profitability. Large companies in the chemical and oil and gas in-
dustries have shifted their research activities toward supporting existing businesses. Com-
panies have set shorter-term goals for corporate research, and their support for university
research now focuses more on basic science and has grown only slightly since the 1980s

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

30 New Directions for Chemical Engineering

(NSB, 2020). These changes have further widened the “valley of death” between discov-
ery and commercialization. To fill that gap, startups, innovation incubators, and many
other intermediate models have emerged in and near universities, but the bidirectional
connection of research and the marketplace is still challenged. The exception is the phar-
maceutical industry, where fundamental drug discovery is still carried out. A comparison
of R&D expenditures as a function of total revenue or total chemical sales (Figure 2-3)
shows the marked difference between the pharmaceutical industry and the chemical and
petroleum industry, respectively.
In a relatively recent development, companies have been expanding or replacing
parts of their R&D effort with open innovation models in which they are seeking solutions
from outside their R&D divisions or even their companies. Companies are also forming
alliances with other companies, universities, and/or national laboratories to solve complex
problems, such as the “end of plastic waste” (the Alliance to End Plastic Waste includes
more than 40 companies that are pooling funding and expertise to address the problem of
plastic waste).
Education and research have historically been viewed as separate activities. To-
day, there is an opportunity to design new models whereby learning and innovation are
connected through new forms of public–private partnerships. Many opportunities exist to
build a connected model in which students can take the latest ideas and technology to the
marketplace, have experience with a business operation, and return with a perspective on
the needs of the market to inspire new research initiatives. Many of the component steps
for such a model exist, but new cross-sector collaborations to support and accelerate the
growth of this model are needed.

FIGURE 2-3 Research and development (R&D) expenditures in 2020 by the top 10 U.S.-based
pharmaceutical (light blue circles) and chemical (dark blue triangles) companies (for which data
were available), each compared with their revenue or chemical sales, respectively. Data from Buntz
(2021); C&EN (2021); Macrotrends (2021a,b).

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Chemical Engineering Today 31

GROWTH OF INTERDISCIPLINARY WORK

The growth of interdisciplinary research reflects the increasing complexity of the


problems to be solved and the growing sophistication of the tools needed to solve them.
Nonetheless, such research also poses a danger. By definition, one cannot have interdis-
ciplinary work without disciplines. The field of chemical engineering needs to recognize
that its graduates are valued for the disciplinary skills they can bring to bear in working
with others on a problem. Thus it is important to continue to educate students in the basic
skills of chemical engineering, albeit with examples less reflective of an olefin-based
business. At the same time, the field needs to be open to the influx of faculty members
and practitioners who have not had a traditional chemical engineering education. Chemi-
cal engineering has benefited enormously from the influx of mathematicians, physical and
other chemists, physicists and materials scientists, biologists, and others to its faculties
over the years, a phenomenon expected to continue. One revealing example is that in the
United States, chemical engineering is home to most programs in polymer science, a field
largely founded in but not embraced by chemistry. This is no time for stasis, but instead
a time for the field to expand and grow at its current frontiers while remaining true to the
core that defines it. As societal challenges become increasingly complex, science and en-
gineering solutions will necessarily come from connections across disciplines, and the
boundaries between disciplines will continue to blur. To contribute to solutions for soci-
etal challenges in the coming decades, chemical engineers will need to become increas-
ingly comfortable working across disciplines and as members of interdisciplinary teams.
Indeed, this report highlights many areas in which chemical engineers will benefit from
interdisciplinary collaborations.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

3
Decarbonization of Energy Systems

 Addressing the existential threat of climate change will require decarbonization of


current energy systems, a challenge rendered all the more difficult by the complex-
ity and magnitude of the energy landscape and the resultant inability of any single
energy carrier to meet the energy demands of all sectors in the foreseeable future.
 The field of chemical engineering continues to make important contributions to the
scalability, delivery, systems integration, and optimization of the mix of energy
carriers that will meet energy needs across different regions and sectors of society
with lower carbon emissions and costs.
 Chemical engineers will enable technological advances at every point in the energy
value chain, from sources to end uses, and bring to bear the systems-level thinking
necessary to balance the economic and environmental trade-offs that will be nec-
essary to transition to a low-carbon energy system.

Energy is a basic human need that is also essential for economic growth. The
global energy demand today is about 155,000 TWh, 82 percent of which is currently sup-
plied by fossil fuels (32 percent petroleum, 23 percent natural gas, and 27 percent coal)
and the rest by nuclear (5 percent); wind and solar (3 percent); and other renewable energy
sources, including hydroelectric, geothermal, and biomass (10 percent) (BP, 2019; IEA,
2021e). In 2018, the primary energy consumption by the various sectors was electricity
generation (38 percent), transportation (28 percent), industrial processes (23 percent), and
residential and commercial spaces (12 percent) (EIA, 2021h). The global energy supply
mix will be altered substantially by the decarbonization efforts of various sectors, partic-
ularly in electricity generation, and by the greater adoption of electric vehicles (EVs) for
light-duty transportation. Refineries are optimized for the production of gasoline or diesel,
and substantial work and investment have gone into the use of biofuels for transportation.
Widespread use of EVs will disrupt both the fossil and biofuels industries.
Chemical engineering has played an essential role in meeting society’s demands
for economical and energy-efficient conversion of natural resources into liquid and gase-
ous energy carriers while addressing environmental challenges associated with their pro-
duction and use. Between now and 2050, the world population is expected to grow from
7.5 billion to well over 9 billion (OECD, 2012; UN, 2017), and increasingly prosperous
populations will demand more energy; by 2050, the global demand for energy is forecast
to increase by almost 50 percent (EIA, 2020b). Chemical engineering will continue to
enable the equitable delivery of increasing amounts of reliable and affordable energy
while supporting efforts to address the existential threat of global climate change (e.g.,
AIChE, 2020). Doing so will require the development and scale-up of renewable energy

32

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Decarbonization of Energy Systems 33

sources and carbon sequestration and utilization at an unprecedented rate and scale, as
well as consideration of trade-offs in such areas as water consumption, cost, and environ-
mental justice.
This chapter describes the impetus for prioritizing decarbonization of energy sys-
tems and the important role chemical engineers will have in advancing technologies that
minimize the climate impact of the energy sector. The chapter is organized from sources
to end uses. Opportunities for chemical engineers are explored in energy sources; energy
carrier production; energy storage; energy conversion and efficiency; and carbon capture,
use, and storage.

THE NEED FOR DECARBONIZATION

The international climate science community has established a link between


global greenhouse gas (GHG) emissions and human actions and energy usage. The con-
centration of CO2 in the atmosphere has tracked closely its rate of anthropogenic produc-
tion since the start of the Industrial Revolution in 1750 (Figure 3-1). Emissions rose to
about 5 billion metric tons per year in the mid-20th century before reaching more than 35
billion metric tons per year by 2000 (NOAA, 2020). In parallel, global surface air tem-
peratures have increased by 1 °C. Oceans absorb a large amount of CO2 released into the
atmosphere. This absorbed CO2 reacts with seawater to form carbonic acid. Thus, in-
creased CO2 levels in the atmosphere increase the acidity of the ocean, harming shellfish
and other marine life.

FIGURE 3-1 Increase in annual CO2 emissions (right axis, blue) and the subsequent increase in
the concentration of atmospheric CO2 (left axis, magenta) since the Industrial Revolution.
SOURCE: NOAA (2020).

International efforts to mitigate climate change began in 1992, culminating in


2015 with the Paris Agreement (UNFCC, 2015), whose primary goal is to keep the global

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

34 New Directions for Chemical Engineering

temperature rise during this century well below 2 °C above preindustrial levels. This is a
monumental challenge that will require decarbonization of the energy sector, net-zero
emissions, and fast-paced removal of GHG from the atmosphere. Transitioning to net-
zero emissions in the energy sector is likely to cost trillions of dollars and will require
efforts at all levels of government and across all sectors (e.g., the coordinated, systems-
level approach proposed as part of the Sustainable Energy Corps; Alger et al., 2021). It
will take decades, and may never be complete. The time required to decarbonize energy
systems will depend on technological advances, government policies, changing econom-
ics of energy carrier options, and essential modifications in consumer behavior. Given the
magnitude and complexity of the global energy system, no single energy carrier will be
able to satisfy requirements across all sectors in the foreseeable future. Thus, achieving
the goals of the Paris Agreement will require a wide range of energy sources and carriers.
The question of viable energy mixes was addressed in a comprehensive multi-
model study coordinated by the Energy Modeling Forum 27 (EMF27), which examined
13 scenarios for keeping the global temperature increase below 2 °C in this century
(Kriegler et al., 2014). Because of the significant number of uncertainties and assumptions
involved in these 13 scenarios, it is best to use a notional average of the scenario outcomes
to approximate the various energy trends. A notional average view suggests that in 2040,

 petroleum and natural gas will continue to play a significant role in the energy
mix;
 coal usage will decrease significantly;
 energy carriers from nonbiogenic renewables (e.g., solar, wind, hydroelectric
power) and biogenic sources (bioenergy) will grow significantly; and
 carbon capture, use, and storage (CCUS) will become a key technology for
decreasing CO2 emissions.

The generation, distribution, and use of electrons from renewable energy sources
represent some of the most robust enablers of decarbonization. In its Sustainable Devel-
opment Scenario for 2019–2070, the International Energy Agency (IEA, 2020d) con-
cluded that the share of electricity in end-use energy demand will grow from about 20
percent to more than 50 percent. One-third of that electricity demand is expected to be
met by solar power in the form of photovoltaic devices deployed at scale in decentralized
form (Figure 3-2), with another 20 percent met by modular wind resources.
Opportunities exist worldwide across all sectors (electricity generation, indus-
trial, transportation, and residential/commercial) to decrease energy-related emissions.
Meeting the challenge of keeping the global temperature increase below 2 °C will require
advances in four key areas:

 energy efficiency;
 increased use of lower-carbon energy sources;
 development and deployment of novel energy and energy storage technolo-
gies; and
 government policies to promote cost-effective solutions.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Decarbonization of Energy Systems 35

FIGURE 3-2 Projected global power generation by fuel technology type in the Sustainable
Development Scenario for 2019–2070. The scenario assumes net-zero CO2 emissions by 2060 due
to increases in sustainable energy sources. NOTES: CCUS = carbon capture, utilization, and
storage; PV = photovoltaic; STE = solar thermal electricity. “Other” includes geothermal power,
ocean energy, and hydrogen. SOURCE: IEA (2020d).

This chapter describes the critical role of chemical engineering in the transition
from fossil fuels to renewable energy, with contributions ranging from energy carrier gen-
eration, storage, and distribution to energy use and conversion across various sectors.
Some sectors, such as electricity generation, light- and medium-duty transportation, and
heating and cooling for residential and commercial buildings, are less challenging to de-
carbonize than others (e.g., heavy-duty long-haul ground, aviation, and marine transpor-
tation; energy-intensive cement and steel production). Any low-carbon energy transition
“bridging” strategy will rely on the greater use of a mix of energy carriers: electrons,
hydrogen, and lower-emission liquid fuels (e.g., advanced biofuels, synthetic liquid fuels;
Santiesteban and Degnan, 2021). The energy transition will require hybrid systems that
combine different energy carriers to address the challenges of different usage sectors,
challenges that chemical engineers are uniquely positioned to address. At the same time,
it should be noted that, while chemical engineers can develop technological solutions and
improve the economic competitiveness of those solutions, many of today’s barriers to
addressing climate change are social and political, and chemical engineers work within
that larger societal context.

ENERGY SOURCES

Chemical engineers play central roles in the recovery and development of energy
sources and in energy distribution. This section focuses on the recovery and conversion
of energy and opportunities for chemical engineering with respect to the two primary en-
ergy sources—solar and nuclear. The section on solar energy also includes opportunities
for chemical engineers related to secondary sources (fossil fuels, biomass, and intermit-
tent sources) derived from solar energy.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

36 New Directions for Chemical Engineering

Solar Energy

The predominant source of energy for the planet is the flux of solar photons. Pho-
tons, in contrast to energy carriers, cannot be stored or “bottled.” Their energy is converted
via natural processes into thermal, chemical, or electrical forms, with significant conse-
quences for local and global climate and for human, animal, and plant life. Thermal cap-
ture acts as Earth’s thermostat, with surface temperatures balanced by albedo and green-
house effects that also create the weather patterns from which energy is ultimately
recovered in the form of wind and hydroelectric power. The energy of solar photons is
also stored as chemical energy through photosynthetic cycles that convert CO2 and H2O
(with photons as the energy source and coreactants) into biomass. Biogenic sources have
been used as energy carriers throughout history—first soon after their formation as com-
bustion fuels, but much more extensively in modern times as fossil fuels, well after geo-
logical chemical reactions have deoxygenated these photosynthetic residues and increased
their energy density, forming natural gas, crude oil, and coal. The chemical reduction of
primordial biomass that led to its deoxygenation and storage as fossil fuels is now being
reversed, over much shorter periods of time, through their combustion and consequent
conversion to CO2 and H2O, along with the release of heteroatoms sequestered within the
biomass. These processes have had important local and global environmental conse-
quences, but have also enabled the standard of living enjoyed today. This section describes
solar energy as a primary energy source through the direct capture and conversion of pho-
tons to energy carriers (electrons; H2, NH3, or organic fuels; and heat), and as the source
of fossil fuels (coal, natural gas, petroleum), biofuels (lignocellulosic and other sources),
and intermittent sources (wind and marine).

Direct Capture and Conversion of Photons

Chemical engineering, the discipline most adept at transforming stored energy


carriers into more convenient forms and into chemicals and materials, also brings an en-
abling skillset to the harvesting of photons and the storage of their energy until it is con-
verted into its thermal, electrical, chemical, or mechanical forms. The discipline, closely
collaborating with other disciplines, such as materials science, solid-state chemistry, and
physics, will continue to contribute to the efficiency, durability, and reliability of photon
capture materials and device components; the ability to manufacture and deploy them at
scale; and the optimization, control, and systems-level strategies required for embedding
modular devices within efficient electrical grids. Chemical engineers’ expertise in chem-
istry and catalysis and ubiquitous transport processes has been applied throughout the
discipline’s history to improve the efficiency of chemical transformations. This expertise
will be the enabling tool as photons are used, either directly or via electrons as intermedi-
ates, to affect the synthesis of chemicals or various energy carriers from CO2 and water.
The domains of surface catalysis, photocatalysis, and electrocatalysis are firmly planted
within the chemical engineering discipline.
In the case of photons, their capture and conversion to energy carriers (elec-
trons; heat; chemical energy as H2, NH3, or small hydrocarbons/alcohols) that can be

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Decarbonization of Energy Systems 37

stored and transported to consumers and markets are inseparable. The path to clean en-
ergy from photons, mediated by electrons and molecules that can be stored and trans-
ported, will require decentralized capture within “photon conversion factories,” akin in
concept to the integrated refineries used to transform fossil resources into fuels and
chemicals today, but in much smaller and modular forms and deployed at diverse points
of photon capture. These factories will produce heat, electrons, energy carriers, and
chemicals as part of modular integrated systems designed to capture the largest possible
fraction of the solar flux with high quantum efficiencies and convert it into useful forms
of chemical energy.
Direct photon capture and conversion to electrons as energy carriers. There
are several paths to photon capture and utilization. Direct solar conversion to electrons is
carried out using semiconductors that first capture photons through electronic excitations
across their band gap and then collect the excited electrons in the form of photovoltaic
(PV) solar panels. Together, the decrease in the cost of PV panels and the expected in-
crease in the electrification of energy systems are driving the deployment of global PV
capacity at a rate that could not have been envisioned just a few years ago. This global
capacity, only about 500 GW in 2018, will double by 2022 and is predicted to exceed 10
TW by 2030 and 30–70 TW by 2050 (the current global demand is 18 TW; Haegel et al.,
2019).
Silicon (Si)-based PV cells represent about 80 percent of the currently installed
solar capture capacity in the United States, the rest consisting of cadmium-telluride
(CdTe) semiconducting thin-film PV cells (DOE, 2021a). High-purity amorphous and
polycrystalline Si PV cells are manufactured using energy-intensive purification pro-
cesses first developed for the processing of Si wafers for electronic devices. State-of-the-
art Si PV cells operate at near-theoretical capture efficiencies (~30 percent), a limit set by
the solar spectrum and the balance between the band gaps accessible by doping and the
attainable current densities. The low-absorption cross-section of Si requires thick wafers,
precluding the use of tandem devices designed to collect different components of the solar
spectrum through systematic doping. Si PV cells will continue to evolve through incre-
mental improvements in device architecture and design options at the cell/module scale,
as well as through lower manufacturing costs; greater reliability/durability; and the devel-
opment of infrastructure for their installation, maintenance, and seamless insertion into
advanced electrical grids. Chemical engineers will continue to enable the evolution of Si
PV cells, their deployment at a global scale, and the optimization and control strategies
required to integrate them into the grid. Si PV cells represent the medium-term choice for
deployment at scale in direct photon-to-electron conversion.
CdTe thin-film PV cells absorb photons at energies near the maximum flux in the
solar spectrum. Recent improvements in efficiency and manufacturing costs have made
them competitive with Si PV cells. These modules consist of micron-thick films of CdTe
held within layers of conducting transparent oxides. Environmental concerns about the
toxicity of the components in CdTe PV cells will need to be addressed through improve-
ments in reliability and durability, thinner films, and higher efficiencies. These advances
will enable greater market penetration as PV-based solar capture devices become more

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

38 New Directions for Chemical Engineering

prevalent in practice. Such advances in manufacturing, cell/module architecture, and sys-


tems integration will be driven by chemical engineering as an enabling discipline, as il-
lustrated by recent efforts to coordinate research and manufacturing capabilities through
a National Renewable Energy Laboratory–led consortium.1
The parallel developments in dye-sensitized PV cells have recently been punctu-
ated by the emergence of mesoscopic architectures, in which coatings of n-type semicon-
ductor nanoparticles, such as titania, act as mesoporous anodes that provide 1,000-fold
increases in dye-anode connectivity (Hardin et al., 2012; O’Regan and Grätzel, 1991).
These molecular photovoltaics have emerged in parallel with perovskite solar cells (PSCs;
Grätzel and Milić, 2019), leading to a significant disruption in the nature of research on
PV cells and to very rapid advances in capture efficiencies. PSC devices also provide the
benefits of roll-to-roll solution-based manufacturing processes, a tolerance for reagents
lower in purity, and much smaller amounts of active materials relative to Si PV cells. PSC
devices have evolved rapidly in photon capture efficiency, from 4 percent in 2010 to >25
percent in 2019 (Grätzel and Milić, 2019; Figure 3-3). These PSC devices are now ap-
proaching photocurrents near their theoretical maximum, but improvements in efficiency,
open-circuit voltages, and long-term durability, as well as replacement of the toxic water-
soluble components ubiquitous in the best-performing perovskite materials, need to be
addressed before significant commercial deployment at scale can occur (Correa-Baena et
al., 2017). These cells suffer from operational instabilities, short useful lives, and signifi-
cant environmental and health concerns related to the long-term containment and ultimate
disposal/recycling of their toxic constituents. These challenges are being addressed
through significant funding from federal programs2 and an influx of entrepreneurial cap-
ital.
Concerns about toxicity and long-term durability continue to prevent PSC devices
from displacing Si PV cells in the marketplace. Recent developments have led to more
stable perovskite compositions, to the identification and mitigation of extrinsic degrada-
tion mechanisms, and to device configurations that ensure more reliable containment to
prevent the release of toxic components in the most efficient perovskites (e.g., me-
thylammonium lead trihalides and formamidinium analogs). Durability and containment,
however, remain formidable challenges (Correa-Baena et al., 2017; Rong et al., 2018).
Significant ongoing research focuses on perovskite compositions that minimize intrinsic
degradation processes and on modular device architectures that ensure reliable long-term
operations. As in the case of Si PV cells, chemical engineering is well positioned to ad-
dress and resolve these challenges, and to deploy the improvements in practice at scale.
The development of advanced solution-based processes for perovskite cell manufacture,
integration of PSC systems into existing grids, and life-cycle assessment (LCA) of the full
environmental impact of these devices are also encompassed by fundamentals and prac-
tice of chemical engineering.

1
See https://www.energy.gov/eere/solar/cadmium-telluride-research-and-development-consor
tium-coordination.
2
See https://www.energy.gov/eere/solar/solar-energy-technologies-office-fiscal-year-2020-pe
rovskite-funding-program.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Decarbonization of Energy Systems 39

FIGURE 3-3 Improvements in capture efficiency of perovskite-based solar cells since 2010.
SOURCE: Grätzel and Milić (2019).

These PSC systems absorb light via direct electronic transitions, leading to high
photon absorption cross-sections and to efficient capture using thin films, in contrast to
the thick wafers required for Si PV cells, because of the low photon absorption cross-
sections inherent in their indirect electronic transitions. Such thin films minimize the
amounts of active components needed and provide significant opportunities for solution
coating processes and for the scalable manufacturing of PSC devices (Li et al., 2018).
Thin films also allow the synthesis of flexible devices suitable for curved surfaces; their
transparent nature enables tandem cells with stacked layers of different perovskites de-
signed to capture complementary wavelengths in the solar spectrum and a larger fraction
of the impinging solar flux. A life-cycle analysis of Si-free tandem cells consisting of two
perovskite layers recovered the energy required to manufacture them in 0.35 years, a much
shorter period than the 1.44 years for perovskite-Si tandem cells (Tian et al., 2020a).
Conversion of photons to H2, NH3, or organic fuels as energy carriers. The
previous section addresses the conversion of solar energy via electronic excitations and
ejection and capture of the emitted electrons. In this context, electrons are transported to
markets or stored as chemical energy within batteries to mitigate the intermittency of
solar flux. The capture of the energy of photons as energetic molecules provides alterna-
tive paths for transporting the photon energies in a different form. At the point of capture,
such strategies also deal with the intermittency issues inherent in solar capture. In all
cases, these strategies require modular architectures and significant integration and inten-
sification of the photochemical and electrochemical processes involved.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

40 New Directions for Chemical Engineering

Implementation requires one of the following strategies: (1) direct reduction of a


common molecule (such as H2O, CO2, or N2) used as the vehicle for storing and trans-
porting solar energy using photons directly within slurries of particulate photocatalysts
(direct photocatalysis; Goto et al., 2018; Takata et al., 2020); (2) photoelectrochemical
cells (PECs) that couple photovoltaic and electrochemical cells at the device scale to gen-
erate H2 from H2O, organic energy carriers or H2–CO mixtures from CO2–H2O reactants,
or NH3 from N2–H2O mixtures (photoelectrochemical devices); or (3) spatially separate
modules that use PV devices to generate electrons and electrocatalytic cells that use these
electrons as reactants to reduce the carrier molecules (and their mixtures) to the end prod-
ucts listed in (2) (sequential processes). Such systems have the ultimate potential to de-
liver these energy carriers and chemicals at scale, but they have been demonstrated at
practical scales only for H2 production via strategy (3)—the combination of commercial
PV cells and H2O electrolysis modules, each at the state of the art. These strategies will
require advances in the synthesis, characterization, and mechanistic assessment of cata-
lytic solids, as well as the development of materials that can withstand severe chemical,
photochemical, and electrochemical environments within complex hydrodynamics for
systems that couple the required reactions through diffusional controls. Thus, the combi-
nation of, and advances in, various disciplines and such subdisciplines as catalysis, fluid
mechanics, solid-state chemistry and physics, separations, advanced models and simula-
tion at the microscopic and macroscale levels, process design, and process control will
be important for future breakthroughs. These subdisciplines are all within the domains of
chemical engineering research and practice.
H2 generation via photocatalytic water splitting represents the most direct route
to the capture of solar energy as chemicals for either transport to markets or a means of
addressing intermittency at the point of capture. The state of the art and competitiveness
of the three strategies described above are discussed in several reviews (Ardo et al., 2018;
The Royal Society, 2018a).
Direct catalytic water photolysis uses particulate photocatalysts consisting of an
absorber (e.g., SrTiO3, Ta2N5) dispersed as aqueous suspensions. These photocatalysts
generate electrons and holes that are collected separately at metal nanostructures present
at their surfaces to form H2 and O2 at each location, in systems that are simple in design
and applicable at larger scales than are possible with modular integrated photoelectro-
chemical devices. Collecting the H2 and O2 separately and preventing their recombina-
tion, extending the life of the metal-promoted semiconducting photocatalysts, and im-
proving their capture efficiencies, however, pose significant safety, engineering,
materials, and catalysis challenges (Ardo et al., 2018). This approach, and means of over-
coming its challenges, are the current focus of the Japan Technological Research Asso-
ciation of Artificial Photosynthetic Chemical Process.
An extensive review of the challenges associated with reactor scale-up and syn-
thesis, and of the efficiency of inorganic photoelectrodes provides the most detailed and
up-to-date roadmap for the deployment of photoelectrochemical devices for water split-
ting. This review describes the trade-offs between efficiency and complexity as systems
evolve from direct photocatalysis to integrated photoelectrochemical devices and ulti-
mately to integrated systems with PV-electrolysis modules (Moss et al., 2021; Figure 3-4).

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Decarbonization of Energy Systems 41

FIGURE 3-4 Illustration of several solar-driven water-splitting technologies, sorted from low to
high complexity, efficiency, and modularity. NOTES: E = electricity; PCWS = public community
water system; PEC = photoelectrochemical cell; PV = photovoltaic. SOURCE: Moss et al. (2021).

PEC devices, which combine photon capture and electron generation at the device
scale, show higher capture efficiencies relative to direct photocatalytic water splitting.
However, their complexity and modular architectures represent formidable hurdles for
deployment at scale, as do the lifetime of the photoelectrodes and the delicate architec-
tures required to integrate photon capture and electrolysis at the device scale. As in the
case of direct photolysis, the efficiency of PEC devices decreases as more demands are
placed on materials and interfaces to carry out the combined functions of photon capture,
charge separation and collection, charge transport at a catalytic function, and the molecu-
lar-scale evolution of H2 and O2 via electron transfer at catalytic centers. Such PEC de-
vices show very high photon capture efficiencies at the expense of greater cost and com-
plexity; they represent solutions only for niche applications in the immediate future. Their
ultimate use at scale remains uncertain, and any significant progress toward practical sys-
tems will require addressing engineering design, molecular and electronic transport, and
materials discovery in concert.
These PEC devices for the synthesis of solar H2-based fuels, as well as their ar-
chitectural analogs for artificial photosynthesis strategies for converting CO2–H2O mix-
tures to CO and organic energy carriers, remain at the proof-of-concept stage (Lewis,
2016). They will require the development of materials and interfaces that can efficiently
induce charge separation upon photon-induced excitations and transfer these charges to
catalysts that can form H2 and organic solar-derived fuels before recombination. These
modules will need to be robust and relatively inexpensive for deployment at scale. Many

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

42 New Directions for Chemical Engineering

of these challenges are being addressed as part of the work of large multidisciplinary cen-
ters, such as the Joint Center for Artificial Photosynthesis3 and its recently announced
successors, the Liquid Sunlight Alliance and the Center for Hybrid Approaches in Solar
Energy to Liquid Fuels.4 These centers aim to design in concert the different components
required and develop hybrid photoelectrodes that can combine photon capture and molec-
ular catalysis to generate carriers from a broad range of wavelengths in the solar spectrum.
These advances require a bridge between length and time scales inherent in photon-driven
excitation and molecular transformations induced by emitted photoelectrons at a catalytic
function. Systems-based integration, control, and design; reaction-transport formalisms;
and knowledge of the catalytic properties of active surfaces and centers will play an ena-
bling role in the design and selection of cost-effective devices for the direct generation of
energy carriers from photons in a manner that avoids toxic and scarce elements, as well
as containment and sustainability concerns (Montoya et al., 2017). Bringing such consid-
erations into chemical, biochemical, and electron-driven processes is a domain of chemi-
cal engineers.
Formidable challenges remain for the development of electrochemical systems
for direct or sequential conversion of photon energies into chemical energy in the form of
H2 (from H2O), CO (from CO2), small alcohols and hydrocarbons or H2–CO mixtures for
subsequent thermochemical conversion to such molecular carriers (from CO2–H2O), and
ammonia (from N2–H2O). The most enduring and significant of these challenges are

 the modular nature and complex interconnections among functions and the
integration of electrical and chemical processes at scale;
 the ubiquitous requirement for scarce precious metals as electrodes, and toxic
or rare elements as semiconductors, dye sensitizers, dopants, and connectors;
 the need for process intensification and high photon capture efficiencies lim-
ited by transport processes within electrolytes, electrodes, or semiconductors;
 the durability of modules during extended field use and their recyclability
after their useful life;
 the costly extraction of dissolved product molecules from dilute aqueous me-
dia and the separators required to avoid the recombination of photocatalytic
or electrocatalytic products; and
 the energy requirements in fabrication and recovery of the component ele-
ments after use.

These matters involve catalysis and kinetics in complex and nonideal liquid sys-
tems (thermodynamic and hydrodynamic); transport of molecules, ions, and electrons in
fluids and solids; materials assembly with precise nanoscale and mesoscale architectures;
process integration, control, and optimization; and LCA. The challenges, fundamentals,
and coping/solution strategies lie firmly within the domain of chemical engineering,

3
See https://solarfuelshub.org/.
4
See https://www.energy.gov/articles/department-energy-announces-100-million-artificial-photo
synthesis-research.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Decarbonization of Energy Systems 43

which has in the past adeptly tackled challenges of similar character for complex chemi-
cal, biochemical, and electrochemical conversion processes mediated by heterogeneous,
molecular, or biological catalysts.
Electrolysis remains the proven technology for electrochemical generation of H2
via modular systems based on acid polymer, liquid alkaline, or ionic-transport solid elec-
trolytes (Ardo et al., 2018; Moss et al., 2021). Progress has recently been made in the
scaling up of electrolysis, and chemical engineers have an opportunity to contribute to the
development of applications at the scale required to disrupt the energy landscape. Elec-
trolysis systems can be operated in acidic or alkaline regimes, although the rate-limiting
nature of the O2 evolution half-reaction (H2O oxidation) has led to a preference for alka-
line electrolyzers, which also avoid the platinum-based materials required to prevent elec-
trode dissolution in acidic media, thus allowing the use of nonprecious metals (e.g., Ni,
Fe, Cu) as electrodes. Acidic electrolyzers use polymer electrolyte membranes that mini-
mize contact between H2 and O2 through fast proton transport and short anode–cathode
distances. Alkaline electrolyzers have relied on microporous physical barriers that impose
larger anode–cathode distances and ohmic losses. Recent developments in selective anion
transport membranes prevent contact between H2 and O2 and have led to more compact
membrane–electrode modules.
The challenges of deploying electrolyzers at scale include their efficient integra-
tion and control as multimodule stacks, the development of earth-abundant electrode ma-
terials, and thinner and more efficient separator membranes. The challenges are similar
but even more formidable for electrochemical reduction of CO2 via concurrent electroly-
sis with H2O to form mixtures of H2, CO, alcohols, carboxylic acids, and hydrocarbons
(Hori, 2008; Nitopi et al., 2019). These products can be used directly as energy carriers
or as precursors to such carriers on heterogeneous catalysts (e.g., H2–CO conversion to
liquid transportation fuels via Fischer-Tropsch or methanol synthesis). Electrochemical
systems face similar challenges in meeting the scale required for impact:

 more efficient and robust electrodes based on earth-abundant elements;


 thinner and more selective membrane separators;
 higher-temperature electrolyzers;
 integration of electrocatalytic and thermocatalytic systems in sequence or
within a single device to form more suitable energy carriers; and
 the development of supply chains to lower the costs of manufacturing and
integrating the modular devices into molecular weight (MW)–scale distrib-
uted deployments for the synthesis of H2 and other energy carriers.

Conversion of photons to heat as an energy carrier. The modular nature of PV


and electrochemical cells, whether in separate or combined forms, poses significant chal-
lenges, including process integration and intensification and deployment at scale. These
challenges can be addressed by using solar thermal strategies that capture the energy of
photons as heat, which can be delivered to users as thermal energy, or converted to chem-
ical energy for transport or for storage during diurnal or intermittent fluctuations in solar
flux. The end use of the captured thermal energy depends on the temperatures accessible

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

44 New Directions for Chemical Engineering

through solar collectors and heat transfer media and on the location of markets relative to
the point of capture. The specific end-use option of generating H2 at scale has been high-
lighted in recent reports because of its relevance to a hydrogen economy and its inherent
advantages over electrolyzers in large-scale deployment (Gonzalez-Portillo et al., 2021;
Moss et al., 2021; NREL, 2017).
This capture and storage of photon energies as heat can be used directly in ambi-
ent temperature control in commercial or residential spaces; as process heat in existing
chemical processing plants or manufacturing operations; or for conversion into electrons
or hydrogen, typically through the generation of high-pressure steam or through conver-
sion cycles commonly termed “chemical looping.” The latter approach involves the de-
sign of reactors based on the chemical engineering principles of kinetics, transport, chem-
ical absorbents, and construction materials that can withstand the extreme temperatures
required for thermal and chemical efficiencies.
The achievement of very high temperatures (~1400 oC) has been enabled by ad-
vances in parabolic solar thermal concentrators (Sargent & Lundy LLC Consulting Group,
2003); these systems have in turn allowed the generation of very high–pressure steam for
power generation in high-temperature turbines. At such high temperatures, chemical loop-
ing using redox-active oxides enables the cycling of such solids between a reduced state
that reacts with water to form H2 and an oxidized state that evolves O2 in a spatially sep-
arate stage. Such processes use oxides of earth-abundant elements and form separate
streams of H2 and O2 from H2O, thus avoiding the need for gas separators and the use of
costly metals as electrodes, as well as the transport limitations inherent in the use of liq-
uids as electrolytes.
These cyclic processes can be deployed via large-scale devices that allow process
integration and intensification strategies unavailable for modular systems but ubiquitous
in conventional refining, chemical manufacturing, and power generation processes, albeit
at somewhat lower temperatures. The high temperatures required for efficient solar ther-
mal capture pose significant challenges with respect to the design, use, and handling of
the heat transfer media and the durability of the required redox-active oxides. Molten salts
are typically used as heat transfer fluids, but solid particles in fluidized systems have re-
cently emerged as attractive alternatives (Gonzalez-Portillo et al., 2021).

Fossil Fuels

Fossil fuels (coal, petroleum, and natural gas) have been essential for society’s
development and progress, having powered the Industrial Revolution and shaped the mod-
ern world. Technological advances based largely on the ingenuity of chemical engineers
have enabled the efficient extraction, processing, and conversion of fossil fuel raw mate-
rials into useful products. The expertise of chemical engineers remains essential in ena-
bling a transition from the current energy landscape to one based on renewable and sus-
tainable energy sources. However, most studies have concluded that fossil fuels will
continue to play a key role in the energy mix at least until 2050, and in the meantime, the
imperative is to reduce the carbon footprint of fossil fuels—a key opportunity for chemi-
cal engineers.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Decarbonization of Energy Systems 45

Coal. Coal is the most abundant and least expensive of all fossil fuel resources.
It can be converted into gas or liquids to produce chemicals and fuels, but is used primarily
for direct combustion. As a solid fuel, coal generates more CO2 per unit energy than other
fossil fuels—from approximately 30 percent more than diesel to nearly twice as much as
natural gas (EIA, 2021i).5 Coal resources are used mainly to generate electricity, and
while reliance on coal has increased in low- and middle-income countries (e.g., China and
India), the opposite has been true in higher-income countries (e.g., United States, Euro-
pean Union, Japan). In the United States, electricity generation from coal has been declin-
ing since 2008, with the biggest drop (~16 percent) taking place in 2019; in contrast, the
use of natural gas has increased dramatically since 1995, and the use of renewables has
increased since 2005, albeit at a slower pace (Figure 3-5).6

FIGURE 3-5 Annual electricity generation by different sources (natural gas—dark blue; coal—
dark red; nuclear—yellow; wind—green; hydro—light blue; all other sources—gray) in the United
States, 1970–2019. SOURCE: EIA (2020d).

Most investments in coal-fired plants in 2019 (almost 90 percent) were for higher-
efficiency (supercritical and ultrasupercritical) plants; the remaining small portion of in-
vestments were in inefficient subcritical plants, mainly in Indonesia (IEA, 2020c). High-
efficiency coal-fired power plants use water at high, above-critical temperatures and pres-
sures (373 °C and 220 bar, respectively). The efficiency gains thus achieved reduce by
about 20 percent both the amount of coal needed and CO2 emissions. These plants also
emit substantially lower amounts of nitrogen oxides and sulfur oxides (IEA, 2012).

5
Pounds of CO2 emitted per million Btu (gJ): coal, 215 (227); diesel, 161 (170); gasoline, 157
(166); propane, 139 (147); natural gas, 117 (123) (EIA, 2021).
6
Preliminary IEA analysis indicates a sharp drop in power-sector demand in 2020 as a result of
the COVID-19 pandemic, with demand for coal having the greatest uncertainty of all fuels used
for power.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

46 New Directions for Chemical Engineering

To capitalize on coal’s advantages and help mitigate its disadvantages, research


and development (R&D) is needed to increase thermal efficiency, demonstrate cost-ef-
fective and secure carbon capture and storage, further improve emission controls, and
reduce water consumption. Meeting these challenges will require research to improve ex-
isting and develop new breakthrough technologies. The Electric Power Research Institute
has recommended the following key goals for such efforts, all areas in which chemical
engineers can play a role (Maxson and Phillips, 2011):

 improved plant efficiency via high-temperature materials and higher turbine


inlet temperatures;
 cost-effective and scalable CO2 capture in new or retrofitted applications;
 environmentally safe and permanent storage of CO2;
 improved emission control systems that can achieve near-zero emissions of
all pollutants; and
 advanced cooling and water management methods to reduce water demand
and pollutant discharges.

Progress has been made in several of these areas (e.g., improved plant efficiency, im-
proved emission control systems, and water management), but less so in the implementa-
tion of viable CCUS processes.
Natural gas. Natural gas contains mostly methane, but also small amounts of
ethane and varying amounts of heavier hydrocarbons, including propane, butane, and pen-
tane. The ethane and heavier hydrocarbons in natural gas are typically referred to as nat-
ural gas liquids (NGLs). Natural gas can also contain CO2, sulfur, helium, nitrogen, hy-
drogen sulfide, and water, which are removed before it is used as an energy source.
Processing plants remove water vapor and nonhydrocarbon compounds, and the
NGLs are separated from the wet gas and sold separately. The separated NGLs are called
natural gas plant liquids, while the processed natural gas is called dry, consumer-grade,
or pipeline-quality natural gas. Some natural gas is dry enough to satisfy pipeline trans-
portation standards without processing. Odorants (light mercaptan compounds) are added
to aid in the detection of pipeline leaks. Pipelines transport dry natural gas to underground
storage fields or to distribution companies and eventually to consumers (EIA, 2021g; Fig-
ure 3-6).
In the mid-2000s, a step change in natural gas and shale oil production occurred
in the United States. Often referred to as the shale revolution, this innovation was made
possible by a combination of hydraulic fracturing and horizontal drilling techniques that
enabled economical oil and gas production from shale formations. The result has been a
near doubling of U.S. natural gas production, from 18 trillion cubic feet in 2005 to ~34
trillion cubic feet in 2019. The United States is now a world leader in natural gas and oil
production, and a global supplier. In the United States, natural gas is used primarily for
electricity generation (power sector) and for heating (industry, residential, and commer-
cial), with a small fraction used for transportation. Figure 3-7 shows natural gas consump-
tion by the various sectors in 2019 and its evolution over the period 1950–2019.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Decarbonization of Energy Systems 47

FIGURE 3-6 Schematic diagram of natural gas production, processing, and delivery to end users.
NOTE: LNG = liquified natural gas. SOURCE: EIA (2021f).

FIGURE 3-7 Natural gas consumption by sector (residential, industrial, transportation, commer-
cial, and electric power), 1950–2019. A net increase in natural gas consumption occurred over this
time period across all sectors. SOURCE: EIA (2021e).

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

48 New Directions for Chemical Engineering

It is generally accepted that, relative to other fossil fuels, natural gas provides a
cleaner bridge to a renewable energy future, and it is the only fossil energy source pro-
jected to grow in the coming decades (DOE, 2018b). However, the longer-term future for
natural gas is less certain. Innovation throughout the entire value chain will be required if
natural gas is to continue being a key contributor to the future of the low-carbon energy
mix. Areas in which chemical engineers will play a key role include the following:

 Production
- Advances in water-quality management; water recycling for shale or un-
conventional gas production
- Further reduction of methane venting to the atmosphere
- Accelerated development of CO2 to replace water as a fracturing agent
 Processing
- Development of low-energy processes for natural gas separation and pu-
rification
 Storage and transportation
- Methane leakage control
- Higher-efficiency compressors and heat exchangers
- Smart sensors for pipeline operational efficiency
- Materials for intercontinental transport via pipeline versus the current
practice, which involves liquefaction and regasification
- Low-cost pipeline materials to enable cotransport of natural gas and high
concentrations of hydrogen (>20 percent)
 Use
- Development of commercially viable CCUS technologies
- Improved efficiency of the overall natural gas system, including in-
creased combustion efficiency and waste-heat recovery, and develop-
ment of innovative controls and low-cost sensors that enable data-driven
operations
- Development of technologies for trigeneration (combined cooling, heat-
ing, and power systems
- Design of novel processes for integration of natural gas with renewables,
particularly solar and wind
- Design of novel processes for production of low- or zero-CO2 hydrogen
(e.g., “blue” hydrogen [with CCUS] and “purple” hydrogen [with black
carbon and/or carbon nanotubes coproduction]; see the discussion of hy-
drogen below)

Petroleum. Chemical engineers have played a central role in the oil industry from
its beginning, initially converting crude oil into useful products in small and simple refin-
eries, and subsequently optimizing large and integrated refineries to address energy effi-
ciency and environmental concerns in the manufacture and use of transportation fuels and
chemicals. Chemical engineers have also worked closely with geologists to maximize the
recovery of conventional and unconventional fossil resources.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Decarbonization of Energy Systems 49

The oil industry and chemical engineering evolved together. Many advances in
chemical engineering science and technology were driven by the needs of the oil industry,
and these advances in turn have spurred growth in the oil industry. With increasing global
focus on the need to accelerate the transition to low-carbon energy to mitigate climate
change, the expertise of chemical engineers is required now more than ever to help the oil
industry minimize its carbon footprint. As discussed previously, the energy system is
enormous and complex, and the transition to a low-carbon energy mix will take decades;
in the near term, the need for petroleum and its derivative products will continue.
Petroleum is the largest energy source in the United States, used both as a fuel for
transportation (road, aviation, marine, and rail) and as a feedstock for the manufacturing
of such products as plastics, fibers, lubricants, paints, and solvents. In 2020, U.S. crude
oil consumption averaged about 18 million barrels per day, with the transportation sector
accounting for 66 percent and the industrial sector for 28 percent of this total (EIA,
2021k).
U.S. oil production totaled 9.6 million barrels per day in 1970 and declined over
the subsequent 35 years. Production in 2005 was 5.2 million barrels per day, and imports
reached more than 10 million barrels per day—just under 50 percent of total U.S. crude
oil consumption. Since 2010, however, the combination of hydraulic fracturing and hori-
zontal drilling has enabled access to oil trapped in shale, making the United States a top
world producer of crude oil (Figure 3-8).

FIGURE 3-8 Amount of crude oil produced in millions of barrels per day for the top five oil-
producing countries (United States, former U.S.S.R./Russia, Saudi Arabia, Iraq, Canada), 1980–
2019. SOURCE: EIA (2021j).

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

50 New Directions for Chemical Engineering

In 2020, the United States produced more than 11 million barrels of crude oil
per day, with tight/shale oil accounting for about 65 percent of this total (EIA, 2021c).
The U.S. Energy Information Administration, projects that tight/shale oil will remain the
main source of crude oil produced in the United States (EIA, 2021b; Figure 3-9). Relative
to conventional crude oil, tight/shale oils contain lighter hydrocarbons, have higher H/C
ratios, and are generally very light crude (API [American Petroleum Institute] gravity 45–
50) and sweet (<0.1 percent sulfur). They also require less energy to process into desired
products. Thus, they are positioned to play an important role in the oil industry’s efforts
to minimize its carbon footprint.

FIGURE 3-9 Historical crude oil production, 2000–2020, and projected crude oil production,
2020–2050, in the United States. An increasing share will come from tight/shale oil sources.
SOURCE: EIA (2021b).

The challenge of unconventional tight/shale oil lies in improving its extraction,


as it is stranded within geological features that are difficult to image and access because
of their low permeability. Thus, the reservoirs need to be hydraulically fractured to create
paths for the flow of oil and gas. This process requires either hydraulic fracturing of the
geological systems to create paths for flow or horizontal drilling over long distances be-
fore fracturing, using explosives or high-pressure water containing various proppants
(small particles such as sand or ceramic beads) and chemicals (Geoscience News and
Information, 2021). Proppants, as their name implies, prop the fractures open, and the
chemicals create a viscoelastic fluid to carry the proppants. The U.S. Department of En-
ergy (DOE) is sponsoring research aimed at enhancing the ultimate recovery of oil and
gas from both existing and new wells in mature and emerging basins (DOE, 2021b). Areas
in which chemical engineers can contribute to innovation in tight/shale oil production are
described below.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Decarbonization of Energy Systems 51

Improved water management. As discussed in the above section on natural gas,


extraction of tight/shale oil and gas requires a large amount of water and produces a large
amount of water that requires treatment or disposal. Low-cost technologies for produced-
water treatment are required to maximize water usage recycling, thus minimizing fresh-
water usage. Produced water may contain injected chemicals plus naturally occurring ma-
terials such as brines, metals, radionuclides, and hydrocarbons. The flowback and pro-
duced water are usually stored in tanks or pits before treatment or disposal, often through
underground injection (DOE, 2021b; EPA, 2021b). Potential options for either replacing
water as a working fluid or minimizing its use include the use of liquefied propane gas
(LPG; e.g., API, 2021), supercritical CO2 (Song et al., 2019), or microwave fracking (e.g.,
Aresco, 2021).
Increased recovery to extend well life. The amount of oil produced in primary
recovery from an unconventional reservoir is much smaller than that produced from a
conventional reservoir. In addition, production rates from unconventional wells often de-
cline by more than 50 percent in the first year. Improved fracturing technology to create
more efficient and durable oil and gas flow pathways is therefore needed. Chemical engi-
neers can contribute to meeting this challenge by applying their understanding of mass
transport in porous materials.
Data-driven approaches. The DOE national laboratories, in collaboration with
universities and industry, are leading an effort to integrate physics-informed statistical
models; inverse models, such as neural networks; natural language processing; big data
analytics; and other emerging artificial intelligence/machine learning (AI/ML) technolo-
gies to draw meaningful insights from reservoir data for real-time rapid visualization and
prediction to enable effective decision making (DOE, 2021b).
Methane management. The atmospheric concentration of methane, a more potent
GHG than CO2, has risen steadily for more than a decade (Nisbet et al., 2019). This trend
reflects the increased production of shale oil and gas, as well as the natural (e.g., from
wetlands and other flood zones) and biogenic (e.g., from agriculture or waste) emissions
that also play a role. As a result of regulations, the oil industry has made good progress in
reducing methane emissions; however, more progress is needed. The main source of fu-
gitive methane emissions is well venting, followed by pneumatic devices that use natural
gas as the operating fluid, as well as storage and transport venting and leaks.
Chemical engineers can enable significant reductions in methane emissions by

 developing low-cost modular technologies for conversion of methane to liq-


uid products to replace venting;
 developing methods for using air instead of natural gas as the operating fluid
for pneumatic controllers;
 improving techniques and developing smart sensors for methane leakage con-
trol and detection; and
 improving methods for deploying higher-efficiency compressors.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

52 New Directions for Chemical Engineering

Biofuels

The production of biofuels from biogenic carbon, such as waste plant matter, al-
gae, and organic waste, has long been heralded as a means of offsetting GHG emissions
from the combustion of fossil fuels (e.g., Lynd et al., 1991; Pacala and Socolow, 2004).
The combustion of waste plant matter—lignocellulose—and other biogenic carbon for
cooking and home heating has been practiced since before recorded history. The well-
known conversion of carbohydrates into fuel ethanol has also been pursued for more than
a century. The use of ethanol from starch-based sugars as a fuel was advocated by Henry
Ford in the early years of the automobile industry, an approach superseded by the devel-
opment of crude oil production and refining. Fuel ethanol has been produced successfully
at scale since the 1970s in Brazil and later in other countries, predominantly from sugars
derived from sugarcane and cornstarch. Annual production in 2019 reached nearly 18
billion gallons (430 million bbl) in the United States (EIA, 2020a) and 29 billion gallons
(690 million bbl) worldwide (EIA, 2021a).
By 2019, the annual production and use of biodiesel, mainly from waste oils and
fats, had risen to about 2.5 billion gallons (60 million bbl) per year in the United States,
with a global production of 10.9 billion gallons (260 million bbl; EIA, 2021a). The
COVID-19 pandemic notwithstanding, the contribution of biofuels to the global transpor-
tation sector has increased each year over the last two decades. Though reasonably mature
today, starch- and sugar-based ethanol and biodiesel still pose challenges for chemical
engineers in the areas of extracting value from by-products (e.g., glycerol in biodiesel
production), capturing and sequestering CO2 from ethanol production processes, and ex-
panding the range of fuel products beyond ethanol at a scale of impact. These challenges
remain significant barriers to improving the economic feasibility of at-scale biofuel pro-
duction.
Interest in biofuel production as a potential strategy for offseting GHG emissions
and decreasing dependence on fossil resources has seen a resurgence, prompting substan-
tial debate about the long-term viability and utility of biofuels. From a technical perspec-
tive, most biogenic feedstocks have lower energy content than their fossil-based counter-
parts because of the high oxygen content of lignin and plant-derived polysaccharides (e.g.,
sugars contain about 54 percent oxygen by mass; Figure 3-10). Thus, for nearly all pro-
cesses for converting biogenic, oxygenated compounds to energy-dense liquid fuels, the
production of high-density fuels at a reasonable cost relative to the amortized, technolog-
ically mature petroleum industry represents a significant challenge.
The above technological challenges combine with the potential environmental
consequences of harvesting crops for energy use to generate considerable controversy
(Searchinger et al., 2008). The environmental cost of clearing agricultural land and its loss
for growing food crops, complications related to water use, and the potential for an un-
certain landscape and diverse mix of political and tax boundaries around the world make
the future of biofuels uncertain. Furthermore, the potential rapid growth in the adoption
of EVs may reduce demand for transportation fuels more broadly. The impact of biofuels
in the energy sector will depend critically on bringing judicious, rigorous, and transparent

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Decarbonization of Energy Systems 53

economic, environmental, and technical analyses—hallmarks of the chemical engineering


profession—to bear on the selection of viable options for biofuel production.

FIGURE 3-10 Representative oxygen content of various feedstocks and materials ranges from 0
percent to more than 70 percent (weight percent). The oxygen content of energy sources varies
widely. The higher oxygen content of biofuels results in a lower energy content relative to fossil-
based energy sources. NOTES: AA = acrylic acid; FDCA = furan-2,5-dicarboxylic acid; LAB =
linear alkyl benzene; MEG = monoethylene glycol; NG = natural gas; PE = polyethylene; PEF =
polyethylene furanoate; PET = polyethylene terephthalate; PP = polypropylene; PTA = purified
terephthalic acid.

Lignocellulosic feedstocks. An estimated 1 billion tons of lignocellulose could


be sustainably harvested for biofuel production annually in North America (DOE, 2016).
The conversion of lignocellulosic biomass into biofuels and other organic coproducts—
which represents the foundation of the biorefinery concept—poses considerable chal-
lenges for chemical engineers. Lignocellulose is a complex and heterogeneous composite
material whose carbon content is predominantly in the form of two polysaccharides—
cellulose and hemicellulose—as well as the aromatic polymer lignin. Numerous pro-
cessing options have been considered over many decades; they entail either fractionating
biomass into its constituents (thus enabling selective processing options, akin conceptu-
ally to the methods used in petroleum processing) or processing biomass directly into
liquid or gaseous intermediates for subsequent conversion to transportation fuels and
chemicals. Unlike petroleum, biomass is a solid, polymeric material that can be substan-
tially heterogeneous, a feature that presents feedstock-associated challenges beyond the
processing of liquids and gases that form the bedrock of the chemical engineering disci-
pline.
The combined challenges of converting lignocellulosic biomass and offsetting
fossil resources as a feedstock for transportation fuels create considerable opportunities
for chemical engineering to continue making enabling contributions to the at-scale con-
version of lignocellulose into biofuels and other products (flexible feedstocks are dis-
cussed in Chapter 6). These opportunities begin with the crops themselves. Now that bi-
ology is a core component of the chemical engineering discipline, modern molecular
biology techniques are now ubiquitous in the toolkit of many chemical engineers. These
techniques can be used to modify plants so they can be more efficient photon collectors
(Kromdijk et al., 2016), to produce target chemicals in planta (Yang et al., 2020), and to

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

54 New Directions for Chemical Engineering

reduce the recalcitrance of plant components in chemical conversion in a biorefinery


(Chen and Dixon, 2007).
Once plants have been harvested and transported to a biorefinery or centralized
depot, there are many opportunities for overcoming recalcitrance to enable conversion of
plant-based biomass to biofuels. The diversity of conversion pathways precludes a com-
prehensive review; therefore, this section focuses on the challenges and opportunities for
chemical engineers to achieve cost-effective and sustainable biofuel production.
In the conventional biochemical conversion pathways, thermochemical treatment
processes increase the reactivity of biomass, mainly by improving physical access to pol-
ysaccharides. These processes apply acid, base, steam, organic solvents, ionic liquids, or
deep eutectic solvents, usually at temperatures in excess of 100 °C. Pretreatment ap-
proaches can also take the form of fractionation methods that separate polysaccharides
from lignin for more direct and selective processing in parallel process trains. Polysac-
charides are subsequently converted into monomeric sugars via carbohydrate-active en-
zymes or sugars and dehydration products, such as furanics or levulinic acid, through
further use of acid catalysts. Soluble carbohydrates and derivatives are then converted
into fuel molecules or precursors through biological and/or chemical catalysis. In the pi-
oneer cellulosic ethanol plants built in the 2000s, lignin is commonly used to provide heat
and power via on-site combustion.
Many attempts have been made to bring biochemical conversion–focused biore-
fineries to scale (1,000 to 2,000 metric tons/day), especially with the aim of converting
nonfood crops or agricultural residues (e.g., wheat straw, corn stover), supported by sub-
stantial government investments. Yet the formidable challenges of economical feedstock
collection, feedstock handling, biomass pretreatment, aseptic solid–liquid separation, and
sterile bioconversion continue to prevent these facilities from achieving the required ca-
pacity factors for economic viability at scale of impact.
The lessons learned from these early facilities inform the many opportunities for
chemical engineers to advance this field in moving from the process paradigm described
above. For example, because on-site lignin combustion is estimated to be the most expen-
sive single unit operation in a biorefinery, as well as a major source of non-GHG emis-
sions, the development of alternative uses for lignin, 40 percent of which can be made of
biomass carbon, is a major opportunity for research (Davis et al., 2013; Eberle et al.,
2017). Successful conversion of lignin into value-added biofuels or biorefinery coprod-
ucts represents a major frontier for the chemical engineering community. Process inten-
sification (PI) through the consolidation of biomass deconstruction into fewer unit opera-
tions and a focus on the elimination of costly steps is critical. An important component of
PI involves separations, which are often key cost drivers in biorefineries. The challenges
in lignocellulosic separations differ from those in petroleum processes because they in-
volve the handling of solids and because biomass-derived compounds consist of high–
boiling point oxygenates. Thus, separation technologies that operate wholly in the con-
densed phase will likely be necessary for the biorefinery.
Beyond biochemical conversion strategies, other lignocellulosic conversion
routes employ fast, thermal deconstruction of the whole biomass or fractions thereof. Hy-
drothermal liquefaction uses liquid water at temperatures above 250 °C to produce a liquid

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Decarbonization of Energy Systems 55

biocrude stream that can be catalytically converted into biofuel molecules. Alternative
biomass pyrolysis routes use oxygen-free environments at or above 500 °C to produce
bio-oil, light gases, and char from lignocellulosic biomass; some of these products are
deoxygenated catalytically, either in the pyrolysis reactor or in subsequent process steps.
At an even higher temperature (>700 °C), synthesis gas (CO and H2) can be produced
from biomass via gasification in mildly oxidizing hydrothermal environments. Research
opportunities common to these high-temperature biomass conversion processes include
the need to understand the complex reaction networks, the design of catalysts and catalytic
processes for substantial deoxygenation and operation in the presence of common catalyst
poisons entrained in and originating from biomass, and the challenges of operating con-
tinuous high-pressure processes.
In most biofuel production processes, chemical coproducts are often invoked as
a requirement for economic viability, with the associated challenges of the very large–
scale disparity between these two value streams. Even ethylene, the chemical produced in
largest amounts from fossil sources, is produced in amounts approximately an order of
magnitude smaller than diesel and gasoline, while other commodity-scale chemicals are
dwarfed by the scale of ethylene. Ultimately, expensive, small-volume coproducts cannot
serve as adequate justification for expensive biofuel production processes, and chemical
engineers will play an important role in finding realistic, scalable solutions. It is important
to note the annual scale of global petroleum production: 5.0–5.5 billion metric tons
(100.69 million bbl per day) of crude oil and 4.1 trillion cubic meters (3.6 billion tons of
oil equivalent) of natural gas (IEA, 2021c,d; EIA, 2021d). Petroleum refineries have
scales ranging from 10 million to 130 million metric tons per year globally. For biofuels
to compete economically with fossil fuels, they need to be produced at similar scales, and
even that may not be sufficient because of their unfavorable (oxygen) stoichiometry.
Nonetheless, government regulations and/or the implementation of a carbon tax may bring
the production of biofuels to a scale of impact, as in the recent case of ethanol in the
United States and Brazil.
Feedstocks beyond lignocellulose. Waste plant biomass is not the only source
of renewable or waste carbon for producing biofuels, as is evident from global efforts to
use algae, which can grow on marginal lands and in ocean or brackish water. The eco-
nomical conversion of algae to lipids and other biofuel precursors has not been achieved,
however, despite extensive research over decades. Many engineering challenges remain,
including cost-effective cultivation in open ponds or controlled photobioreactors, greater
CO2 and photon capture efficiency, separation of target products from cells, and catalysts
and processes for the downstream conversion of algae-based intermediates (e.g., fatty ac-
ids and carbohydrates) to biofuels.
Some organic waste feedstocks are also of potential use in biofuel production (see
the discussion of feedstock flexibility in Chapter 6). Given that municipal solid waste
(MSW) contains substantial organic matter (e.g., food waste, paper, and cardboard),
chemical engineering has many opportunities to increase the use of such feedstocks
through research in fractionation and separations, as well as combined conversion ap-
proaches. Similarly, industrial and consumer-based food production yields substantial,
often highly reduced (deoxygenated) waste feedstocks, such as oils and fats, that can be

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

56 New Directions for Chemical Engineering

catalytically converted to biofuels, although that process poses substantial challenges. In


moving toward a zero-waste society, these organic waste feedstocks, among others, offer
substantial opportunities for chemical process development.
Lastly, it is noteworthy that the conversion of CO2 or gaseous mixtures, such as
flue gas from coal-fired power plants or gases from steel processing, is of interest to the
chemical engineering community. Considerable effort is currently devoted to realizing
the potential of CO2 and other gas conversions via biological, electro-, and thermal catal-
ysis routes (e.g., Ye et al., 2019). It will be critical to consider substrate concentrations
(e.g., direct CO2 air capture and conversion is a major challenge), the source of reducing
equivalents, and the cost and sustainability of any type of process in this vein, as discussed
earlier in this chapter in the context of photon capture.
Overall, biofuels will play a role in reducing GHG emissions associated with
transportation fuels. However, judicious analyses of process feasibility, economics, and
environmental impact will be critical for deciding among the many options; LCA will
provide the rigor needed for such analyses. The challenge for chemical engineers is to
identify those options that will ultimately be economically successful and sustainable at
the scale required to meet society’s fuel needs.

Intermittent Energy Sources

Wind. Wind has provided a source of power for centuries. Three main types of
wind turbines are used today:

 distributed or “small” wind turbines (<100 kW), which are used to power a
home, farm, or small business directly and are not integrated into the electri-
cal grid;
 utility-scale wind turbines (100 kW to several MW), which deliver electricity
to the grid for distribution to end users; and
 larger offshore wind turbines (up to 15 MW).

Wind energy, a niche option a few decades ago, is now the largest source of re-
newable electricity in the United States. In 2019, wind energy output represented about 7
percent of the U.S. electricity supply, with 100 GW of capacity—equivalent to powering
about 32 million homes (AWEA, 2020). On a global scale, wind energy accounts for
about 5 percent of electricity demand (IEA, 2020d).
Chemical engineers are involved in several areas of wind energy production
(Veers et al., 2019). Specific challenges in materials research, development, and imple-
mentation include

 carbon composites and/or recycled materials for turbine blades;


 cement and steel manufacturing with lower CO2 emissions for wind turbine
structural units, such as motors and gear boxes; and
 metallurgy and lubricants for state-of-the-art wind turbines.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Decarbonization of Energy Systems 57

The large diameter of modern turbines poses significant manufacturing and trans-
portation challenges and the need for modular manufacturing and on-site assembly for
both onshore and offshore installations. The decentralized deployment of installations for
capturing wind energy also requires local energy storage and robust sensor and control
systems, and often creates environmental concerns regarding the impact on coastal eco-
systems and land and ocean animal life (NREL, 2020). Finally, as with all renewable
energy sources, challenges exist with respect to the integration of wind energy into chem-
ical production (Centi et al., 2019) and end-of-life considerations for turbine components.
Marine. Marine energy includes energy derived from ocean waves, tidal move-
ments, ocean and river currents, salinity gradients (i.e., where a river empties into the sea),
and thermal conversion (i.e., based on the temperature difference between surface sea-
water and deep [~1 km] seawater). Economical production of tidal energy requires tidal
waves larger than 3 m. The United States has several demonstration projects in tidal-en-
ergy power production, but none are producing power at commercial scale. Overall, ma-
rine energy’s development level is similar to that of wind energy roughly 30 years ago,
which is to say that wind and solar energy are at commercial scale, while wave energy is
at precommercial scale (see IRENA, 2020, for marine energy status and prospects).
Chemical engineers can potentially make contributions to marine energy through the de-
velopment of

 materials capable of withstanding seawater corrosion;


 flexible materials capable of handling the fatigue loads imparted by waves
with their fast cycles of 8–10 s;
 antifouling coatings for submerged equipment; and
 electroactive polymers—polymers that generate electricity from mechanical
stimuli (e.g., dielectric elastomers, piezoelectric materials, ionic polymer
metal composites, and triboelectric materials)

Nuclear Energy

Nuclear power plants use heat produced during nuclear fission to produce steam,
which is used to spin large turbines that generate electricity. The current technology is
based on nuclear fission in pressurized water-moderated reactors (light water reactors).
Fast breeder reactors have been in development for several decades, and some are now in
commercial operation in Russia. While there is renewed interest in nuclear fusion, its po-
tential commercial deployment is still decades away, and the role of chemical engineering
in this area is likely to be marginal and is therefore not discussed here.
The United States is the world’s largest producer of nuclear power, accounting
for more than 30 percent of worldwide nuclear electricity generation. Nuclear power has
contributed almost 20 percent of electricity generation in the United States reliably and
economically over the past two decades. It has been the single-largest contributor (more
than 70 percent) of U.S. non-GHG-emitting electric power generation (DOE, 2021c).
However, the actual and/or perceived hazards of nuclear power plants and the public’s
negative perception of nuclear power have contributed to its slow growth. As of January

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

58 New Directions for Chemical Engineering

2021, the United States had 94 operable reactors (96,550 MW); 39 inactive reactors
(18,140 MW); and two new reactors under construction in Georgia, with a planned elec-
tricity generation capacity of about 1,100 MW each (WNA, 2021).
Advanced nuclear industrial cogeneration offers a potential pathway with suffi-
cient heat and energy intensity to address the problems of industrial emissions at scale.
Traditional nuclear reactors rely on large light water reactors operating at maximum tem-
peratures below 300 °C—a temperature high enough to make steam for power generation
but too low to drive industrial processes. Consequently, the nuclear power industry is
currently focused solely on power generation. Advanced reactors have higher output tem-
peratures relative to light water reactors—up to 600 °C for molten salt reactors and 900
°C for high-temperature gas reactors. These higher temperatures are sufficient to drive
most petrochemical processes. During the past decade, DOE has explored using this heat
for industrial processes with the Next Generation Nuclear Plant.
Advanced nuclear reactors have the potential to provide the heat required by var-
ious industrial processes. Significant cost reduction is required for this advanced technol-
ogy to be affordable for industrial heat generation, but with some new concepts based on
low-cost natural gas, competitive, cost-effective solutions are not out of reach. Efforts to
drive down cost are focused in three areas:

 New qualified fuels—TRISO (TRi-structural ISOtropic) particle fuel and


molten salts—offer fundamentally better process safety profiles. Each TRISO
particle is made up of a uranium, carbon, and oxygen fuel kernel, which is
encapsulated by three layers of carbon- and ceramic-based materials that pre-
vent the release of radioactive fission products. These fuels are designed so
that processing shuts down automatically if they overheat, thus allowing for
inherently safer reactors that are much simpler to operate relative to tradi-
tional reactors.
 Safer fuel allows for extensive or complete automation, which significantly
lowers operating costs.
 Well-supervised factory production of standard reactors attempts to drive
down unit costs by applying the fixed manufacturing facility costs over many
units and driving annual improvements in efficiency. This factory-built ap-
proach has been used to achieve dramatic cost improvements in the wind and
solar industries.

Increased demand for nuclear reactors and efficiency gains associated with the
corresponding manufacturing learning curve could lower the cost of nuclear reactors. The
petrochemical industry sector is capable of driving demand for these units for decades.
High-temperature reactors provide high-quality heat directly to industrial facilities, and
the integration of this heat will require chemical engineers working within an interdisci-
plinary team that understands process safety, integration, and intensification.
While existing as a separate discipline, nuclear engineering borrows heavily from
physics, as well as from mechanical and chemical engineering. Setting aside the particle
physics associated with fission and fusion reactions, a nuclear reactor is effectively a

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Decarbonization of Energy Systems 59

chemical reactor that takes in fuel as a feed, produces fission or fusion products as waste,
and produces heat as a product. Process integration and process design are necessary to
extract energy most efficiently from the steam that is generated by a nuclear reactor.
Viewed through the lens of the fundamental pillars of chemical engineering (transport
phenomena, reaction engineering, thermodynamics, and applied mathematics), the design
and safe operation of nuclear power plants are a good match for the skillset of a well-
trained chemical engineer. Optimal thermodynamics and heat transfer are key to an effi-
cient process design, as is process control to operate a power plant effectively and safely.
Further development of advanced reactor designs, as well as storage solutions for nuclear
waste, will also benefit from the same chemical engineering fundamentals. Advances in
nuclear energy present a clear opportunity for chemical engineers to collaborate with nu-
clear and other engineering disciplines.

ENERGY CARRIER PRODUCTION

Energy carriers are intermediates in the energy-supply chain, located between


primary and/or secondary sources (Thollander et al., 2020) and end-use applications. For
convenience and economy, energy carriers have shifted continually from solids to liquids,
recently from liquids to gases, and more recently to electricity, a trend that is expected to
continue and even accelerate to address climate change concerns. Currently, about one-
third of final energy carriers reach consumers in solid form (as coal and biomass), one-
third in liquid form (consisting primarily of oil products used in transportation), and one-
third through distribution grids in the form of electricity and gas. It is projected that the
share of all grid-oriented energy carriers could increase to about 50 percent of all con-
sumer energy by 2100 (Sims et al., 2007).
The following sections describe the opportunities for chemical engineers to con-
tribute to electricity generation and the production of low-carbon fossil fuels, hydrogen,
and synthetic fuels; production of advanced liquid biofuels was discussed previously in
this chapter.

Electricity

Reduction of GHG emissions during electricity generation, as well as electrifica-


tion of light- and medium-duty vehicles and residential/commercial heating, is crucial for
decarbonization of the energy sector in the near and medium terms.
Electricity generation from coal-fired plants, the largest CO2 emitters, has de-
creased in the United States since 2009, while electricity generation from both natural gas
and renewable energy sources has increased. In 2020, about 4,000 TWh of electricity was
generated at utility-scale electricity generation facilities in the United States, with about
60 percent of that total being generated from fossil fuels—coal, natural gas, petroleum,
and other gases; about 20 percent from nuclear energy; and about 20 percent from renew-
able energy sources (EIA, 2021b).
The shift from coal to natural gas and renewables was made possible by a steep
decline in the cost of key technologies associated with shale gas, wind power, solar power,

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

60 New Directions for Chemical Engineering

and grid-connected electricity storage (DOE, 2015). New wind and solar technologies
offer the lowest levelized cost7 of electricity over most of the Earth’s surface (IRENA,
2020). Since 2009, the levelized cost of wind has declined by 70 percent and that of solar
photovoltaics by almost 90 percent, providing an important means of supplying electricity
with no direct CO2 emissions (Lazard, 2019).
Recent decarbonization studies (e.g., IEA, 2021b; NASEM, 2021a) indicate that
deep decarbonization of electricity generation can be accelerated, but further innovation
is required. Chemical engineers are contributing to the scale-up, cost reduction, and reli-
ability of improved and novel technologies, particularly in the solar and wind energy sec-
tors. (Specific opportunities in these sectors were discussed earlier in this chapter.)

Low-Carbon Liquid Fossil Fuels

Liquid hydrocarbons from crude oil have been the preferred energy carrier for the
transportation sector because of their high energy density, easy distribution and storage,
low cost, and well-established and extensive infrastructure along the value chain. If they
are to play a role in the low-carbon energy mix of the future, however their carbon foot-
print will need to be significantly reduced (Figure 3-11).
In 2019, the global demand for liquid hydrocarbons was about 100 million barrels
per day, approximately 58 percent of which was for the transportation sector, 14 percent
for feedstock for chemicals, 12 percent for power generation/residential/buildings, and 16
percent for other industrial use (ExxonMobil, 2019). Demand is expected to increase until
at least 2040, although not uniformly across all sectors. Demand for hydrocarbon liquid
fuels for industrial use and for power generation, residential uses, and buildings is pro-
jected to decrease, being replaced by energy carriers from renewable sources. In the chem-
ical sector, demand for liquid hydrocarbons used as feedstock to manufacture consumer
products is expected to increase. In the transportation sector, overall demand is projected
to grow, but not uniformly across transportation types. Gasoline demand for light-duty
vehicles is projected to decrease as a result of greater market penetration of EVs, while
demand for liquid fuels in commercial transportation (heavy-duty, aviation, and marine)
is expected to increase, particularly in the heavy-duty long-haul sector. Changes in global
demand for oil, based on new policy scenarios, are projected to follow similar trends (IEA,
2021a).

Petroleum Refining

Most of the CO2 emissions associated with liquid fuels come from the use/com-
bustion of the fuel itself, but there are also opportunities to reduce emissions during oil
production, extraction, and refining. The majority of current refineries were designed and
optimized to manufacture primarily gasoline; diesel, aviation, and other heavy fuels and

7
Levelized cost is the sum of total lifetime costs divided by the amount of energy produced,
and thus represents the present value of the total cost of building and operating a power plant over
an assumed lifetime.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Decarbonization of Energy Systems 61

chemicals are normally secondary products. To maintain a low cost of production, the
amount of lower-cost heavy (high C/H ratio) and “dirty” crude oils in the feedstock mix
is maximized. Refineries will require significant reconfiguration not only to satisfy prod-
uct demand and meet challenges associated with GHG emissions, but also to maintain
low production costs for liquid fuels. Chemical engineering is already playing a central
role in addressing these challenges.

FIGURE 3-11 Energy density of transportation fuel types, indexed to gasoline = 1. The data points
represent the energy content per unit volume or weight of the fuels themselves, not including the
storage tanks or other equipment they require. For instance, compressed fuels require heavy storage
tanks, while cooled fuels require equipment to maintain low temperatures. SOURCE: EIA (2013).

Opportunities to reduce CO2 emissions during refinery operations include the fol-
lowing:

 Increase energy efficiency through further improvement of energy manage-


ment systems to ensure that refineries are run according to the most energy-
efficient standards.
 Replace combustion of liquid fuel for heat generation with renewable sources
(e.g., green electricity).
 Produce renewable (green) hydrogen with electrolyzers using imported or
self-generated renewable electricity.
 Use low-grade heat generated during operations to produce electricity for in-
ternal and external use.
 Eliminate flaring in refineries.
 Replace steam-driven rotating machines and fired heaters with electric coun-
terparts.
 Deploy CCUS from refinery flue gases.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

62 New Directions for Chemical Engineering

Opportunities for chemical engineers to contribute to reductions in the carbon


footprint of refinery feedstocks include the following:

 Use lighter and sweeter crude oil (higher H/C ratio and fewer heteroatom
contaminants, such as sulfur and nitrogen compounds) instead of carbon-in-
tensive and harder-to-process heavy oils (lower H/C ratio and more heteroa-
tom contaminants).
 Coprocess crude oil with biomass.
 Integrate bio- and oil and refineries.
 Produce and use renewable (green) hydrogen for hydroprocessing needs
and/or heat generation.
 Integrate electrofuels (e-fuels) within existing refineries to decrease low-
carbon liquid fuel production costs.
 Further integrate production of petrochemicals to optimize the product slate
(i.e., minimize production of gasoline/distillate and maximize that of petro-
chemicals and higher-quality lubricants).

Hydrogen

Hydrogen is a versatile energy carrier with significant potential to contribute to a


clean, low-carbon energy system. To realize this potential, chemical engineering and re-
lated fields can contribute in the following ways:

 Increase production from nonfossil sources.


 Significantly increase clean hydrogen production via water hydrolysis using
clean electricity.
 Develop a hydrogen infrastructure for distribution and storage at scale to sat-
isfy demand.

Globally, 96 percent of hydrogen is produced from fossil sources (48 percent nat-
ural gas, 30 percent liquid hydrocarbons, and 18 percent coal), with only about 4 percent
produced from electrolysis of water (IRENA, 2018). In 2018, the global demand for pure
hydrogen was above 70 million metric tons, and the demand for hydrogen as part of a
mixture of gases, such as synthesis gas, was about 45 million metric tons. The vast ma-
jority of this hydrogen (~95 percent) was used for production of chemicals (mainly am-
monia for fertilizers and methanol) and for oil refining (IEA, 2019). Hydrogen is currently
produced in large quantities, but a vast expansion of the world’s production capacity
would be needed for hydrogen to replace a significant fraction of oil in transportation.
Current hydrogen demand corresponds to ~8.4 EJ (2,333 TWh), or about 6 percent of the
annual global transportation energy demand (IRENA, 2021).
A growing number of countries have policies that directly support investment in
low-carbon hydrogen technologies, and global demand for hydrogen as an energy carrier
is projected to increase significantly after 2030 (IEA, 2020e). This increase is projected
mainly in sectors that are relatively more challenging to decarbonize and that have needs

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Decarbonization of Energy Systems 63

that cannot be met by electrification, including fuel for heavy-duty long-haul transporta-
tion; synthetic fuels for aviation and shipping, for which available low-carbon fuel options
are limited; ammonia as fuel for shipping; and a source for heat generation in the industrial
and buildings sector. Hydrogen is a promising option for storing renewable energy.
Today, hydrogen is produced primarily by steam reforming of natural gas; partial
oxidation (catalytic and not) and autothermal reforming of natural gas technologies are
also used to a lesser extent. In some countries, particularly in Asia, gasification is used
commercially for hydrogen production from coal. Production by water electrolysis is also
a commercial technology, but at a much smaller scale and higher cost. Methane pyrolysis,
or methane splitting of natural gas using renewable electricity to produce hydrogen and
black carbon, is currently in the commercial demonstration scale.8 Many other routes to
low-carbon hydrogen, such as thermochemical water splitting, direct photocatalysis, and
biological production from microorganisms, are in various stages of R&D.
It is generally accepted that hydrogen from fossil fuels without CCUS will remain
the main source of hydrogen production. After 2030, it is estimated that almost all of the
growth in hydrogen production will come from low-carbon hydrogen (IEA, 2020d) made
from renewables-based electricity or from fossil fuels, particularly natural gas, in combi-
nation with CCUS. By 2070, electrolytic hydrogen is projected to account for nearly 60
percent of global hydrogen production (IEA, 2020d).
Hydrogen produced from different feedstocks is identified by colors. “Black,”
“gray,” and “brown” refer to hydrogen produced from coal, natural gas, and biomass,
respectively. “Blue” is commonly used for hydrogen produced from fossil fuels, with CO2
emissions reduced by the use of CCUS. “Green” hydrogen is produced from water elec-
trolysis using renewable electricity.
Blue and green hydrogen have a path to competitiveness with gray hydrogen (Hy-
drogen Council, 2019). The competitiveness of blue hydrogen depends primarily on scale-
up of CCUS facilities and the value attributed to sequestered CO2. Carbon prices or taxes
of USD 50/ton—a figure consistent with near-term milestones of major economies with
net-zero commitments (e.g., in the European Union by 2030; Argus, 2020)—would make
blue hydrogen competitive. The competitiveness of green hydrogen will require steep cost
reductions for electrolyzers, as well as reductions in renewable energy costs. The produc-
tion of green hydrogen is projected to break even with that of gray hydrogen before 2030
in regions with low renewable-energy costs, and before 2035 in regions with average re-
newable-energy costs (Hydrogen Council, 2019). A combination of green and blue hy-
drogen production pathways will be required to satisfy the potential demand for low-car-
bon hydrogen. Their supply mix will depend on a range of technical and societal factors,
production costs, existing infrastructure (such as power supply and transmission net-
works), and emerging hydrogen trade routes.
Many challenges and opportunities related to the production of low-carbon hy-
drogen can be addressed by chemical engineers. For blue hydrogen, the challenges and
opportunities center on improved CCUS (i.e., at lower cost) and assessment of geological

8
See www.monolithmaterials.com.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

64 New Directions for Chemical Engineering

sites for long-term CO2 storage; for green hydrogen, they center on the need for a source
of sustainable, low-cost, renewable electricity and access to freshwater.
Larger-scale electrolysis plants, whose development will depend on achieving
lower costs and improved electrical efficiency for the electrolyzers, are also needed. Three
main electrolyzer technologies exist today: alkaline electrolysis, proton exchange mem-
brane (PEM) electrolysis, and solid oxide electrolysis cells (SOECs). Alkaline electrolysis
is a mature, commercial-scale technology with relatively low capital costs. PEM electro-
lyzers use pure water as an electrolyte solution and avoid the recovery and recycling of
electrolyte solution necessary with alkaline electrolysis; however, they use expensive
electrode catalysts and membrane materials. Lifetimes of PEM electrolyzers are currently
shorter than those of alkaline electrolyzers, and overall costs are higher. SOECs are the
least-developed electrolysis technology and not yet commercialized. Ceramics serve as
the electrolyte, resulting in lower material costs. Because steam is used for electrolysis,
SOECs require a heat source. If the hydrogen produced were to be used for the production
of synthetic hydrocarbons (in power-to-liquid or power-to-gas schemes), the waste heat
from these synthesis processes (e.g., Fischer-Tropsch synthesis and methanation) could
be recovered to produce steam for further SOEC electrolysis (IEA, 2019).

Synthetic Fuels

Synthetic fuels are produced by converting hydrogen and a carbon source into
compounds that can be used as energy carriers, such as methane; methanol; ethanol; and
such higher-carbon-number products as gasoline, diesel, and aviation fuel. Ammonia is
also increasingly seen as a non–CO2-generating synthetic fuel, although ammonia pro-
duction itself is currently a major source of CO2 emissions. Electrification and advances
in electrochemical routes to ammonia synthesis may increase the potential for ammonia
as a promising fuel. Low-carbon or carbon-neutral synthetic fuels require green hydrogen
and carbon from a bioenergy source or from CO2 captured from flue gases or the atmos-
phere using direct air capture (DAC) technologies. In the case of DAC, however, signifi-
cant energy and technology advances will be necessary to make this route viable. The
low-carbon or carbon-neutral synthetic fuels are referred as power-to-fuels (PtF) or elec-
trofuels (e-fuels).
The production of e-fuels requires significant amounts of electricity, and from a
thermodynamic point of view, the electricity should be used directly. For example, 25
kWh of energy is required to produce 1 L of synthetic kerosene from electrolytic hydrogen
together with CO2 captured through DAC. More than 80 percent of the energy is used to
produce hydrogen; around 15 percent is used for the capture of CO2 through DAC; and
the rest is used in the Fischer-Tropsch synthesis. Currently, only about 40 percent of the
energy input is stored in the final liquid product, although with process optimization, the
overall conversion efficiency could potentially increase beyond 45 percent (IEA, 2020d).
However, e-fuels have the potential to help some sectors—particularly the aviation trans-
portation sector, which is difficult to decarbonize and for which electrification is not an

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Decarbonization of Energy Systems 65

option because of high energy-density requirements for the energy carrier. These consid-
erations would need to be important enough to justify the thermodynamic inefficiencies,
however.
A few aviation e-fuel demonstration plants have been announced. For example,
the Norsk e-fuel project is planning a first plant in Herøya, Norway. That plant is expected
to become operational in 2023, with a production capacity of 10 million L annually, scal-
ing up to 100 million L annually by 2026.9 Cost reduction for e-fuel production is expected
to benefit from the economy of scale and experienced manufacturing learning curve. It
should be emphasized that DAC technology is in its infancy, with substantial room for
technology and process improvement and cost reduction. For e-fuels to be competitive
with conventional fossil fuels and even bio–jet fuel, a combination of low electricity costs
and a high CO2 tax or cost will be needed.

ENERGY STORAGE

Energy storage occurs when an energy carrier is held in a fixed location until it
can be deployed. When the energy carrier uses chemical bonds or potential energy, as is
the case for liquid fuels and pumped hydropower, respectively, energy storage is concep-
tually simple. When the energy carrier is electrons, storage requires batteries or capacitors
or conversion of the electrical energy into another energy carrier. Because many of these
are mature technologies, chemical engineers have limited potential to impact some modes
of energy storage, especially those involving mechanical energy, such as pumped hydro-
power and compressed gas storage. For energy carriers that involve electrons or chemical
bonds, however, chemical engineers can play a critical role in the development and de-
ployment of scalable and economical energy storage.
The changing nature of the world’s energy system has continued to create signif-
icant demand for energy storage, and this trend is likely to accelerate in the coming years.
One obvious example involves intermittent renewable electricity sources such as solar
and wind, whose full contribution to decarbonization of electricity generation can be re-
alized only if they are coupled with grid-scale storage. A second example is the trend
toward electrification of light-duty vehicles, which will require massive deployment of
on-board batteries.
The properties of an energy storage system can vary with both scale and applica-
tion. In vehicle applications, for example, the weight and size of batteries are critical; in
grid-scale storage, by contrast, the battery weight is unimportant. The typical storage time
and desired charging and discharging rates are also very different for these two applica-
tions. This example illustrates why future energy storage needs will be met by a wide
array of technologies—there is no one-size-fits-all approach to energy storage. The rapid
deployment of electrical storage systems is underpinned by the commercial sector’s large
investments in technology development. If academic research is to have an impact in this
environment, researchers will have to seek solutions in frontier areas in which well-

9
See https://www.norsk-e-fuel.com/en.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

66 New Directions for Chemical Engineering

funded development efforts in industry are less likely to overwhelm the scale of any aca-
demic effort.
An important distinction between energy storage in electrons and chemical bonds
lies in the number of times a particular piece of storage medium can be cycled. For fuels
based on chemical bonds, storage simply involves a tank, and little to no change in the
storage medium is expected even after thousands of cycles. Here, chemical engineers have
a role to play in understanding evaporation and erosion, as well as improving leak detec-
tion methods. Batteries, however, tend to degrade during cycling. Many of these degra-
dation processes stem from chemical reactions or nucleation and growth of phases that
are driven by fundamental chemical engineering principles. This observation indicates
that chemical engineers can play a key role in developing concepts that increase battery
lifetimes. At the same time, for batteries to be a core part of a truly sustainable energy
system, their end-of-life disposal or recycling needs to be considered. Similarly, the global
availability of battery components (e.g., lithium) is an important consideration in LCA
comparisons of competing technologies. The intentional design of batteries for end-of-
life disposal and the use of earth-abundant elements are areas in which chemical engi-
neering researchers are poised to play an important role.
As discussed in the previous section, hydrogen is a versatile energy carrier with
significant potential to contribute to a clean, low-carbon energy system. Given its low
energy density, however, its long-distance distribution and storage poses challenges, par-
ticularly if it needs to be shipped overseas. It has been estimated (IEA, 2019) that for
distances above ~1,000 miles, shipping hydrogen as ammonia or liquid organic hydrogen
carriers (LOHCs) is likely to be more cost-effective than shipping liquified hydrogen.
Chemical engineers have opportunities to significantly lower the cost of conversion be-
fore export and reconversion back to hydrogen in the case of LOHCs, and before con-
sumption or the direct use of ammonia as fuel.
Vast sectors of the global energy system rely on liquid or solid energy carriers,
such as gasoline and coal. The energy density and ease of transporting these carriers rel-
ative to gases or electrons give them strong intrinsic advantages. Chemical engineering
played a central role in what might be termed the hydrocarbon economy of the 20th cen-
tury, and the discipline will also be central to the deployment of more sustainable and
environmentally benign liquid energy carriers in the 21st century. Unlike high-value-
added products such as pharmaceuticals, new or improved liquid energy carriers need to
be competitive with low-cost alternatives and deployable at massive scale in order to be
viable. In many cases, achieving these goals will require integrated process development
rather than a singular focus on improved catalysts or similar chemical steps. The ability
of chemical engineers to use LCAs and technoeconomic assessments (TEAs) to focus
research on approaches with a plausible path to economic viability ensures that their con-
tributions will have impact in the domain of energy storage.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Decarbonization of Energy Systems 67

ENERGY CONVERSION AND EFFICIENCY

Ultimately, whatever the source and carrier, energy is converted to heat or work,
commonly referred to as energy use or consumption. This section is organized by appli-
cation within the transportation, industry, residential, and commercial sectors, and focuses
on opportunities for chemical engineers to contribute to energy efficiency and decarbon-
ization in those sectors.

Transportation Sector

Transportation is a large and diverse sector that encompasses road (passenger and
freight vehicles), aviation, marine, and rail transportation. In 2018, the transportation sec-
tor accounted for nearly a quarter of global CO2 emissions (IEA, 2021a), and efforts to
decarbonize the transportation sector are therefore critical to achieving the goals of the
Paris Agreement. As noted previously, the transition to a net-zero emissions transporta-
tion sector will take decades and cost hundreds of billions of dollars, and may never be
complete (Ogden et al., 2016). Net-zero emissions aside, just reducing transportation CO2
emissions significantly in the coming decades is a formidable challenge. Tackling this
challenge will require structural shifts in the transportation of people and freight, much
larger gains in energy efficiency, major advances in technology, effective government
policies, significant levels of investment in infrastructure for low-carbon energy carriers,
and the manufacture of low-carbon and zero-emission vehicles.
No single energy carrier can, in the foreseeable future, satisfy the requirements
across all segments of the transportation sector (EPA, 2021a). Some segments of the sec-
tor are easier to decarbonize (light- and medium-duty vehicles) than others that require
high-energy-density fuels (heavy-duty long-haul, aviation, marine). A “bridging” low-
carbon energy transition strategy will rely on the combined increased use of such energy
carriers as electrons, hydrogen, and low-carbon liquid fuels, particularly advanced biofu-
els and synthetic liquid fuels (Santiesteban and Degnan, 2021). This section focuses on
opportunities for chemical engineers to contribute to technology improvements in EVs,
hydrogen-powered fuel cell engines, and internal combustion engines.

Electric Vehicles

The global EV market has grown over the past decade as a result of both techno-
logical advances and supportive government policies. About 7.2 million EVs were in use
in 2019, and this number is predicted to increase to nearly 140 million within the next
decade (IEA, 2020a). Many automobile manufacturers have announced plans to stop pro-
duction of internal combustion engines by 2030. The key technological enabler of the
growth in EVs is the progress made in lithium-ion batteries, in terms of both improved
performance and cost reduction (BloombergNEF, 2020; Figure 3-12). To accelerate mar-
ket penetration of EVs, several remaining challenges for lithium-ion batteries need to be
overcome, many of which chemical engineers are poised to help address, as described
below.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

68 New Directions for Chemical Engineering

FIGURE 3-12 Survey of lithium-ion cell and pack integration prices for automotive years 2013–
2020, adjusted to real 2020 $/kWh. SOURCE: BloombergNEF (2020).

The chemical industry has played a critical role in the successful development
and manufacturing of lithium-ion–based batteries and continues to drive innovation to
meet booming demand. Improvements are being pursued in mining, metals processing
and purification, battery design and manufacturing, battery chemistries, and performance.
The advances needed for lithium-ion batteries in automotive applications are de-
scribed by Masias and colleagues (2021):

 further increases in energy density per unit weight and volume;


 continued cost reduction on a $/kWh basis, which is both challenging and
necessary to increase the opportunity for wide-scale adoption of electric
transportation;
 the ability to validate and predict long battery life quickly and accurately;
 fast recharging while preserving long battery life and overall safety; and
 direct recycling of cathodes and anodes to reduce both the costly geological
extraction and initial material processing steps of rare elements.

Lithium-ion–based batteries are approaching the theoretical energy density limit


imposed by their inherent chemistry; therefore, they may be unable to satisfy future EV
demand, and several next-generation lithium batteries are being investigated (Wu et al.,
2020), the three main technologies being lithium-air, lithium-sulfur, and lithium-metal.
Despite some research progress, however, development and scale-up have been slow. The
lithium-air battery has faced battery life and energy efficiency challenges. The lithium-
sulfur battery has shown more promise than the lithium-air, but it, too, has been limited
by life and volume concerns because of its low energy density. Lithium-metal batteries
have advanced the furthest in the past decade but are still in the development stage (Masias
et al., 2021).

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Decarbonization of Energy Systems 69

All of the lithium battery types described above use a liquid electrolyte. One strat-
egy for improving overall performance is the use of a solid electrolyte material. Such
batteries are referred to as solid-state or all-solid-state batteries (ASSBs). The discovery
of highly conductive solid-state electrolytes has led to tremendous progress in the devel-
opment of ASSBs (Tan et al., 2020a), making it possible to overcome such technical chal-
lenges as poor interfacial stability, scalability, preservation of high energy density, pro-
duction safety, and cost reduction.
Solid-state batteries with lithium-metal anodes have the potential to achieve high
energy densities. Several start-ups, many in close collaboration with large automakers,
are focusing on the development of technologies required to optimize solid-state electro-
lytes. Substantial increases in the energy density of solid-state batteries, together with
improvements in other performance metrics, such as cost and durability, could make elec-
trification a viable and more attractive commercial option for medium- and heavy-duty
regional and long-haul vehicles. These advances could also have a significant impact on
short-distance and small-freight marine and aviation transportation.

Hydrogen-Powered Fuel Cell Vehicles

The two primary options for zero-emissions transportation are electric drivetrains
powered by batteries (battery electric vehicles [BEVs]) and by hydrogen fuel cells (fuel
cell electric vehicles [FCEVs]). Both are used for light-, medium-, and heavy-duty vehi-
cles. These two technologies have complementary strengths and meet different applica-
tion and customer needs. BEVs are emerging as the technology of choice for light- and
medium-duty vehicles, while FCEVs are better suited for heavy-duty commercial vehicles
that are driven long distances, carry heavy loads, and require relatively quick refueling.
PEM fuel cells are used in FCEVs because they offer higher power density, lower
overall weight, and lower total volume compared with other fuel cell types (NRC, 2015).
They are also easily scaled for different vehicle classes. Because these cells operate at
fairly low temperatures (<100 °C), they have a short warm-up time and better durability
relative to other fuel cell types. These features also contribute to much greater overall
efficiency compared with high-temperature solid oxide or molten carbonate fuel cells in
a frequent start-up/shutdown vehicle application.
In a PEM fuel cell, the anode–separator–cathode structure is known as the mem-
brane electrode assembly (MEA). It consists of a solid polymer electrolyte membrane
(e.g., Nafion) with a catalyst layer and a gas diffusion layer on each side. The catalyst
layers are platinum-group metal nanoparticles (usually platinum or platinum alloy) on a
porous carbon/ionomer support. The gas diffusion layers are multilayer carbon fiber/pol-
ytetrafluorethylene paper sheets that facilitate mass transfer of reactants and products. The
MEA is sandwiched between bipolar plates that have channels for gas flow and conduct
the electric current. Like batteries in an EV, the fuel cell unit in the vehicle is actually a
stack of many individual fuel cells arranged in series to provide sufficient voltage. The
stack is rated by maximum power output.
The key challenges and opportunities to which chemical engineers are contrib-
uting in this domain are as follows:

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

70 New Directions for Chemical Engineering

 Reducing costs—The cost of platinum-based catalysts is the primary driver


of fuel cell costs, a fact that has motivated research on technologies for re-
ducing the platinum content of MEA catalysts and even developing platinum-
free MEA architectures.
 Increasing durability—The on-road durability of PEM fuel cells is currently
less than desirable for large-scale commercial deployment. The primary
cause of degradation is loss of catalyst activity (e.g., due to sintering) and
deterioration of the PEM. One important lever for increasing durability is to
increase catalyst loading (i.e., add more platinum), but doing so increases
cost. A better alternative is to design sintering-resistant catalysts or replace
platinum with lower-cost metals.
 Increasing efficiency—PEM fuel cells for on-road FCEVs have a peak power
efficiency of up to 60 percent in terms of converting the available fuel energy
of hydrogen into electrical energy, well below the theoretical maximum effi-
ciency of approximately 80 percent.

Hydrogen is stored on the vehicle as a compressed gas at a pressure of 700 bar in


one or two carbon-fiber composite tanks. Since even highly compressed hydrogen has a
much lower energy density than gasoline and diesel fuel, the volume of the tank(s) needs
to be large enough to enable a driving range comparable to that of conventional internal
combustion engine (ICE) vehicles. Research, development, and demonstration (RD&D)
in this area targets primarily the cost of the composite tanks (e.g., cheaper carbon fibers
or other materials and cheaper construction methods) and auxiliary components. Early-
stage research also is being conducted on other technologies for storing hydrogen on
FCEVs, such as

 adsorbents (e.g., metal organic framework materials [MOFs] and metal hy-
drides) that would allow lower hydrogen pressures;
 cold/cryocompressed conditions in insulated tanks (e.g., −75 °C/500 bar or
−235 °C/70 bar); and
 chemical storage, in which a hydrogen-dense chemical (e.g., NH3BH3,
methylcyclohexane) is loaded into the tank, H2 is disassociated through heat
or chemical reaction, and then spent chemical is collected and regenerated at
a central facility.

Without a breakthrough, vehicle manufacturers would be unlikely to opt for these


alternatives in the foreseeable future because of the need to develop additional technolo-
gies and/or fueling infrastructures to accommodate them.

Internal Combustion Engine Powertrain Vehicles

For both heavy- and light-duty ICE powertrain vehicles, chemical engineers can
play important roles in advancing technology for increasing fuel efficiency, thereby re-

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Decarbonization of Energy Systems 71

ducing CO2 emissions. Advances in ICE technology have traditionally been led by me-
chanical engineers, making this a key opportunity for chemical engineers to collaborate
with another engineering discipline. Areas for such collaboration include

 advanced combustion schemes (e.g., low-temperature combustion);


 electrified accessories and waste-heat recovery;
 hybridization;
 lightweighting, with new, aerodynamic cabs and trailer components;
 advanced communication and logistics;
 energy-efficient tires and monitoring systems; and
 advanced GPS-based predictive control technology.

GHG emissions could be reduced in ICE-powered heavy-duty vehicles by the use


of low-carbon liquid fuels. Codevelopment of more efficient fuels and vehicle engines is
a major collaborative opportunity for chemical and mechanical engineers, with chemical
engineers bringing expertise in reaction mechanisms and transport and mechanical engi-
neers bringing expertise in transport, computational fluid dynamics, and engine design.
Opportunities in this area include

 blends of high-quality, low-carbon fossil-diesel fuel with advanced biofuels


and/or synthetic fuels (e.g., e-fuels) to achieve specific tailpipe GHG emission
levels; and
 increased control of the variability of diesel-fuel properties.

Industrial Sector

The industrial sector produces the goods and raw materials people use every day.
Industry is energy intensive—nearly half of the world’s energy is dedicated to industrial
activity; accordingly, it is also responsible for a large portion of global CO2 emissions.
Production of cement, steel, and chemicals accounts for the largest portion of industrial
CO2 emissions—about 70 percent (IEA, 2020d). Reducing manufacturing-related emis-
sions in the United States and China will be critical to reducing CO2 global emissions from
manufacturing (Figure 3-13).
The lack of commercially available and scalable low-carbon alternatives to fossil
fuels makes deep reductions in CO2 emissions from industry highly challenging in the
short and medium terms. This fact is reflected in the Sustainable Development Scenario
projections (IEA, 2020d), in which the industrial sector emerges as the second-largest CO2
emitter in 2070, after the transportation sector, accounting for around 40 percent of resid-
ual emissions, even though its emissions are projected to be 90 percent lower overall than
in 2019 (Figure 3-13). Currently, energy inputs to the industrial sector are approximately
70 percent from fossil fuels (IEA, 2020d). In the Sustainable Development Scenario pro-
jections, the use of fossil fuels in industry would be reduced by more than 60 percent by

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

72 New Directions for Chemical Engineering

FIGURE 3-13 Projected decrease in global direct CO2 emissions of industry by subsector (paper,
aluminum, cement, steel, chemicals, other) and region (United States, European Union, China,
India, rest of world [ROW]) in the International Energy Agency’s Sustainable Development Sce-
nario, 2019–2070. SOURCE: IEA (2020d).

2070, being replaced primarily by electricity and bioenergy, while more than 75 percent
of the remaining CO2 emissions would be captured and stored permanently. The following
sections focus on opportunities for chemical engineers in the cement, steel, and chemical
production subsectors, as well as cross-cutting approaches for decarbonization of the in-
dustrial sector.

Cement Production

More than 70 percent of the energy used in the U.S. cement industry comes from
coal and petroleum coke. More than 50 percent of the total CO2 emissions attributable to
cement production are process related (from calcination of limestone in the kiln), not en-
ergy related. To advance decarbonization of the cement industry, in addition to energy-
efficiency improvements, these process-related CO2 emissions will need to be reduced.
Key strategies for deep decarbonization of the cement industry are clinker substitution
(supplementary cementitious materials [SCMs]), a switch to lower-carbon fuels, CCUS,
low-carbon cement and concrete chemistries, and kiln electrification. While several of
these strategies can be combined to approach net-zero emissions, decarbonization efforts
related to demand reduction, use of SCMs, and waste-carbon utilization will also be
needed. For example, Hasanbeigi and Springer (2019) recently showed that, compared
with 2015, total CO2 emissions from California’s cement industry could decrease by 68
percent by 2040 even though the state’s cement production is projected to increase by 42
percent (Figure 3-14).

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Decarbonization of Energy Systems 73

FIGURE 3-14 Impact of various options for CO2 emissions reduction in California’s cement in-
dustry. NOTES: BAU = business as usual; CCUS = carbon capture, use, and storage. SOURCE:
Hasanbeigi and Springer (2019).

Further RD&D is necessary to decarbonize cement processes, including technol-


ogies to

 improve concrete performance and durability using alternative raw materials


while meeting construction codes and standards, especially for natural SCMs
(e.g., pozzolans, calcined clay);
 develop innovative ways of using large-scale nonpurified CO2 for different
applications and product streams with low energy penalty and cost;
 develop electrified kilns capable of operating at very high temperatures suit-
able for producing cement;
 advance the use of waste biomass and green hydrogen in cement kilns and
improve understanding of resulting effects on the kiln and final product;
 achieve high carbon-capture efficiency (more than 90 percent) on retrofitted
and new cement plants;
 develop better catalysts and process designs to deliver higher efficiency lev-
els, reduce costs, and lower material consumption or waste production for
CCUS in the cement plant; and
 improve CO2 transportation and storage infrastructure for CCUS.

Steel Production

Around 70 percent of steel in the United States is produced by electric arc fur-
naces; the remainder is produced by blast furnaces (BFs) or basic oxygen furnaces. Key
technologies needed to decarbonize the steel industry include improvements in energy
efficiency; a switch to low-carbon fuels; use of green hydrogen instead of natural gas in
direct reduction of iron (DRI); production of iron by electrolysis of iron ore using only
renewable electrical energy; postcombustion CCUS (such as top-gas recycling in BFs);
DRI with CCUS; HIsarna (smelting reduction) with CCUS; carbon utilization (carbon to

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

74 New Directions for Chemical Engineering

ethanol or chemicals); and green hydrogen plasma smelting reduction. Although some of
these decarbonization technologies have been commercialized, some require further
RD&D, such as

 industrial-scale plant design for producing iron by electrolysis, and develop-


ment of detailed cost models for assessing the commercial viability of the
process; and
 development of methods to analyze and evaluate the BF operation continu-
ously when extending the use of hydrogen to all tuyeres (injection nozzles for
air/H2) in BFs.

Chemical Manufacturing

The chemical industry is highly diverse, producing more than 70,000 products
globally. Yet 18 large-volume chemicals—including light olefins, ammonia, BTX (ben-
zene, toluene, xylene) aromatics, and methanol—account for 80 percent of the energy
demand and 75 percent of the total GHG emissions attributable to global chemical man-
ufacturing (IEA, ICCA, and DECHEMA, 2013). For perspective, the global production
volumes in 2012 were 220 million metric tons for ethylene and propylene, 198 million
metric tons for ammonia, 58 million metric tons for methanol, and 43 million metric tons
for benzene. Together, production of these four products used about 7.1 EJ of energy per
year (or a specific energy consumption of about 13.7 MJ/kg of product), and the top 18
large-volume products used a total of about 9.4 EJ of energy per year (Figure 3-15).
Because about 90 percent of chemical manufacturing processes use catalysts, cat-
alyst and catalyst-related process improvements could reduce the energy intensity of these
18 products by 20–40 percent by 2050, amounting to energy savings of about 13 EJ and
a CO2 emissions reduction of 1 metric gigaton (IEA, ICCA, and DECHEMA, 2013). In-
cremental improvements will suffice in the short to medium term; in the longer term,
however, the deployment of new technologies, such as biomass feedstocks and green hy-
drogen, will be necessary. The challenges for chemical engineers are clear: to identify top
catalyst and catalyst-related process opportunities for these 18 and other chemicals.
Key technologies for decarbonization of the chemical industry include energy
systems engineering; fuel switching; novel catalytic processes (e.g., olefin production via
catalytic cracking of naphtha or renewable methanol); advanced separation processes;
electrification; transitions to low-carbon feedstocks and processes (e.g., biomass, hydro-
gen-based production of ammonia and methanol, artificial photosynthesis, renewable-en-
ergy electrochemistry, biobased plastics production, gas-to-liquids gas); and CCUS.
Chemical engineers can contribute to the RD&D necessary to achieve these technological
advances, including the following examples:

 Achieve viability for natural gas crackers and improve catalyst production;
improve the efficiency of olefin production via catalytic cracking of naphtha
or via methanol; and address environmental issues for used catalysts, includ-
ing regeneration and catalyst selectivity and lifetime.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Decarbonization of Energy Systems 75

FIGURE 3-15 (a) Global energy consumption versus production volumes, and (b) global green-
house gas (GHG) emissions versus production volumes for the top 18 large-volume chemicals in
2010. NOTE: BTX = benzene, toluene, xylene. SOURCE: IEA, ICCA, and DECHEMA (2013).

 Improve the hydrogen peroxide propylene oxide process—oxidation of pro-


pylene with hydrogen peroxide yields propylene oxide and water as a by-
product; its energy consumption could be about 35 percent lower than that of
the traditional process.
 Resolve waste management issues for selective membranes, including mem-
brane washing/cleaning, and drive step-change advances in separations, in-
cluding the use of ceramic membranes.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

76 New Directions for Chemical Engineering

 Reduce technical and economic hurdles for such low-carbon processes as an-
aerobic digestion and gasification, biobased production of plastics, and hy-
drogen-based production of ammonia and methanol.
 Advance processes related to water electrolysis, such as optimized processes
for variable operation, improved stability for operations under pressure (30–
40 bar), electrodes with low-content noble metals and other rare elements,
and photocatalytic water splitting (e.g., non–noble metal electrodes, corro-
sion-resistant photoelectrode materials improved over potential).
 Reduce technical hurdles for fuel switching to hydrogen and other lower-car-
bon fuels, electrification, and CCUS.

Petroleum Refining

The United States is one of the largest producers of liquid transportation fuels and
refined petroleum products in the world, and energy engineering could reduce the fuel
used in producing these products by 50 percent (Morrow et al., 2015). Key technologies
for decarbonization of petroleum refining include fuel switching, electrification, biomass
hydrothermal liquefaction and biocrude oil, synthetic fuel synthesis, reduction or elimi-
nation of flaring, selective membranes, advanced control and improved monitoring,
CCUS, catalytic cracking, progressive distillation, self-heat recuperation, and biodesulfu-
rization. Further RD&D is critical for these technologies, and chemical engineers can
make contributions to

 advance synthetic fuel synthesis;


 integrate biofeedstocks;
 scale up technology for conversion of CO2 and H2 to hydrocarbons with lower
electricity consumption;
 improve chemical separations with lower energy demand for selective mem-
branes;
 apply advanced modular nuclear reactors for low-temperature steam genera-
tion; and
 improve electric heating to achieve high temperatures and large scales effi-
ciently and economically.

Food and Beverage Industry

The food and beverage industry is the third-largest consumer of energy in the
United States; its energy consumption is dominated by mechanical systems, compressed
air, refrigeration, and process heat in the moderate-to-low temperature range. Key tech-
nologies for decarbonization of the food and beverage industry include fuel switching to
lower-GHG sources, such as electricity and renewables, and plant efficiency measures,
particularly because many of these plants tend to be operated by small- and medium-sized
manufacturers with limited energy-management capacity. Electrification of dewatering,

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Decarbonization of Energy Systems 77

drying, and process-heating applications using heat pumps, hybrid boilers, induction heat-
ing, dielectric heating, and advanced cooling/refrigeration represents an important oppor-
tunity for this subsector. In addition, advanced processing and preservation to reduce deg-
radation in processing, along with improvements at the supply chain and consumer levels,
are important because on a global scale, about a quarter of the food supply is wasted (see
Chapter 4; Buzby et al., 2014; D’Odorico et al., 2018; Finley and Seiber, 2014). A key
challenge for this subsector is that food and beverage processors must comply with mul-
tiple regulations that complicate the implementation and slow the adoption of new tech-
nologies. Although some decarbonization technologies are commercially available, fur-
ther RD&D is needed in such areas as

 shifting from steam and fossil fuels to electric and solar heating technologies;
 demonstrating and certifying alternative processing technologies to reduce
GHG emissions while maintaining product safety and quality, and reducing
degradation in the supply chain;
 using waste to produce bioenergy; and
 adopting fuel switching and expanded implementation of energy efficiency.

Cross-Cutting Approaches for Decarbonization of the Industrial Sector

To achieve the net-zero emissions goal for industry, five broadly applicable de-
carbonization pillars require vigorous pursuit in parallel over the next decades: demand
reduction, energy efficiency, fuel switching and electrification, transformative technolo-
gies in sectors, and abatement. The applicability and selection of these pillars will vary
across sectors, with weighing of trade-offs in costs and accessible resources (de Pee et al.,
2018). Except for sector-specific transformative technologies, these approaches can be
considered cross-cutting and can facilitate reduction of GHG emissions across multiple
sectors.
Barriers abound across the landscape of deployment, development, scale-up, and
whole-system integration of current, emerging, and transformative technologies that can
advance these cross-cutting approaches. Nonetheless, chemical engineers have opportu-
nities to help overcome these barriers (e.g., in the areas of competitiveness, carbon capture
and use, and advanced materials). Many low-carbon technologies are in the early stages
of development and will require extensive RD&D for effective deployment. RD&D needs
range from advancing cross-cutting technologies (e.g., improved electrolysis of water to
lower-cost H2) to making radical changes (e.g., applying high-temperature heat for ethane
crackers). Figure 3-16 summarizes a more comprehensive set of recommendations from
a recent study by the National Academies of Sciences, Engineering, and Medicine on low-
carbon technologies, approaches, and infrastructure needing RD&D investment in the
2020–2050 period (NASEM, 2021a). A portfolio of collaborative RD&D initiatives, mul-
tigeneration plans, agile management, and durable support will be needed to face these
challenges successfully and drive progress going forward (NASEM, 2019f, 2021b).

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

78 New Directions for Chemical Engineering

FIGURE 3-16 Research, development, and demonstration investment needs to advance low-car-
bon technologies (carbon capture, use, and storage [CCUS]; low-carbon fuels; electrification; and
energy efficiency [EE]) and achieve decarbonization over the period 2020–2050. NOTES:
CHP/WHP = combined heat and power/waste heat to power; DAC = direct air capture; HT = high-
temperature; LCA = life-cycle assessment; SEM = strategic energy management. SOURCE:
NASEM (2021b).

Commercial and Residential Sectors

The impact of the commercial and residential sectors on net energy use and GHG
emissions in the United States is similar in scale to that of the transportation sector. Alt-
hough significant opportunities exist for improving energy efficiency in commercial and
residential settings, the highly dispersed nature of this sector creates challenges not shared
by the industrial subsectors discussed in the previous section, which have large, fixed
facilities. A recent National Academies report on decarbonization identifies key strategies
that employ existing technologies, including electrification of energy use and dramatically
increased use of electric heat pumps for heating and hot water (NASEM, 2021a). The
future energy needs of residential buildings in the United States and elsewhere in the
world are quite different. In the United States, the critical need is for more efficient use
of energy for heating and air conditioning as existing systems for these purposes are re-
placed or updated. Low- and middle-income nations, on the other hand, are likely to see
enormous growth in the use of air conditioning, driven by rising prosperity and, to a lesser
extent, by the impact of climate change.
To contribute to significantly reducing energy use in the commercial and residen-
tial sectors, chemical engineers will need to work closely with other engineers, including
mechanical and civil engineers, who are more closely affiliated with these sectors, in the

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Decarbonization of Energy Systems 79

development of technologies that can drive progress. As is the case with other topics dis-
cussed in this chapter, putative technologies will need to be deployable at low cost with
high reliability to have any chance of success.
Although systems for refrigeration and air conditioning are ubiquitous, improve-
ments in their efficiency and sustainability will require overcoming key technological
challenges. Chemical engineers can play a pivotal role in the development of new heat-
transfer fluids that are nontoxic and nonflammable and lack the very high GHG intensity
of many chlorofluorocarbons (CFCs) and other halocarbons. Nontraditional heat-transfer
cycles, such as adsorption cooling, and materials with caloric properties (e.g., Moya and
Mathur, 2020) also have considerable potential that intersects strongly with chemical en-
gineering applications in other domains.
Improving the properties of building materials is a key path toward improving
commercial and residential energy efficiency. Opportunities include the development of
passive materials, such as paints for roofs that reflect the solar spectrum more effectively
and, as a result, reduce the cooling load required for buildings (Li et al., 2021a); and active
materials, such as “smart glass,” which adapts to external sunlight to reduce the net energy
demands in buildings (Alias et al., 2019). Chemical engineers have many opportunities to
contribute creatively in these areas, provided that researchers remain focused on meeting
the cultural and economic needs of end users rather than on fashioning “pure technology”
solutions.

CARBON CAPTURE, USE, AND STORAGE

CCUS will be centrally important to controlling the concentration of carbon in


the atmosphere. Achieving net-zero emissions to halt growth in the concentration of at-
mospheric CO2 will not require achieving zero anthropogenic CO2 emissions, but rather
balancing anthropogenic CO2 emissions with natural and anthropogenic CO2 sinks. An
extensive portfolio of mitigation strategies for GHG emissions could contribute to the
achievement of net-zero emissions (Figure 3-17). Six negative emissions technologies
(NETs) remove carbon from the atmosphere and sequester it (Fuss et al., 2018; NASEM,
2019b): coastal blue carbon, terrestrial carbon removal and sequestration, bioenergy with
carbon capture and sequestration, DAC, carbon mineralization, and geological sequestra-
tion. Other approaches have been proposed, as well, such as cloud alkalinity, biomass
burial, enhanced ocean upwelling and downwelling, DAC by freezing, marine bioenergy
with carbon capture and sequestration, and electrochemical lining (The Royal Society,
2018b). The removal of other gases, such as methane, N2O, and CFCs with significant
global warming potential will also be important in the overall effort; however, the con-
centration of these gases in the atmosphere is much smaller than that of CO2, making them
very difficult to remove once they have been released.
The broad area of CCUS is a very rich one for chemical engineering; it presents
numerous opportunities for major contributions, as well as challenges to address over the
next years and decades, in the technical, economic, LCA, and integrated assessment man-
agement (IAM) areas. IAM is a quantitative tool for combining information from diverse

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

80 New Directions for Chemical Engineering

fields (i.e., science, economics, and policy) to assess the impact of emissions or their re-
duction. To address these challenges effectively, however, chemical engineers will need
to partner with engineers from other disciplines, such as environmental engineering (e.g.,
NASEM, 2019c).
DOE has identified priority research directions for CCUS (DOE, 2018c). Of these
priorities, chemical engineers could contribute in the following areas:

 Capture—designing high-performance solvents and developing environmen-


tally friendly solvent processes, designing sorbent materials and integrated
processes, developing membranes and related processes, and producing hy-
drogen from fossil fuels with CO2 capture.
 Utilization—designing interfaces for enhanced hydrocarbon recovery with
carbon storage; valorizing CO2 from catalytic, electrochemical, and photo-
chemical transformations to fuels, chemicals, and new materials; and tailor-
ing microbial and bioinspired approaches to CO2 conversion.
 Storage—advancing multiphysics and multiscale fluid flow to achieve Gt-
per-year capacity; locating, evaluating, and remediating existing and aban-
doned wells; and optimizing injection of CO2.
 Cross-cutting—integrating experiments, simulations, and machine learning
across multiple length scales to guide the development of materials and pro-
cesses; intensifying CCUS processes; incorporating social aspects into deci-
sion making; and integrating LCA and technoeconomic assessment (TEA),
along with environmental and social considerations, to guide technology
portfolio optimization.

Near-zero- and positive-emissions technologies emit almost zero GHGs or emit


GHGs to a lesser extent compared with alternative technologies, respectively. They in-
clude enhanced energy efficiency, clean or renewable electrification, bioenergy, hydrogen
and hydrogen-based fuels, and CCUS technologies. LCA and TEA (including scalability
of the specific technology) are two primary tools used to evaluate and rank these technol-
ogy options. Chemical engineers can play a role in the development of many, technologies
that have been reviewed extensively elsewhere (e.g., Bui et al., 2018; Fuss et al., 2018;
Hepburn et al., 2019; IEA, 2020d; NASEM, 2019b). Opportunities for chemical engineers
in the areas of DAC and CO2 and CH4 utilization are briefly summarized in this section.

Direct Air Capture

DAC appears to be a relatively easy fix for climate change and has the additional
advantage that it can be located close to the sequestration reservoir, thus avoiding the need
to use a pipeline for CO2 transportation. However, dilute systems, such as those with CO2
in the air at a concentration of about 400 ppm, require more energy than concentrated
systems, such as those with CO2 in flue gas from ammonia manufacture at a concentration
exceeding 98 percent; from coal-fired power plants at a concentration of 12–15 percent;
from cement, iron/steel, and glass production at a concentration of 20–35 percent; and

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Decarbonization of Energy Systems 81

from natural gas–fired power plants at a concentration of 3–4 percent on a volume basis
(NASEM, 2019a). Thus, the concentration of CO2 in flue gases is between almost 100
and 300 times that of CO2 in the air, and as a result, the energy required to capture CO2
from the air is 2 to 3 times greater than that required to capture CO2 from the flue gases
(Bui et al., 2018).
Areas on which chemical engineers will focus in the future include the develop-
ment of low-cost solid sorbents, highly CO2-selective materials that require reduced re-
generation energy, materials that are highly active in ambient conditions, and processes
with increased mass-transfer coefficient and high throughput and low pressure drop
(NASEM, 2019b). Other areas of focus will include packing designs, process intensifica-
tion, catalytic additives, and long-term stability of sorbent matrices.

FIGURE 3-17 Stocks and net flows of CO2, including potential uses and removal pathways. The
numbers 1 through 10 represent various pathways for CO2 use and removal: 1 = chemicals from
CO2; 2 = fuels from CO2; 3 = products from microalgae; 4 = concrete building materials; 5 = CO2
in enhanced oil recovery (EOR); 6 = bioenergy with carbon capture and storage; 7 = enhanced
weathering; 8 = forestry techniques; 9 = soil carbon sequestration techniques; and 10 = biochar.
SOURCE: Hepburn et al. (2019).

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

82 New Directions for Chemical Engineering

Carbon Utilization

Use of CO2 and CH4 as feedstocks is an important mechanism for CCUS. How-
ever, only a small amount of CO2 and CH4 emitted each year is currently being captured
and used. The main pathways for CO2 utilization include mineral carbonation, chemical
utilization, and biological utilization, while the main pathways for CH4 utilization are
chemical utilization, biological utilization, and direct uses as fuel. Most carbon utilization
technologies are early in their development phase. A number of research areas are de-
scribed in the above-referenced National Academies report (NASEM, 2019a); specific
opportunities for chemical engineers are highlighted in the following sections.

Mineral Carbonation

CO2 is used to make carbonates, such as cement and concrete, for use in the con-
struction sector, as well as in paper and food production. The conversion of CO2, which
is a low-energy molecule, into solid mineral carbonates in near-ambient temperatures is
one of the few thermodynamically favorable reactions of CO2 (NASEM, 2019a). For this
reason, as well as the sheer size of the construction materials market, the use of CO2 for
mineral carbonation is considered the largest and most favorable CO2 utilization pathway.
Challenges and opportunities for chemical engineers in this area include control of car-
bonation reactions, process design, accelerated carbonation and crystal growth, green syn-
thesis routes for alkaline reactants, structure property relationships, analytical and char-
acterization tools, and construction methodologies.

Chemical Conversion of CO2

Urea, polycarbonate, ethylene and propylene carbonates, salicylic acid, and pol-
yether carbonate are currently produced from CO2 at commercial scales; however, the
CO2 used is not derived from carbon capture processes. Commodity and fine chemicals
and fuels currently produced from CO2 at pilot plants are methanol, methane, CO, fuel
via a CO2-based Fischer-Tropsch process or direct pathway from CO2 to fuels, diphenyl
carbonate, and oxalic acid (NASEM., 2019a). Specific challenges and opportunities for
chemical engineers in the area of CO2 conversion to chemicals and fuels include the de-
velopment of long-lasting and stable catalysts that can also work when the CO2 feed
stream contains the impurities typically present in flue gases, low-temperature electro-
chemical conversion processes, enhanced conversion per pass and avoidance of carbonate
formation, and lower energy requirements for the anode in the electrochemical reduction
of CO2.

Chemical Conversion of CH4

Sources of methane waste gas include emissions from oil and gas plants, landfills,
sewage, manure, and other waste operations. The methane waste gas from oil and gas

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Decarbonization of Energy Systems 83

plants is primarily methane, whereas that from waste management operations, called bio-
gas, is a mixture of CO2 and methane. In contrast with CO2, methane is a high-value and
high-energy chemical and has no equivalent pathways to mineral carbonation. Because of
its high energy, it is used primarily as fuel, and thus any conversion to chemicals needs to
compete with the fuel value of methane. The cost trade-offs are likely to change with
increased use of CCUS technologies. Challenges and opportunities for chemical engineers
in this area include the development of catalysts, integration of catalyst and reactor tech-
nology, and identification of new chemical targets. Tools such as LCA and TEA will also
be critical in identifying situations in which methane conversion is cost-competitive with
the use of methane as fuel.

Biological Conversion of CO2

Biological conversion of CO2 holds great promise because some microorganisms


have a natural ability to capture and covert CO2. Photosynthetic pathways to CO2 utiliza-
tion include approaches using algae (products include biofuels, dietary protein and food
additives, and commodity and specialized chemicals), green algae (products include bio-
diesel, dietary protein, polyunsaturated fatty acids, pigments, lipids, and terpenoids), and
cyanobacteria (products include ethanol, butanol, fatty acids, heptadecane, limonene,
bisabolene, ethylene, isoprene, squalene, and farnesene). By using CO2 as feedstock, these
photosynthetic pathways mitigate the problem of the high-cost sugar feedstocks needed
for microbial pathways; however, the slow growth rates of algae and cyanobacteria pre-
vent them from achieving industrially relevant productivity and scale-up. Challenges and
opportunities for chemical engineers in this area include bioreactor and cultivation opti-
mization, analytical and monitoring tools, genome-scale modeling and improvement of
metabolic efficiency, bioprospecting, valorization of coproducts, genetic tools, and path-
ways to new products.

Biological Conversion of CH4

Methanotrophs can use methane as their carbon and energy source. However, sig-
nificant challenges arise, such as the risk of contamination during fermentation, buildup
of toxic intermediates, and the high cost of additives. Some commercial activity has taken
place in this space. Calysta™ has commercialized FeedKind® protein as an alternative
feed for fish, livestock, and pets, using no agricultural land and less water than is required
for similar agricultural products. Intrexon™ has used methanotrophs to produce high-
value chemicals, such as isobutanol and farnesene. And Mango Materials™ plans to con-
vert biogas into polyhydroxyalkanoate (PHA), which is a biodegradable plastic. Chal-
lenges and opportunities for chemical engineers in this area are the same as those for
biological conversion of CO2.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

84 New Directions for Chemical Engineering

CHALLENGES AND OPPORTUNITIES

In the energy sector, the overarching challenge for chemical engineers is to ad-
dress the environmental and climate impacts of current energy systems, particularly the
use of fossil-based energy sources. The transition to a low-carbon energy system will re-
quire a bridging strategy that relies on a hybrid system with a mix of energy carriers.
Chemical engineering is rooted in the transformation of stored energy carriers into more
convenient forms and into chemicals and materials. Chemical engineers have an important
opportunity to continue to apply their skillsets to non–fossil-fuel-based energy sources
and carriers and thus contribute to the decarbonization of the energy sector.
In the long term, achieving net-zero carbon emissions will require significant ad-
vances in photochemistry, electrochemistry, and engineering to enable efficient use of the
predominant source of energy for Earth—the solar flux. To this end, novel systems will
be required to improve the efficiency of photon capture and conversion to electrons; im-
prove the storage of electrons; and advance the direct and/or sequential conversion of
photons to energy carriers via reactions with H2O, N2, and CO2 to produce H2, NH3, and
liquid fuels, respectively.
Specifically, greater market penetration of PV solar panels will require a contin-
ued decrease in the cost of these panels, as well as increased electrification of energy
systems. Chemical engineers have an opportunity to play an enabling role in addressing
this challenge by advancing incremental improvements in device architecture and design,
lowering manufacturing costs, and improving reliability and durability. Beyond PV tech-
nologies, critical challenges hindering the greater use of PSC systems include operational
instabilities, short useful lives, and the containment and ultimate disposal of some toxic
components. Research conducted by chemical engineers and others will be critical to de-
veloping perovskite compositions that minimize degradation and ensure reliable long-
term operation. Conversion of photons to H2, NH3, or organic fuels will require advances
in the synthesis, characterization, and mechanistic assessment of catalytic solids, as well
as the development of materials that can withstand severe chemical, photochemical, and
electrochemical environments within complex hydrodynamics for systems that couple the
required reactions through diffusional controls, all of which are research opportunities for
chemical engineers.
Successfully mitigating climate change will require a long-term transition to re-
newable and sustainable sources of energy. In the short term, however, chemical engi-
neers have many opportunities to reduce the carbon footprint of fossil fuels. For coal,
these opportunities include research and technological advances to increase thermal effi-
ciency, further improve emission controls, and reduce water consumption. While natural
gas is a cleaner bridge fuel compared with other fossil fuels, innovations are still needed
throughout the value chain. Chemical engineers can enable advances that will minimize
or replace water use as a fracturing agent, improve storage and transportation, and better
integrate natural gas with renewable energy sources. For petroleum, challenges and re-
lated opportunities for chemical engineers include improved water management,

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Decarbonization of Energy Systems 85

increased recovery to extend well life, data-driven approaches to reservoir management,


and improved methane management. Decreasing the GHG emissions associated with all
types of fossil fuels will require the demonstration of cost-effective and secure carbon
capture and storage methods, a critical opportunity for chemical engineers.
Because most biogenic feedstocks have lower energy content than their fossil-
based counterparts, the greatest challenge for increased use of biofuels is the production
of high-density fuels at a reasonable cost that is competitive with that of existing, fossil-
based fuels. This challenge, combined with the need to account for the environmental
consequences of harvesting crops for energy use, creates opportunities for chemical engi-
neers to use systems-level economic, environmental, and technical analyses to select the
most viable biofuel options. Furthermore, increasing the market penetration of EVs for
personal transportation will require reimagining petroleum refineries that were designed
to produce gasoline or diesel fuel as their main products. Refineries will require reconfig-
uration to shift their product slate toward petrochemicals and low-carbon liquid fuels
needed for the difficult-to-decarbonize commercial transportation sector (e.g., heavy-duty
long-haul ground, aviation, and marine transportation). Full integration of existing petro-
leum refinery assets with biorefineries and the greater use of renewable energy will enable
significant reductions in carbon footprints and lower-cost low-carbon liquid fuels, another
area that presents considerable opportunities for chemical engineers.
For intermittent energy sources, chemical engineers can contribute to the devel-
opment of advanced materials that can increase the viability of wind and marine energy.
Chemical engineering research will also be critical to advancing low-carbon fuels; im-
proving petroleum refining; advancing clean hydrogen production; and developing im-
proved synthetic fuels for sectors, such as aviation, that are difficult to decarbonize. A
successful transition to low-carbon energy systems presents the challenge of energy stor-
age. Chemical engineers can enable the development of new battery materials, as well as
contribute to LCAs of competing battery technologies and the design of batteries for safe
end-of-life disposal.
For end uses, the production of cement, steel, and chemicals presents the clearest
opportunities for chemical engineers to contribute to decarbonization of the industrial sec-
tor. To achieve the net-zero emissions goal for industry, five broadly applicable decar-
bonization pillars require vigorous pursuit in parallel over the next decades: demand re-
duction, energy efficiency, fuel switching and electrification, transformative technologies
in sectors, and abatement. All will benefit from the contributions of chemical engineers.
Finally, CCUS will be centrally important to controlling the concentration of car-
bon in the atmosphere. This broad area is rich with opportunities for chemical engineers,
including LCA, integrated assessment management, and the science and technology ad-
vances necessary to advance direct air capture and carbon utilization.

Recommendation 3-1: Across the energy value chain, federal research funding
should be directed to advancing technologies that shift the energy mix to lower-car-
bon-intensity sources; developing novel low- or zero-carbon energy technologies; ad-
vancing the field of photochemistry; minimizing water use associated with energy

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

86 New Directions for Chemical Engineering

systems; and developing cost-effective and secure carbon capture, use, and storage
methods.

Recommendation 3-2: Researchers in academic and government laboratories and


industry practitioners should form interdisciplinary, cross-sector collaborations fo-
cused on pilot- and demonstration-scale projects and modeling and analysis for low-
carbon energy technologies.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

4
Sustainable Engineering Solutions for Environmental Systems

 While water, food, and air have historically been the focus of other disciplines,
chemical engineers bring both molecular- and systems-level thinking to pioneering
efforts in this highly interconnected space.
 The positive impact of chemical engineers will be magnified as they adapt to think-
ing beyond the traditional unit operation scale to focus at a global scale on the
water–energy–food nexus.
 The Earth’s atmosphere, with its large range of spatial and temporal scales, pre-
sents intriguing challenges for chemical engineers, who will continue to aid in im-
proving air quality.

By 2050, the Earth’s population is projected to grow to more than 9 billion, lead-
ing to a 60 percent increase in food demand, an 80 percent increase in energy demand,
and a 55 percent increase in water demand (OECD, 2012; UN, 2017). Food, energy, and
water are highly interconnected, with production or consumption of one usually directly
linked to production or consumption of another. Agricultural crops produce biofuels and
provide food for animals and humans. Energy is used to purify, transport, heat, or cool
water; to produce fertilizers; and to power farm equipment, food processing, and cooking.
Land and water diverted for energy production are no longer available for food production
and vice versa. Water is used for the production of fuels and electricity, and for agricul-
ture, food processing, livestock, and cooking. In addition to these complex relationships,
air quality affects or is affected by all three sectors and has a direct impact on human
health and well-being. Although water is a renewable resource that is conserved in the
Earth, freshwater can be depleted locally, and the policies for its local allocation are set
in a highly political context.
The concept of a water–energy–food (WEF) nexus was first introduced in 2011
by the World Economic Forum in Water Security: The Water-Energy-Food-Climate
Nexus. To better contextualize this WEF nexus, it is important to intertwine an additional
nexus, one that considers sustainability and environmental conditions (including climate),
as well as economic and social (including human health) components, sometimes referred
to as the “triple bottom line” (Das and Cabezas, 2018; Figure 4-1).
This chapter explores the role of chemical engineers in ensuring adequate food
supplies and clean water and air. After a discussion of the WEF nexus, scientific gaps in
understanding of the fundamental properties of water, as well as the need for engineering
solutions for water quality and supply, are reviewed. Opportunities for chemical engineers
to both pioneer and contribute to multidisciplinary efforts to advance agricultural and food
processing technologies are then described, followed by a discussion of the research needs
for understanding and improving air quality.

87

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

88 New Directions for Chemical Engineering

FIGURE 4-1 The concept of the water–energy–food nexus is further contextualized to understand
solutions to global problems and support human livelihoods and prosperity by showing its inter-
connectedness with the environment–economy–social nexus.

THE WATER–ENERGY–FOOD NEXUS

Solutions in the WEF nexus need to be sustainable; thus, environmental, eco-


nomic, and social factors need to be carefully considered. Increased resource demand pre-
sents a monumental challenge for chemical engineers. They have historically played a
central role in the energy sector, and their contributions in the interconnected space of
water, food, and air quality have been important and are now growing. Fundamental in-
sights from chemical engineering disciplines form the foundational knowledge necessary
to understand and create solutions in the WEF nexus, which inherently spans multiple
disciplines. Examples include the structure and dynamics of water, the nature and physics
of aerosol particles, and the scaling of synthetic protein production. Additionally, chemi-
cal engineers’ knowledge of biochemical engineering and its applications to agriculture,
and of separations with applications to water and air pollution, as well as their systems-
level understanding, is critical to solving global problems.
Fossil fuels (petroleum, natural gas, and coal) make up most of the landscape of
primary energy sources, contributing globally about 82 percent of the total energy pro-
duced in 2019 (IEA, 2021e). The continuing reliance on fossil fuels contributes to green-
house gas (GHG) emissions and air pollution, as well as associated human health, envi-
ronmental, and climate problems (D’Odorico et al., 2018). Today, one in five people lack

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Sustainable Engineering Solutions for Environmental Systems 89

access to modern electricity in their homes, and 3 billion people use wood, coal, charcoal,
or animal waste for heating and cooking, all of which have deleterious impacts on air
quality (D’Odorico et al., 2018). Note that energy generation and consumption are cov-
ered extensively in Chapter 3 and are discussed in the remainder of this chapter only in
the context of interdependencies with food, water, and air resources.
Over the past 50 years, global crop production has increased by more than 300
percent, and animal production by more than 250 percent. Dairy and meat production is
expected to increase by more than 60 percent by 2050 (D’Odorico et al., 2018). All of
these increases are attributable to the wider use of fertilizers and higher-yielding crop
varieties, which enable greater food production. Of the global land surface, 38 percent is
used for agriculture, with roughly one-third of that total used for crops and the other two-
thirds used for animal grazing (FAO, 2020).
Food waste is a major problem around the world. In the United States, the per-
centage of food wasted ranges from 16 percent for meat to ~25–30 percent for grains,
dairy, and eggs (Finley and Seiber, 2014). On a global scale, about one-quarter of the food
produced for human consumption is lost or wasted in the supply chain; therefore, the per-
centage of arable land and freshwater resources used to produce that food is also wasted
(D’Odorico et al., 2018).
The interdependency within the WEF nexus is massive (Figure 4-2). Approxi-
mately 15 percent of global water withdrawals are used for energy, 70 percent for agri-
culture, and the remaining 15 percent for other applications. About 8 percent of global
energy is used to transport, purify, and pump water, and about 30 percent to produce food.
About 1 percent of all food is used to produce energy (Garcia and You, 2016). The inter-
connection between water and energy in various sectors of the U.S. economy in 2011 is
depicted, from withdrawal and extraction through use, in the Sankey diagram in Figure
4-3 (DOE, 2014). The competition of water use for energy and food is at the core of the
WEF nexus landscape: the growing population and the larger number of people now in
the middle class drive demands for food, energy, and water, while at the same time fresh-
water resources are fixed and limited. This situation generates challenges for the environ-
mental, economic, and social aspects of life on Earth (Albrecht et al., 2018; DOE, 2014;
Simpson and Jewitt, 2019).
In 2016, the United Nations estimated that 800 million people suffer from food
insecurity, approximately 1.2 billion people lack access to electricity, and 800 million
people lack access to safe drinking water (Scanlon et al., 2017). The world’s population
is projected to increase to well over 9 billion in 2050, and the global middle class will
continue to expand. The resulting projected drain on WEF resources is so great that it will
be necessary to address the nexus holistically rather than as separate sectors. Future re-
source security (defined as the uninterrupted availability of resources at an affordable
price) will require not only the integrated management of water, energy, and food re-
sources but also a transition to a circular economy, with special attention to challenges of
sustainability and climate (Biggs et al., 2015; D’Odorico et al., 2018; see Chapter 6 for
further discussion of the circular economy). This integrated management approach is at
the core of the capabilities and focus of chemical engineers, from the chemical to the
system scale, with respect to reducing demand (conservation), increasing supplies, and

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

90 New Directions for Chemical Engineering

managing storage and transport (i.e., managing the spatial and temporal imbalances of
production and consumption) while working toward a circular economy in the WEF
nexus. The U.S. Department of Energy (DOE, 2014) frames integrated solutions for the
WEF nexus around six pillars:

 optimizing the freshwater efficiency of energy production, electricity gener-


ation, and end-use systems;
 optimizing the energy efficiency of water management, treatment, distribu-
tion, and end-use systems;
 enhancing the reliability and resilience of energy and water systems;
 increasing safe and productive use of nontraditional water sources;
 promoting responsible energy operations with respect to water quality, eco-
system, and seismic impacts; and
 exploiting productive synergies among water and energy systems.

FIGURE 4-2 The water–energy–food nexus emphasizes the interconnectivity of these three major
resources. The top illustration quantifies the reliance of each on the other two, while the bottom
illustration describes their connectedness and how they contribute to one another. SOURCE: Gar-
cia and You (2016).

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Sustainable Engineering Solutions for Environmental Systems 91

FIGURE 4-3 This Sankey diagram shows energy (top) and water (bottom) consumption associated
with various end uses. Numerical values shown are based on 2011 data. SOURCE: DOE (2014).

Chemical engineers can lead and contribute to advances in the many technology
vectors required within each of these pillars. Some key examples include the following:

 Reducing water demand (conservation)


– reduction of food waste and development of technologies that reduce
spoilage, and use of food waste to produce chemicals (e.g., bio-oils) and
to feed livestock (Balicka, 2020);
– implementation of a “more-crop-per-drop” (i.e., water productivity in ag-
riculture) approach, use of engineered crops with higher water efficiency
and/or drought tolerance, and development of better pesticides (Scanlon
et al., 2017);
– use of brackish groundwater instead of freshwater for energy (D’Odorico
et al., 2018; Scanlon et al., 2017);
– development of advanced sensors to avoid waste and improve process
reliability, as well as use of relevant data collection, analysis, and report-
ing (DOE, 2014);
– development of alternative fluids to replace freshwater in various pro-
cesses (D’Odorico et al., 2018; Scanlon et al., 2017);

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

92 New Directions for Chemical Engineering

– improvement of the water efficiency of industrial processes, and replace-


ment of freshwater (e.g., with supercritical CO2, nitrogen, nanomaterials,
or liquid hydrocarbons) in oil and gas extraction (D’Odorico et al., 2018);
– development of efficient, less expensive, and lower-water-use or water-
less cooling options for thermoelectric power plants, and options for re-
covering waste heat and reducing the need for cooling in thermoelectric
power plants (D’Odorico et al., 2018; Scanlon et al., 2017);
– improvement of water reuse within homes (e.g., new uses for greywater)
and development of technologies for waterless products, processes, and
activities (D’Odorico et al., 2018; Scanlon et al., 2017); and
– increased use of renewable energy (wind and solar) to reduce the water
demand for electricity generation (DOE, 2014; IChemE, 2015).
 Increasing supplies
- Food
- closing the yield gap of crop productivity between low- and middle
income countries and higher-income countries while minimizing en-
vironmental, social, and other impacts (“sustainable intensifica-
tion”);
- producing genetically modified (GM) crops that are insect- and herb-
icide-resistant and tolerant to drought;
- producing GM livestock to change the fat content in milk;
- producing in vitro meat that does not involve raising livestock (cell-
cultured meat) or plant protein–based meat; and
- developing hydroponics- and aquaponics-based technologies
(D’Odorico et al., 2018; Scanlon et al., 2017).
- Water
- further developing desalination options for brackish groundwater or
seawater so that desalination expands beyond specific uses that deal
with small volumes of water (e.g., drinking water) and population
groups that can accommodate the higher costs;
- capturing stormwater;
- developing advanced materials for removing chemical and biological
contaminants from water (e.g., removing lead contamination from
drinking water to address contamination scenarios such as those
faced by residents of Flint, Michigan); and
- treating municipal wastewater (D’Odorico et al., 2018).
- Energy
- developing second- (i.e., from agricultural waste) and third- (i.e.,
from nonedible biomass) generation biofuels,1
- developing advanced energy crops,
- using waste for energy production,

1
Today’s ethanol from fermentation of sugars and biodiesel from plants are considered first-
generation biofuels, and in both cases, the feedstocks can be used for food as well.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Sustainable Engineering Solutions for Environmental Systems 93

-improving the energy efficiency of production of chemicals and


products, and
- developing technology for cost-effective recovery of dissipated en-
ergy from electricity generation that could also help with carbon cap-
ture and storage (DOE, 2014).
 Managing storage and transport
- developing technologies that enhance food preservation and storage
(D’Odorico et al., 2018), and
- improving battery and other energy storage technologies (IChemE,
2015).

Interdisciplinary research that integrates the physical, agroecological, and social


sciences, as well as economics, is needed, along with the involvement of academia, in-
dustry, and government. While chemical engineering’s role in the energy sector is dis-
cussed in Chapter 3, the ties among food, water, air, the environment, and energy are
apparent and are key to the future of both the discipline and society writ large.

MOLECULAR SCIENCE AND ENGINEERING OF WATER SOLUTIONS

Chemical engineers have a leading role in molecular science and engineering that
requires a fundamental understanding of water structure, dynamics, and interactions, as
well as in the development of new complex separation processes. However, issues of wa-
ter scarcity, preservation, and purification are on a global scale, and solutions to these
complex issues therefore require an unprecedented ability to think across scales ranging
from the atomic to the geologic. The ability of chemical engineers to solve problems from
the molecular to the system-design level will be critical to meeting these challenges, but
they will also have to learn from and interact with civil and environmental engineers and
to understand system boundaries that go far beyond a unit operation, process, or plant
scale.

Water Purification

The purification of water involves the removal of a variety of chemicals or solid


materials using a range of processes (Figure 4-4). Several water sources and associated
opportunities for chemical engineering are highlighted below.

Desalination

The conversion of ocean or brackish water to drinking water is one of the great
engineering advances of recent decades, and this technology is applied at scale in numer-
ous locations (e.g., the Ashkelon Seawater Reverse Osmosis Plant in Israel2). The two
main technologies applied to this problem are distillation, in several forms, and reverse

2
See https://www.water-technology.net/projects/israel.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

94 New Directions for Chemical Engineering

osmosis (RO) membrane systems, the latter of which account for 65 percent of global
capacity (Abdelkareem et al., 2018; Bhojwani et al., 2019). Modern RO plants operate
near the thermodynamic limit but are still very large consumers of energy. Opportunities
to improve the overall energy efficiency of a seawater RO plant involve energy and waste-
heat management as much, if not more, than the development of new membrane materials
or processes. Many RO plants are located in regions where renewable power sources
(wind or solar) can be contributors, or in remote regions where nuclear power is a viable
alternative. In the latter case, it is possible to use an arrangement whereby the power plant
runs at full capacity, with electricity fed into the grid, during periods of high demand, and
is otherwise used to purify water, which is much easier to store than electricity.

FIGURE 4-4 Comparison of several conventional and membrane processes for water purification
and the subsequent mechanisms of water transport through membranes based on solute size and
molecular weight. NOTE: ED = electrodialysis; MF = microfiltration; NF = nanofiltration; NOM
= natural organic matter; RO = reverse osmosis; UF = ultrafiltration. SOURCE: Landsman et al.
(2020).

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Sustainable Engineering Solutions for Environmental Systems 95

Produced Water

While sea- or brackish-water desalination is well developed, current and future


challenges relate to dealing with water contaminated in different ways. For example, oil
production yields oil along with so-called produced water, which in many cases is larger
in volume than the oil. This water contains suspended oil, additives, and solids, and in
many cases, heavy metals or radioactive elements. If hydraulic fracturing has been used,
at least some of the produced water will contain polymeric or surfactant-based “fracking”
fluids (this water, sometimes called flowback, was, ironically, initially formulated with
freshwater). The oil and water may be produced as an emulsion (either water-in-oil or oil-
in-water, or as a multiple emulsion), which if stabilized by asphaltenes or resins can be
quite stable. Economical treatment and reuse are very difficult, so this water is often
reinjected into the formation. The development of technology for cost-effective treatment
of this water, particularly on site, presents a substantial opportunity for chemical engi-
neers.
Similar challenges continue to exist in, for example, managing the water pro-
duced during coal and other mining; the iron-ore treatment in the taconite process; and
multicomponent radioactive liquid wastes such as those at the Hanford, Washington, nu-
clear reactor site (where 45 years of operation resulted in an estimated 440 billion gallons
of wastewater [Washington State Department of Ecology, 2021]). Other challenges in-
clude the removal of boron and other neutral materials, as well as the removal of environ-
mentally persistent perfluorocarbon molecules.
In the case of oil-field or mining water, the separations need to be carried out in
steps to accomplish the sequential removal of colloidal and larger particles, then oils, and
finally salts. Each of these steps could be considered a unit operation. The science of
coagulation and flocculation is well known and practiced in municipal water treatment
plants, but the implementation of new coagulants or flocculates is often too costly. None-
theless, there are substantial opportunities to develop new polymer or surfactant chemis-
tries that can address these challenges and to leverage the principles of self-assembly to
treat these waste streams.
After solids and oils have been removed, membrane processes can be applied.
The membranes used for water treatment are usually porous polymeric films with pore
structures and sizes designed for the application at hand. The flux of water and solutes
across the membrane always depends on a balance between permeability and selectivity,
as outlined by Landsman and colleagues (2020). In this process, the design parameters
are the pore size and distribution; solute size and shape; solute concentration; osmotic
pressure; and, importantly, interactions between the solute and the membrane, which can
be electrostatic, chemical, or biospecific (Landsman et al., 2020). Each of these parame-
ters can be exploited to design separations for water contaminated in various ways.
A range of research-based approaches have been suggested for new membranes.
Molecularly heterogeneous polymers (multiblock copolymers) can be designed to form
bicontinuous solids, and then treated to make solids of controlled porosity or synthesized
with carbon nanotube channels. Such membranes could be functionalized to be electroac-

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

96 New Directions for Chemical Engineering

tive or biomolecularly specific. They could incorporate elements of ion-specific ex-


change, or be catalytic and react with and separate target molecules simultaneously. This
is a rich field for chemical engineers with an interest in transport, catalytic and reaction
chemistry, thermodynamics, self-assembly, and materials science. To make an impact on
real-world applications, research in this area needs to focus on the factors that typically
limit existing technology solutions (e.g., biofouling, durability, cost). Moving forward, an
important consideration will be management of the life cycle of such membranes and
incorporation of recycle or upcycle characteristics into their design. For example, the pro-
duction of polymeric membranes often relies on polar aprotic solvents, such as N,N-di-
methylformamide (DMF), or 1,4-dioxane and tetrahydrofuran (THF). These solvents pose
considerable environmental challenges, and it would be preferable to replace them with
environmentally responsible alternatives.

Trace-Element Recovery

There is an increasing need to recover recyclable and reusable nutrients and minor
or trace elements—particularly phosphorous, rare earth elements, and energy-related ele-
ments—from waste streams. In many cases, natural deposits of their associated minerals
are limited within the United States (Jyothi et al., 2020). In addition, challenges remain
for the extraction and/or destruction of both emerging and legacy trace contaminants that
threaten human and ecological health. The low concentrations of trace elements make
removal challenging, especially in waters containing high salinity or highly complex or-
ganic matrices. A wide variety of conventional separations have been employed (e.g.,
chemical precipitation, coagulation-flocculation, flotation, solvent extraction, ion ex-
change, adsorption, membrane processes, filtration, reverse osmosis, and electrochemical
techniques), with varying levels of technoeconomic feasibility depending on the contam-
inant properties, background water composition, and treatment goals (Naidu et al., 2019;
Pereao et al., 2018). Two of the more promising approaches, especially for recovery of
trace contaminants, use sorption (including adsorption, absorption, ion exchange, and pre-
cipitation) and/or membrane processes. For many trace elements, however, significant
advances in these technologies are needed to expand recovery and reuse and reduce treat-
ment costs (Elbashier, et al., 2021; Li et al., 2021b). Several promising new techniques
employing novel sorbents have emerged, including electrospun nanofibers with a highly
specific surface for adsorbent applications (e.g., Sharma, 2013) and ionic imprinted pol-
ymers (e.g., Branger et al., 2013; Luo et al., 2015). In the case of membrane technologies,
novel materials targeted for recovery of rare earth elements include metal organic frame-
works and liquid membranes (Smith et al., 2019; Tursi et al., 2021).
The recovery of phosphorus from water presents formidable challenges. Reserves
of phosphate rock are rapidly being depleted, and the current cost of recovering phospho-
rus from wastewater or agricultural runoff exceeds the cost of separating conventional
phosphate from rock. Therefore, new and cost-effective technologies for recovering phos-
phorous from wastewater are needed. To this end, as Peng and colleagues (2018) indicate,
more than 50 such technologies, including biological, chemical, and physical processes,
have been developed. One of the most common end products for phosphorus recovery is

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Sustainable Engineering Solutions for Environmental Systems 97

precipitated struvite or vivianite. Fluidized bed reactors that achieve upwards of 80 per-
cent recovery from wastewater have been reported (Nelson et al., 2017), but the energy
consumption, cost, and footprint of available processes are all too high to encourage their
widespread adoption. Newer alternatives based on adsorption technologies, ranging from
alginates to peptide-based materials, as well as biobased systems, are rapidly being ex-
plored, with encouraging results (e.g., Jama-Rodzenska et al., 2021; Su et al., 2020; Zhang
et al., 2020a). Alternative approaches, such as Donnan dialysis with ion-exchange mem-
branes, are also being investigated for concentrated water containing high concentrations
of ions, organic matter, and suspended solids in concentrated waste streams (Shashvatt et
al., 2021). However, much research is still needed to bring such technologies into practice.
The extraction of lithium or uranium from water streams poses similar challenges,
including extraction from seawater and water produced during oil and gas operations. In
the case of lithium, concentrations range from about 5 mg/L to about 500 mg/L, and re-
moval can be accomplished using solar evaporation, adsorbents, membrane-based pro-
cesses, and electrolysis-based systems (Kumar et al., 2019). Current challenges include
accelerating concentration processes and dealing with lower lithium concentrations. In
the case of oil and gas wastewater, metal oxide adsorbents and membrane technologies
are promising. However, current membrane materials lack sufficient lithium-ion selectiv-
ity, and novel membrane approaches, such as 12-Crown-4-functionalized polymer mem-
branes (Warnock et al., 2021) and Cu-m-phenylenediamine (MPD) membranes (Wang et
al., 2021a), require further research.
Finally, concerns resulting from the widespread contamination of drinking water
by legacy contaminants, such as perfluorinated compounds and lead released from water
distribution systems, highlight the need for innovative treatment technologies that address
recalcitrant organics and metals with known human health risks at trace concentrations.
Per- and polyfluoroalkyl substances (PFAS) have no known half-life and have been found
in more than 2,000 locations within the Unites States. The toxicity of the more than 4,000
PFAS compounds varies widely; state-level water quality regulations for some PFAS
compounds, such as perfluorooctanoic acid (PFOA), are at the part-per-trillion level. Sig-
nificant research into treatment strategies includes technologies focused on contaminant
destruction (e.g., advanced oxidation and reduction, photolysis, electrochemical oxida-
tion, and incineration) and those aimed at separating the compounds from water (e.g.,
activated carbon adsorption, polymeric adsorbents, ion exchange, ozofractionation, and
membrane separation; Ross et al., 2018). To date, however, no single technology has been
identified that is cost-effective, not energy intensive, and universally effective at treating
the range of PFAS compounds (e.g., varying functionality and chain length) within the
complex matrices of source waters (Crone et al., 2019; Gagliano et al., 2020). The devel-
opment of treatment technologies for these persistent chemicals is a major opportunity for
chemical engineers.
Lead contamination of drinking water resulting from dissolution of lead solder,
fixtures, and piping is another concern requiring innovative solutions, including consid-
eration of point-of-use treatment technologies. Maintenance of lead scales within distri-
bution systems is the typical control mechanism for ensuring water quality; however,
changes in water quality can dramatically affect lead release and compromise drinking

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

98 New Directions for Chemical Engineering

water quality either across the distribution system or within premise plumbing (Wahman
et al., 2021). Thus, the development of effective point-of-use technologies (e.g., reverse
osmosis, adsorbents) that can remove both dissolved and particulate lead is needed
(Brown et al., 2017; Verhougstraete et al., 2019). Nanoenabled technologies are also use-
ful for lead and other metal contaminants, such as copper, Cu(II); chromium, Cr(VI); and
arsenic, As(V). For example, Greenstein and colleagues (2019) have examined polymer-
iron oxide nanofiber composites and iron oxide–coated polyacrylonitrile fibers for re-
moval of lead, Pb(II), and these other metal ions, as well as suspended solids. In these
cases and others, solutions to such separation problems depend on fundamental chemical
engineering principles, such as thermodynamics, transport phenomena, chemical kinetics,
engineering of nanoscale materials, and process design.

Removal of Microplastics

Chemical engineers have opportunities to address the challenges of remedying


the damage done by pollution in large bodies of water. A significant challenge is the re-
mediation of microplastics in oceans and large lakes. An estimated 8 million metric tons
of plastic enters waterways annually (NOAA, 2021b), and a significant portion of that
total is in the form of so-called microplastics, which are less than 5 mm in size. Micro-
plastics enter the water directly as waste from consumer products, such as cosmetics, but
are also produced by attrition of larger pieces of plastic or incomplete incineration of
plastics. The Great Pacific Garbage Patch (GPGP) covers approximately 1.6 million
square kilometers in the North Pacific (Lebreton et al., 2018). By comparison, the Deep-
water Horizon oil spill covered about 150,000 square kilometers in the Gulf of Mexico,
or 10 percent of the area of the GPGP (NOAA, 2021a). Microplastics pose a risk to larval
fish and thus can potentially impact a significant source of food protein (Gove et al.,
2019).
Remediation of such a large area of open sea is challenging. Some microbial treat-
ments have been proposed (Auta et al., 2017), but appear to be applied more effectively
in treatment plants. Open-sea cleanup of microplastics will likely have to rely on their
physical removal. Such harvesting will require the concentration of microplastics into
flocs, which may be promoted by highly effective flocculating agents (Roh et al., 2019).
Such flocs could then also be subjected to bioremediation processes.
Removal of microplastics from water and other media presents opportunities for
chemical engineers. Removal methods include physical sorption and filtration, biological
removal and ingestion, and chemical treatments (Iyare et al., 2020; Padervand et al.,
2020). Physical sorption methods include adsorption on green algae (depending on the
microparticles’ surface charge), removal using membrane technology, and removal using
filtration technology (e.g., ultrafiltration, disc filter, rapid sand filtration, and dissolved
air flotation [Talvitie et al., 2017]). Chemical removal methods include coagulation, ag-
glomeration, and settling processes (Lapointe et al., 2020) using Fe3+- and Al3+-based salts
and other coagulants (e.g., polyacrylamide [PAM]); electrocoagulation; sol-gel reactions;
and photocatalytic degradation. Biological removal and ingestion methods include bio-
logical degradation using microorganisms (Padervand et al., 2020).

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Sustainable Engineering Solutions for Environmental Systems 99

Fundamental Properties of Water

The technological advances needed to secure a global freshwater supply, as de-


scribed above, rely on fundamental breakthroughs in understanding the structure of water
and its dynamics (Debenedetti and Klein, 2017). For example, the structuring of water
near an interface plays a critical role in the fouling of that surface, whether it is on the
outside of a naval vessel or the inside of a pipe in a water treatment plant. Similarly,
whether ions, for example, are depleted or concentrated at the air–water interface is of
considerable importance in the development of new water-purification technologies or of
water transport models for agriculture or climate prediction. Chemical engineers, primar-
ily in academia, are heavily involved in experimental and theoretical studies of water both
with and without other additives. This section reviews and summarizes some of the op-
portunities and challenges therein.
Considerable advances have been made over the past two decades regarding the
structure and dynamics of water under a wide range of conditions and environments.
Much of that progress has been fueled by theoretical and computational advances, coupled
with the development of experimental techniques capable of probing structure and dy-
namics over a wide range of length and time scales (Figure 4-5).

Structure of Pure Water

Some of the more intriguing developments in understanding the structure of pure


water have been enabled by developments in synchrotron light sources and vibrational
spectroscopy, coupled with progress in models and computation that has permitted a
deeper interpretation of such measurements (Bjorneholm et al., 2016). Various types of
vibrational spectroscopy, as well as elastic and inelastic neutron scattering, have gradually
been refined to provide detailed insights into the structure and dynamic processes in bulk
water and at interfaces, particularly when coupled with selective deuteration, at time
scales ranging from fractions of a picosecond to tens of nanoseconds. These time scales
encompass a wide range of characteristic molecular processes, from O–H bond vibrations
to the motion of water and water-bound molecules, such as lipids in bilayer membranes
or polymers in solution. Two-dimensional infrared spectroscopy, for example, continues
to push its limits of sensitivity and applicability and has provided previously inaccessible
information about how molecules, such as amyloid proteins, self-assemble in aqueous
solution (Middleton et al., 2012; Shim et al., 2009). Sum frequency generation (SFG) is
providing much-needed insights into chemical reactions at interfaces, including those oc-
curring at electrodes for energy generation (Neri et al., 2017). In situ experiments, such
as inelastic neutron scattering measurements, can now provide critical information about
industrially relevant reactions as they occur in real-world processes, such as the formation
of NaOH and Na2SiO3 activated in low-CO2 cements (Gong et al., 2019). Experiments at
light sources (synchrotrons) have for decades been important for structural characteriza-
tion, such as crystallographic studies of water, hydrates, and aqueous solutions of biomol-
ecules, including proteins. Synchrotron light sources continue to be upgraded, and with

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

100 New Directions for Chemical Engineering

FIGURE 4-5 Illustration showing the structure and dynamics of water when probed by various
spectroscopies at multiple length and time scales. NOTE: AFM = atomic force microscope;
APXPS = ambient-pressure X-ray photoelectron spectroscopy; GISAS = grazing-incidence small-
angle scattering; HB = hydrogen bond; IR = infrared; SANS/SAXS = small-angle neutron scatter-
ing/small-angle X-ray scattering; SFA = smooth factor analysis; SFG = sum frequency generation;
STM = scanning tunnel microscopy; ODNP = Overhauser dynamic nuclear polarization; O NMRD
= oxygen nuclear magnetic resonance dispersion; QENS = quasielastic neutron scattering; XAS-
FY = X-ray absorption spectroscopy—fluorescence yield. SOURCE: Monroe et al. (2020).

those upgrades, new methods are providing an unprecedented view into the structure of
water in a wide range of scenarios, including as ultrathin films or in samples undergoing
unusual phase transitions (Byrne et al., 2021).
Over the past 30 years, the issue of water polymorphism in pure water and aque-
ous solutions has gradually been fleshed out (Bachler et al., 2019; Debenedetti, 2003;
Gallo et al., 2016; Handle et al., 2017). The phase diagram of water has been expanded to
recognize the existence of several forms of amorphous ice, low-density amorphous (LDA)
ice, and high-density amorphous (HDA) ice. LDA ice can be formed by rapid vapor dep-
osition of water onto ultracold substrates (below 120 K); it has a density of 0.94 g/cm3,
but its viscosity is higher than that of water. In contrast, HDA ice has a density of 1.17
g/cm3 and can be formed by pressurizing a particular phase of ice (ice Ih) at temperatures
in the vicinity of 140 K. A third form of amorphous ice, referred to as very-high-density
amorphous ice, has a density of 1.26 g/cm3 and can be formed by compressing HDA to
pressures above 1 GPa. The existence and origin of these newly found forms of ice have

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Sustainable Engineering Solutions for Environmental Systems 101

been understood largely based on sophisticated molecular simulations, many of which


have been carried out by chemical engineers (Palmer et al., 2018). The discovery of these
ices has helped explain several astrophysical observations, as they arise, for example, in
comets or icy moons and in some of the coldest regions of the Earth’s atmosphere. Amor-
phous ice is also relevant for engineering applications and could be used for the chemical
and structural stabilization of biological molecules at low temperatures (as needed, for
example, in cryo-transmission electron microscopy) over extended periods. The discovery
of LDA ice formed by vapor deposition has inspired the creation by chemists and chemi-
cal engineers of other classes of emerging engineering materials, such as ultrastable va-
por-deposited glasses, which offer unusual mechanical characteristics and important ad-
vantages for applications in electronics, such as organic light-emitting diodes and
polymorphic metallic films that respond to light.

Structure of Water at Interfaces

Considerable progress has also been made in understanding the structure of water
in inhomogeneous environments—for example, water at interfaces and under extreme
confinement. Consensus has gradually emerged, primarily from various types of infrared
measurements and molecular models, that an individual —OH group from the molecule
is freed from the hydrogen-bonding network and points toward the vapor phase. Water
molecules near the interfacial “layers” are believed to be slightly more disorganized than
bulk, fully hydrogen-bonded water, and considerable debate continues around the
acid/base characteristics of interfacial water. This last issue is of particular significance to
electrochemistry and, more generally, to chemical reactions at interfaces, and presents
exciting opportunities for chemical engineering. It is also important for the interpretation
and design of adsorption processes at aqueous interfaces. Clathrate formation, for exam-
ple, with both methane and CO2, is believed to be nucleated at the water interface (Li et
al., 2020a; Liang et al., 2019). A better understanding of such interfaces would therefore
enable the discovery and development of clathrate formation inhibitors or stabilizers.
Beyond pure water, the structure of aqueous solutions at interfaces presents im-
portant challenges (Bjorneholm et al., 2016). The issue of whether ions, for example, are
depleted or concentrated at the air–water interface is still being debated. Current thinking
is that large, more polarizable halide anions are depleted from the air–water interface to a
lesser extent relative to smaller and less polarizable ions (Ghosal, 2005; Jungwirth and
Tobias, 2006; Tong et al., 2018). Iodide, however, is believed to be enriched at that inter-
face. These trends are strongly influenced by the presence of molecular solutes, and de-
pending on the polarity and size of such solutes, it is difficult to predict the structure of
an aqueous solution at an interface, posing considerable challenges for the design of ad-
sorption operations, for example.
Theory and simulation have the potential to provide the tools needed to describe
such systems, but molecular models of water and ions are currently unable to describe the
solubility of ions in water or to predict the segregation or enrichment effects observed in
experiments (Mester and Panagiotopoulos, 2015). Such models have been developed for
bulk pure water, and the effects of ions have been included as an afterthought, thereby

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

102 New Directions for Chemical Engineering

restricting the models’ ability to describe solutions or mixtures and interfaces. Attempts
to circumvent this problem by relying on quantum mechanical methods have been hin-
dered by the computational demands of such calculations, which continue to exceed avail-
able resources. Recent approaches involving various combinations of advanced sampling
concepts from statistical mechanics, quantum mechanical calculations of intermolecular
interactions, and emerging concepts from machine learning offer considerable promise
for reducing the description of water and aqueous solutions and their interfaces to a trac-
table problem. As chemical engineers strive to conceive and design chemical processes
involving aqueous interfaces, it will be important for such predictive models to be devel-
oped and brought to bear on the design and optimization of modern water-based technol-
ogies.
The challenges and gaps in understanding that arise at the air–water interface are
exacerbated at solid interfaces with metals; oxides; or organic matter, including biomole-
cules. Water–metal interfaces are central to catalysis and electrochemistry, and probes
capable of directly reporting the structure of interfacial water are severely limited. Im-
portantly, existing measurements require interpretation based on molecular models, and
such models continue to suffer from the same shortcomings highlighted in the context of
air–water interfaces. Much work remains to be done not only in understanding the struc-
ture of water at such interfaces but also in manipulating that structure to control reactivity
and transport (Ruiz-Lopez et al., 2020).
Considerable effort has been focused on understanding both hydrophobicity and
hydrophobic surfaces. Experimental evidence from SFG experiments has now shown that,
as with the air–water interface, water molecules have an individual “dangling” —OH
group at a hydrophobic surface (Bjorneholm et al., 2016). Results of experiments on in-
dividual surfaces also indicate that, beyond the first monolayer of water molecules at such
an interface, water rapidly adopts a bulk, isotropic structure. Those findings are in conflict
with measurements of the forces between two hydrophobic surfaces, which indicate that
an attractive force can already be felt at separations on the order of 20 nm (Kekicheff et
al., 2018), suggesting that some level of structural influence is felt at relatively large dis-
tances. Simulations of water structure and forces between hydrophobic surfaces have
added some clarity to this ongoing debate, with some calculations finding evidence of
long-range attractions and others finding only short-range interactions, depending on the
type of surface and water model adopted in the calculations (Altabet and Debenedetti,
2017; Eun and Berkowitz, 2011; Hua et al., 2009).
Even less is known about how these interactions might change in the presence of
ionic species, including ions or charged molecules. Recently, however, chemists and
chemical engineers have made important advances by relying on synthesis of carefully
conceived organic molecules and surfaces that present hydrophobic groups and charges
in precise arrangements, coupled with atomic- or surface-force apparatus measurements
(Ma et al., 2015). Those experiments have revealed that chemical heterogeneity at the
nanometer level plays a significant role in the structure of water and the resulting hydro-
phobic interactions. A key advance is the realization that hydrophobicity is not an inherent
characteristic of nonpolar domains but is instead controlled by functional groups sepa-
rated by nanometer scales. Cleverly designed simulations have helped explain and exploit

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Sustainable Engineering Solutions for Environmental Systems 103

some of these observations (Kim et al., 2015). From an engineering perspective, these
findings suggest that judicious positioning of charges next to hydrophobic domains rep-
resents a strategy for tuning hydrophobic forces, with a wide range of implications, from
the design of water-repellent coatings, antifreeze proteins, and self-assembling systems to
the stabilization of colloidal suspensions.

Dynamics of Water

For chemical engineering, the relevance and excitement of many of these new
discoveries surrounding the structure of water are matched only by new observations per-
taining to the dynamics of water under confinement or near interfaces. Early experiments
with membranes consisting of carbon nanotube pores (less than 2 nm in diameter) re-
vealed that flow rates through such pores are roughly three orders of magnitude faster
than anticipated by continuum flow calculations (Holt et al., 2006). Fast transport of gases
in nanopores had been anticipated on the basis of simulations by chemical engineers
(Skoulidas et al., 2002), and the experimental observations were explained by invoking
the formation of a frictionless interface at the nanotube wall. It has also been reported that
such a phenomenon occurs with both organic solvents (e.g., alkanes) and hydrogen-bond-
ing solvents, and that in the case of water, the flow rate gradually drops over time, pre-
sumably as a result of some level of ordering in the pore. Subsequent experiments and
simulations have shown that such ordering can in fact lead to even faster transport of
water—as fast as 1 meter per second, and at small dimensions, in the range of a nanometer
(Marbach et al., 2018; Striolo, 2006; Tunuguntla et al., 2017).
Chemical engineers have been particularly adept at drawing inspiration from bi-
ology to develop solutions for problems ranging from reengineering proteins to co-opting
bacteria for water reuse. Research on technologies for water transport through protein
pores, such as ion pumps or aquaporins, has been used in the design of the high-flux sys-
tems mentioned above or selective pores for ion transport. Fundamental theoretical and
computational research has paved the way for the design of intriguing separation or en-
ergy-generation processes, including recent discoveries about the transport of aqueous
ionic solutions through Janus nanopores—designer pores consisting of adjacent sections
with different diameters and surface charges—that could potentially become important
sources of energy from simple salinity gradients (Yang et al., 2018; Zhang et al., 2017;
Zhu et al., 2018). More work is needed to scale up and optimize observations that have
thus far been limited to laboratory-scale observations, but these and other “tailored-pore”
approaches offer considerable promise.
Similarly, additional research on aqueous electrocatalysis, photocatalysis, and
photoelectrocatalysis (Li et al., 2020b; McMichael et al., 2021; Ochedi et al., 2020)—for
which the basic idea is to use electrons and/or photons to transform readily available sub-
stances, such as water and CO2, into sustainable chemical fuels such as hydrogen and
methanol—is likely in the near future to see application in commercial projects that could
shape the industrial landscape for generations. In the face of rapidly changing climate
patterns, fundamental research on topics related to ice nucleation, clathrate formation or
destabilization, and the fate of aerosols can inform technologies aimed at mitigating the

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

104 New Directions for Chemical Engineering

impacts of climate change. As water sources change and conservation is accelerated, mon-
itoring will be of critical importance at a global scale. For example, “dating” efforts to
gauge the age of aquifers using atom trap trace analysis of krypton and noble gases are
only a few years old and are starting to yield intriguing data, but are as yet not widely
known or utilized (Gerber et al., 2017).

FEEDING A GROWING POPULATION

The advent of farming-based agriculture was an essential enabler of human civi-


lization. Innovative strides in agriculture in the 20th century, many of which were enabled
by chemical engineers or researchers working in what would ultimately become the chem-
ical engineering discipline, allowed substantial improvements in food yields per land area,
the ability to efficiently prolong the life of foods, and the ability to deliver food to con-
sumers over long distances. These innovations are exemplified by the development and
scale-up of the Haber–Bosch process for ammonia production, which rivals global bio-
logical nitrogen fixation in magnitude (Galloway et al., 2008). The development of refrig-
eration, today a standard lesson in chemical engineering’s thermodynamics courses, is
another critical example of the technological march toward improved efficiency in food
production.
Today, agricultural pursuits use an astounding 38 percent of all land on the planet
outside of Antarctica (FAO, 2020), and the raising of livestock is responsible for 14.5
percent of global GHG emissions (FAO, 2021). Researchers estimate that besides decar-
bonizing the energy sector (as discussed in Chapter 3), better management of food pro-
duction and agriculture represents the other primary way to reduce global GHG emissions
(Mbow et al., 2019; see Figures 4-6 and 4-7). As a result of continued population growth
and anthropogenic climate change, humankind will need to rethink and reinvent agricul-
tural and food production practices toward more sustainable land and resource use (Til-
man et al., 2011).
These global-scale challenges present significant opportunities for chemical en-
gineers, especially in the application of systems-level approaches to evaluate and optimize
products and processes. Examples of specific challenges are mentioned below, but are by
no means exhaustive. Generally, opportunities for innovation can be viewed in terms of
increasing efficiency in agricultural practices and developing “leapfrog” technologies that
are more sustainable than current agricultural methods.

Improving Agricultural Efficiency

The last three decades have seen significant technological innovations in agricul-
ture, resulting in more precise application of resources, such as fertilizer; the advent of
modern biomolecular techniques for engineering improvements in plant- and animal-
based products; and the use of automated machinery and, more recently, robotics to re-
place manual labor. These and other agricultural advances will inevitably continue.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Sustainable Engineering Solutions for Environmental Systems 105

FIGURE 4-6 Chart illustrating the magnitude of greenhouse gases emitted at distinct points across
the supply chain for several food products. NOTE: Greenhouse gas emissions are given as global
average values based on data across 38,700 commercially viable farms in 119 countries. SOURCE:
Our World in Data (2021b).

Since the emergence of metabolic engineering and synthetic biology, tools of


modern molecular biology have become commonplace in chemical engineering. Tremen-
dous opportunities exist to combine these tools with systems-level approaches to improve
per-land-area crop yields; enhance the efficiency of water and nutrient use; and increase
resistance to fungal infections, insect infestations, drought, and other adverse cli-
mate/weather and biological effects. While the public has generally viewed genetic mod-
ifications of plants with mistrust, techniques based on CRISPR (clustered regularly inter-
spaced short palindromic repeats) technology may ultimately enable precise genomic
editing without being considered genetic modification, representing a major boon for ap-
plication in food crops. The use of genome-wide association studies, especially in the
broad gene pools that exist in undomesticated crops, and translation to domesticated food
crops present an opportunity to apply computational approaches common to chemical en-
gineering curricula in concert with analysis-driven research.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

106 New Directions for Chemical Engineering

FIGURE 4-7 A side-by-side comparison of two leading estimates of global greenhouse gas emis-
sions from the global food system. NOTE: Crippa and colleagues (2021) include emissions from
a number of nonfood agricultural products, including wool, leather, rubber, textiles, and some bio-
fuels. Poore and Nemecek (2018) do not include nonfood products in their estimate of $13.6 billion
metric tons of CO2e. This may explain some of the difference. SOURCE: Our World in Data
(2021b).

Protein-rich products—including dairy, eggs, fish, meat, and poultry—account


for an increasing amount of the human diet, with demand growing rapidly in low- and
middle-income countries. In higher-income countries, protein-rich foods are often avail-
able in excess and make up a substantial fraction of food-related consumer waste (Spiker
et al., 2017). Reducing net consumption of animal-based protein is likely to be an im-
portant contributor to improving sustainability and curbing methane emissions. Achieving
this reduction will require, among many other efforts, both the development of substitutes
for animal-based proteins and, in cases in which animals are still harvested for meat prod-
ucts, understanding of methods for mitigating related emissions. Exciting avenues of re-
search in the latter area include understanding of the microbiome interaction of agricul-
tural animals with feed. Examples include recent research wherein dairy cows were fed
small amounts of seaweed, which reduced overall per-animal methane emissions (Roque
et al., 2021).

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Sustainable Engineering Solutions for Environmental Systems 107

Beyond the development of improved agricultural practices for plants and ani-
mals, improved water usage is a core element of the role of chemical engineering in the
future of agriculture. Anthropogenic climate change, among other forces, is already shift-
ing the global and regional balances of water resources. Systems-level approaches, in-
cluding life-cycle assessment, will play a critical role in meeting this local, regional, and
global challenge. Technological approaches to the use of water in agriculture are a prime
example for which modular processes will be vital, as deployment is likely to require
many units operating in a localized and autonomous manner rather than large, centralized
facilities. Many of the same technical challenges apply to reducing the energy use associ-
ated with agriculture.
The judicious application and effective recycling of reactive nutrients, tradition-
ally exemplified by ammonia and urea as fertilizer but also including phosphorous and
potassium, is another area presenting multiple chemical engineering challenges. The dead
zone in the Gulf of Mexico (and in bodies of water around the globe) due to fertilizer
runoff is a legacy issue presenting an opportunity for chemical engineers to contribute to
environmental protection through targeted applications of nutrients to enhance nitrogen
management. Chemical engineers can also pioneer more efficient production of ammonia
(beyond the Haber–Bosch process), which has reemerged as a grand challenge and has
been the focus of intense research in the past decade (e.g., Boerner, 2019; Garnier, 2014;
Smith et al., 2020). Any potential technology for this process needs to ultimately produce
ammonia in very large quantities to be globally relevant. Addressing challenges in catal-
ysis, reaction engineering, and process development overall will be critical to this end.
Some nutrients beyond ammonia/reactive nitrogen are finite relative to the expected
growth in agriculture, and their availability today still relies on mining, as was the case
with ammonia in the late 19th century. These finite nutrients include phosphorous and
potassium, which, along with ammonia, are commonly applied to feed and food crops.
Meeting challenges in their economical and efficient recovery along multiple supply
chains will be critical as agricultural demands increase.
Modern agriculture and food processing produce massive amounts of industrial
and postconsumer waste. For example, nut processing results in the production of remark-
able amounts of waste lignocellulose. These waste streams offer a direct and immediate
pathway for the needed development of biofuels, biochemicals, and biomaterials—areas
in which chemical engineers already play a leadership role. Going beyond anaerobic di-
gestion or simple combustion of organic waste to produce products of higher value than
methane (often used for combustion) or enable direct production of power is an area ripe
for immediate contributions from chemical engineers.
The above discussion includes but a sampling of ways to improve modern agri-
cultural practices in which chemical engineers can play pivotal roles. There are undoubt-
edly many others. The problem of food production is inherently global in nature, and
systems-level thinking at multiple scales—a hallmark of the chemical engineering pro-
fession—will be critical to enable positive change toward a more sustainable agricultural
system.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

108 New Directions for Chemical Engineering

Reinventing Agriculture

Today’s agricultural practices are inherently land, water, and nutrient intensive.
Almost universally, agriculture today is practiced in two dimensions and with batch pro-
cesses. The use of land means that agricultural productivity scales with surface area,
whereas most chemical engineering manufacturing processes, as described further in
Chapter 6, scale volumetrically. Moreover, crops and animal products are typically not
harvested continuously but in a manner that depends on the time of year, with many ex-
ternal factors (e.g., weather, climate, disease, fire, drought) affecting the process conver-
sion efficiency, yield, and rate. For many generations, these observations appeared to be
intrinsic to the production of food and therefore barely worth stating. However, now is
the time to consider thoughtfully how immutable these notions are. Examples of this shift
in thinking can already be seen with the surge of companies focused on the production of
non–animal-based meat, dairy, egg, and oil substitutes. In various ways, these companies
are exploring the concept of transitioning agriculture from an areal farming practice to a
volumetric, continuous industrial manufacturing process. The outcomes of this approach
will likely have effects across the entire food and agriculture supply chain.
One example—among many—of plant-based products used today is lipids in oil-
producing crops. The production of palm oil, for example, is quite energy intensive, lead-
ing to substantial land-use change effects, especially in tropical regions. The application
of synthetic biology, metabolic engineering, bioprocess development, and analysis-driven
research to produce oil products that can displace plant-based oils in a cost-effective and
sustainable manner will likely have positive effects on reforestation and preservation of
the natural environment (Parsons et al., 2020). Producing food in a continuous, industrial
manufacturing setting could also vastly increase the number of feedstocks available for
food production beyond CO2, NH3, and sunlight. As discussed in the section on feedstock
flexibility in Chapter 6, the use of waste-based feedstocks for valuable products, espe-
cially for food production enabled by biological and catalytic transformations pioneered
by the chemical engineering community, offers a clear path toward a more circular carbon
and nitrogen economy that is more sustainable than today’s agricultural practices, and
chemical engineers will play a critical role in this much-needed transition.

Food Engineering and Processing and Storage

Supplying the world’s population with food that is nutritious, affordable, and sus-
tainable is a global challenge. Societal-scale pressures associated with climate change and
population growth and distribution will demand substantial changes in the world’s food
sources in the coming decades. Despite the large number of chemical engineers working
in the food industry, food science has historically not been an area of strong emphasis in
U.S. chemical engineering research. Multiple opportunities exist for chemical engineers
to play a critical role in the future of food engineering and processing. These topics share
several features with more traditional chemical manufacturing, including the need to op-
erate at very large scales at low cost and with stringent safety, sustainability, and quality

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Sustainable Engineering Solutions for Environmental Systems 109

standards. Any proposed solution that fails to meet any of these criteria will be ineffective
in leading to real change in the world’s food system. Of critical importance is to evaluate
the cost and sustainability of new concepts using systems-level life-cycle assessment
while taking account of the role of local environmental conditions and traditions.

Food Engineering

Chemical engineers have led in adapting the tools of molecular and systems biol-
ogy for applications in diverse areas; however, the application of these methods to food
engineering is in its early stages. Although genetic modifications can be controversial,
molecular biology has the potential to enable enormous advances in crop science. An ex-
ample is golden rice, which could eliminate vitamin A deficiencies in large populations.
The pursuit of research and development (R&D) in chemical engineering with plants ra-
ther than single-celled organisms, such as yeast, poses both technical and nontechnical
challenges. The cell walls in plants make extracting molecules from plant cells more chal-
lenging relative to animal or microbial cells, and the longer life cycles of plants compared
with single-celled organisms or model organisms such as C. elegans create logistical chal-
lenges for research. Nevertheless, the potential for chemical engineers to combine their
skills in systems thinking with a rich set of biomolecular tools to improve food-related
plants is great, particularly if they make concerted efforts to work in concert with related
fields of biological and agricultural science.
In addition to improvements in food at the molecular level, the coming decades
are likely to see significant changes in the macroscopic sources of food, particularly pro-
tein and lipid sources. Meat-free protein alternatives are already becoming mainstream in
the United States, but considerable advances will be required if these kinds of foods are
to be used on a global scale. Scaling up the volume of products that involve processing
solids and liquids at low cost and with high reliability has been a core focus of chemical
engineering since the field’s inception, and chemical engineers will have a major impact
if they can adapt their existing expertise in chemical engineering and biotechnology to the
challenge of producing new foods at scale. Valuable opportunities exist for collaboration
between chemical engineers and researchers who are pioneering initial demonstrations of
“lab-grown” foods on small scales and for diversification in such areas as self-assembly
into food-relevant applications.
Although the impact of new food sources will accelerate, the scale of primary
crops and agriculture will remain vast. Chemical engineers can play a role in continuing
developments in precision agriculture, defined as the precise delivery of fertilizer, pesti-
cides, and herbicides to minimize environmental impact. In the past, only a limited con-
nection has existed in this area between the ambitions and needs of industry and farmers
and the scope of academic research in the U.S. chemical engineering community.

Food Processing and Storage

While producing food in better ways is critical to meeting global food needs, re-
ducing food waste could also have an enormous impact. The challenges associated with

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

110 New Directions for Chemical Engineering

food waste differ around the world: broadly, in higher-income countries, food waste is
dominated by high-quality food that is not completely used by consumers, whereas in
low- and middle-income countries, food waste more often results from a shortfall in the
delivery of food from its original source to consumers. Consequentially, although signif-
icant advances in packaging and food treatment to prolong shelf life have the potential to
make a major impact in both low- and middle-income countries and higher-income coun-
tries, there can be no single solution to the overall challenge of reducing food waste. Even
so, chemical engineers have the potential to enable important advances in this globally
important area. To ensure impact in low- and middle-income countries, researchers need
to focus on ultra–low-cost solutions and appropriate technologies that account for local
context and cultural traditions. In higher-income countries, the sustainability of food pro-
duction and packaging is a major concern that can be addressed through the use of non-
food containers to hold food during shipping or storage and edible coatings that can be
applied to food to prolong shelf life. Chemical engineers are well suited to the challenging
task of applying systems-level life-cycle assessment to the development of solutions in
these areas.

UNDERSTANDING AND IMPROVING AIR QUALITY

Even when air pollutants are at very low levels—and despite the progress that has
been made in this area since 1970—air pollution continues to harm the environment, af-
fect the climate, and harm human health (Figure 4-8; EPA, 2021a). On a global scale, the
World Health Organization (WHO) estimates that about 4.2 million people died prema-
turely worldwide in 2016 because of outdoor air pollution in cities and rural areas, and 91
percent of those deaths occurred in low- and middle-income countries (WHO, 2018). The
causes of death were heart disease, stroke, chronic obstructive pulmonary disease, and
lung cancer. Additionally, the WHO estimates that of the 3 billion people who still use
solid fuels (such as wood, crop wastes, charcoal, coal, and dung) and kerosene for cook-
ing, about 3.8 million died prematurely in 2016 from diseases attributable to indoor air
pollution caused by these types of cooking practices (WHO, 2016). Globally, air pollution
is the fourth-highest contributing risk factor for death, behind high blood pressure, smok-
ing, and high blood sugar (Figure 4-8). Indoor air quality3 is far less well studied and not
as uniformly regulated as outdoor, yet people in the United States and other high-income
countries spend more than 85 percent of their time indoors (EPA, 2018). Efforts to address
indoor air pollution are complicated by the highly heterogeneous building and ventilation
types that exist across even relatively small regions. Building codes related to ventilation
have focused thus far on energy efficiency and less on air quality, and in many cases,
improvements related to energy efficiency have led to less air exchange and thus poorer
indoor air quality (EPA, 2020).

3
The National Academies’ Committee on Emerging Science on Indoor Chemistry is currently
examining the links among chemical exposure, air quality, and human health in indoor environ-
ments.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Sustainable Engineering Solutions for Environmental Systems 111

FIGURE 4-8 Number of deaths caused by various risk factors globally in 2017. Air pollution is
the fourth-largest risk factor, contributing to 4.9 million deaths annually. SOURCE: Our World in
Data (2019).

The six most common outdoor air pollutants are particulate matter (PM), ground-
level ozone (O3), sulfur dioxide (SO2), nitrogen dioxide (NO2), airborne lead, and carbon
monoxide (CO). While many of these pollutants are still present in the air, their concen-
trations in the United States have declined significantly since 1970 (Figure 4-9). In addi-
tion, 187 toxic air pollutants that can cause cancer and birth defects—including such
chemicals as acetaldehyde, asbestos, cadmium and chromium compounds, benzene, di-
oxin, epichlorohydrin, and methanol—are listed in the U.S. Clean Air Act (42 U.S.C. §
85 [1955]); some of these chemical species form atmospheric aerosols, while the rest re-
main in the gaseous phase. Another air pollution challenge is to protect stratospheric
ozone, which was achieved to a great extent by banning chlorofluorocarbons (CFCs), hy-
drochlorofluorocarbons (HCFCs), and other harmful compounds (UNEP, 2020).
Still another challenge is posed by acid rain, caused by the reaction of atmos-
pheric SO2 and NOx with water, oxygen, and other chemicals to form nitric and sulfuric
acids. Acid rain can turn lakes and streams acidic, harming fish and wildlife, and flowing
through soil can leach aluminum and remove minerals and nutrients from the soil. The
impact of acid rain in the United States and Europe has largely been reduced by the reg-
ulation of sulfur and nitrogen emissions. Other air pollutants include CO2 and other
GHGs, which are discussed further in the context of energy (Chapter 3) and the circular
economy (Chapter 6).

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

112 New Directions for Chemical Engineering

FIGURE 4-9 Graph displaying the gradual decrease in the emissions of three major air pollu-
tants in the United States over time. SOURCE: Our World in Data (2021a).

Atmospheric aerosols are suspensions of microscopic and inhalable solid or liq-


uid (typically, aqueous) particles in the air. About 90 percent of aerosols, by mass, occur
naturally (e.g., from volcanic eruptions, ocean phytoplankton emissions, sea and salt
sprays, forest and grassland fires, or dust aerosols from wind erosion of arid land; Voiland
and Simmon, 2010). Anthropogenic aerosols include those from fossil fuels (e.g., those
composed of sulfates) and agricultural waste burning, road dust, cooling towers, industrial
processes, and deforestation fires (which are composed primarily of carbon). Aerosol con-
centrations can be as high as 106–107 particles per cm3. The typical size of aerosol particles
is 1 nm to 10 µm; they are classified as PM10 when their diameter is less than 10 µm and
PM2.5 when it is less than 2.5 µm. Aerosols greater than 10 µm in diameter settle to the
ground in a matter of hours because of gravity, but those less than 1 µm in diameter remain
in the atmosphere for weeks and are ultimately removed by impacting the surface of the
Earth (dry deposition) or via precipitation (wet deposition). Aerosols are characterized as
either primary (i.e., emitted as particles at the source) or secondary (i.e., formed from
gaseous precursors; Haywood, 2016). This class of pollutants has garnered considerable
attention with respect to indoor air quality during the COVID-19 pandemic because of the
ability of exhaled virus-laden aerosols to cause infection.
Aerosols have both direct and indirect effects on climate. Models estimate that
direct cooling effects have counteracted about half of the warming caused by GHGs since
the 1880s (Voiland and Simmon, 2010). Indirect effects include nucleating the formation

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Sustainable Engineering Solutions for Environmental Systems 113

of clouds and controlling their numbers and lifetimes. For humans, inhalation of PM and
its subsequent deposition in the lungs and even entrainment into the bloodstream are the
main causes of aerosol-related health problems. Specifically, PM2.5 can cause cardiovas-
cular, mental, dermatological, and reproductive health problems (McNeill, 2020).
A major challenge for improving air quality is bridging the molecular scale of a
chemical reaction and the massive scale of an atmospheric model. Some of the founda-
tions of chemical engineering education (e.g., transport phenomena, thermodynamics, and
chemical reaction engineering), along with a systems approach focused on complexity
and scale, can be applied directly to solving environmental engineering problems. In many
respects, the atmosphere (more generally, the environment) functions as a large chemical
reactor in which pollutants are transported, mixed, and transformed while being distrib-
uted across the atmosphere, land, and water. More specifically, the extremely wide ranges
of the temporal and spatial scales of the atmosphere are well suited to a chemical engi-
neering modeling approach. Spatial scales span about 15 orders of magnitude, from the
nanometer range of molecular dimensions to thousands of kilometers, and temporal scales
span about 12 orders of magnitude, from the millisecond range of fast chemical reactions
to the thousand-year scale of waste disintegration.
The chemical engineering profession has made key contributions to improving
air quality, including the development of catalytic converters for vehicles (Box 4-1),
cleaner-burning fuels, flue gas desulfurization and selective catalytic reduction systems
for NOx conversion to nitrogen gas and water, wet flue gas scrubbing methods, and better
coal gasification technologies (Haywood, 2016). Furthermore, most chemical industries
are required to control their aerosol emissions, and the principal equipment for that control
includes cyclonic separators, fabric filter collectors (baghouses), electrostatic filters and
precipitators, and wet scrubbers, all of which were developed with the help of chemical
engineers.
Moving forward, chemical engineers have an opportunity to minimize and elim-
inate the formation of air pollutants at the source. When “benign by design” is not possi-
ble, chemical engineers can play a role in minimizing or eliminating emissions of these
pollutants into the environment or even apply treatments to make the emissions safe for
the environment. Ultimately, it may even be possible to treat the environment to remove
air pollutants (AIChE, 2017; Sánchez, 2019). Specific opportunities for chemical engi-
neering include the following:

 reduction of emissions from smokestack and exhaust tailpipes through im-


provements in engines and industrial plants (e.g., engines with better com-
puter control and fuel economy, and multivalve engines), advanced monitor-
ing and diagnostics, enhanced maintenance, development of catalytic
pathways to destroy pollutants, use of cleaner fuels (e.g., low-sulfur and re-
newable fuels) and filters (e.g., diesel filters), and development of electric
and fuel cell vehicles;
 better management of waste, such as the use of environmentally friendly sol-
vents, and equipment for capture of methane gas emitted from waste sites and
its use as biogas (e.g., Cao et al., 2018b);

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

114 New Directions for Chemical Engineering

BOX 4-1
Control of Nitrogen Oxide Emissions from Diesel Engines

Energy-efficient diesel-powered engines have enabled the growth and harvesting of the
food supply, the mobility of people and goods over large distances, and the construction of
modern infrastructure. The higher air-to-fuel ratio in diesel engines leads to lower greenhouse
gas emissions relative to gasoline-powered engines, but removing NOx and soot from diesel
exhaust poses formidable challenges. Modern exhaust treatment systems integrate several cata-
lysts and filters to trap carbon particulates, oxidize CO and organics, and reduce NOx. For gas-
oline engines, three-way precious metal catalysts have succeeded spectacularly in mitigating
emissions, representing one of the most meaningful contributions of catalysis and chemical en-
gineering to human health. But these catalysts are ineffective in removing NOx from air-rich
diesel effluent. Particulate and NOx emission limits required the urgent development of alterna-
tive after-treatment strategies for diesel engines. A timely combination of advances in materials
design, mechanistic insights, and systems control and optimization led to the successful deploy-
ment of Cu-exchanged chabazite (Cu-CHA) zeolites as catalysts for urea-based selective cata-
lytic reduction (SCR) of NOx.
These breakthroughs required effective, robust, and regenerable materials expected to func-
tion over broad ranges of temperature and effluent flow rates and compositions, including brief
excursions to about 900 °C, and devices able to meet emission targets for 100,000–400,000
miles under harsh hydrothermal environments containing strong catalyst poisons (S, P) derived
from fuels and lubricants. The relevant reaction networks are complex (with more than 20 par-
allel and sequential steps). They include urea decomposition to NH3, which then selectively
reduces NOx to N2 and H2O. This reaction must occur without parallel reactions of NH3- with
O2, which is about a thousand times more concentrated than NOx. Research led to the develop-
ment of catalysts with isolated Cu cations grafted onto zeolite exchange sites. The microporous
zeolite scaffold confines and protects Cu active sites from contact with fuel components that
form carbonaceous residues, while at the same time providing acid sites for additional synergis-
tic catalytic functions. These developments were enabled by engineering concepts in chemical
kinetics and spectroscopy, and were leveraged by using theoretical methods to show how the
unique reactivity of the Cu centers reflects their solvation and mobilization by NH3 to form the
Cu dimers required for the O2 activation steps in the catalytic redox cycles. CHA zeolites also
proved to be remarkably robust in the harsh hydrothermal environments that degrade most other
zeolites. These innovations in catalyst design led to the deployment of Cu-CHA catalysts for
urea-SCR, representing the first large-scale implementation of an exhaust after-treatment strat-
egy using catalysts containing earth-abundant elements.
In practice, these SCR catalysts are deployed in compact devices that seamlessly integrate
them with catalytic particulate filters, the CO/hydrocarbon oxidation catalysts, and other com-
ponents to form a diesel after-treatment system (see Figure 4-1-1). Robust process models and
control schemes are essential for the system to function, and they require accurate rate and se-
lectivity data, together with thermodynamic, hydrodynamic, and kinetic descriptions. Today,
these modular chemical plants are deployed in millions of diesel-powered vehicles. The vehi-
cles’ efficient engines have helped decrease CO2 emissions, and their exhaust streams no longer
contaminate the environment with NOx and particulates. This achievement represents the result
of a combination of novel materials and core chemical engineering concepts, made to work
through the systems-based approach that characterizes chemical engineering as a discipline.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Sustainable Engineering Solutions for Environmental Systems 115

FIGURE 4-1-1 Emissions of nitrogen oxides (NOx) from diesel vehicles are converted to
environmentally benign products using urea-selective catalytic reduction (SCR) technology
based on copper-exchanged zeolite catalysts that have high reactivity at low temperatures and
are durable over long lifetimes. SOURCE: Purdue University/Mo Lifton.

 use of clean energy solutions for power generation, cooking, heating, and
lighting;
 use of strict emissions control in waste incineration sites; and
 capture of CO2 emissions.

In addition to these mitigating actions, chemical engineers can contribute to the funda-
mental understanding of aerosol formation, aerosol dynamics in the atmosphere, and their
chemical characterization (e.g., Seinfeld, 1991; Seinfeld and Pandis, 2016). Sensor tech-
nologies that would allow better monitoring, chemical characterization, and knowledge
of the spatial distributions of aerosols would also help address air pollution. Finally, the
use of data science and multiscale models to illustrate atmospheric chemistry and the in-
corporation of process modeling to bridge the gap between observations and theory are
opportune areas for the engagement of chemical engineers.

CHALLENGES AND OPPORTUNITIES

Food, energy, and water are highly interconnected, and solutions in this complex
system need to be both environmentally sustainable and economically viable. A continued
increase in the global population will lead to increased resource demands, posing a broad
set of challenges for chemical engineers to address in the coming decades. Across all these

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

116 New Directions for Chemical Engineering

challenges, interdisciplinary and cross-sector collaborations will be critical, as will coor-


dination across the federal agencies with responsibilities in these areas.
Chemical engineers can support water conservation by both designing higher-
efficiency processes and developing methods for using alternative fluids to freshwater.
Specific research opportunities range from better understanding of the fundamentals of
water structure and dynamics to the development of membranes and other separation
methods. In the domain of water use and water purification, U.S. chemical engineers
would benefit from collaborations with civil engineers and other scientists and engineers
in arid regions that have more experience with desalination.
Global pressures associated with climate change and population growth will re-
quire substantial changes in the world’s food sources, a need that chemical engineers can
help address through enabling technologies. Specific opportunities for chemical engineers
include precision agriculture, non–animal-based food and low-carbon-intensity food pro-
duction, and reduction or elimination of food waste. Advanced agricultural practices de-
signed to improve yield while reducing demand for both energy and water will require
collaboration with other disciplines, as well as systems-level approaches such as life-cycle
assessment. A particularly valuable opportunity for collaboration is with researchers who
are pioneering initial demonstrations of lab-grown foods on small scales.
Chemical engineers have an opportunity to help in improving air quality by ad-
vancing understanding of the nature and physics of aerosol particles and applying separa-
tion technologies, as well as the molecular- and systems-level understanding that will be
necessary to address this global challenge. Atmospheric science is already an interdisci-
plinary field that includes chemistry, physics, meteorology, and climatology, making it a
promising area in which chemical engineering can contribute through increased collabo-
ration.

Recommendation 4-1: Federal research funding should be directed to both basic and
applied research to advance fundamental understanding of the structure and dy-
namics of water and develop the advanced separation technologies necessary to re-
move and recover increasingly challenging contaminants.

Recommendation 4-2: To minimize the land, water, and nutrient demands of agri-
culture and food production, researchers in academic and government laboratories
and industry practitioners should form interdisciplinary, cross-sector collaborations
focused on the scale-up of innovations in metabolic engineering, bioprocess develop-
ment, precision agriculture, and lab-grown foods, as well as the development of sus-
tainable technologies for improved food preservation, storage, and packaging.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

5
Engineering Targeted and Accessible Medicine

 Chemical engineers have been involved for at least a century in reactor design and
separations and more recently in cell engineering, formulations, and other aspects
of drug manufacturing, and they have the potential to make many more contribu-
tions to health and medicine at scales ranging from the molecular to manufacturing
facilities.
 Since the first attempts to isolate small molecules from biological organisms and
control and reengineer cell behavior, the development of biologically derived prod-
ucts has increased, with major advances resulting from recombinant DNA technol-
ogy, the sequencing of genomes, the development of polymerase chain reaction,
the discovery of induced pluripotent stem cells, and the discovery and implemen-
tation of gene editing.
 The development of disease treatments is a multidisciplinary enterprise, and chem-
ical engineers can contribute to many aspects of medicine by applying systems
biology to physiology, the discovery and development of molecules and materials,
and process development and scale-up.
 Many health disparities are a result of systemic issues that will require larger social
changes to address, but chemical engineers can develop engineering solutions that
help address disparities requiring more focused efforts.

Spending for health care in the United States is enormous (Figure 5-1) and has
grown over the years, from 5 percent of gross domestic product (GDP) in 1960 to nearly
18 percent in 2019. The largest fractions of this total are for hospital care (31 percent) and
physician and clinical services (20 percent). The costs of prescription drugs have skyrock-
eted as well, despite a decrease in retail drug prices, and represent a growing share of
health care costs (e.g., 3.8 percent in 2018, compared with 5.7 percent in 2019; CMS,
2020; Martin et al., 2021). On the other hand, revenue from the biotechnology sector,
including drugs and agricultural and industrial products, is a significant contributor to the
national economy, accounting for more than 2 percent of GDP in 2012 (Carlson, 2016).
Therapeutic drug molecules are either small molecules (molecular weight [MW]
below 1000 Da) synthesized chemically, or biologic molecules of any MW produced by
biological (cell-based) methods. Drug development today includes both therapeutic clas-
ses, although small molecules now account for roughly 70 percent of U.S. Food and Drug
Administration (FDA) drug approvals and about 60 percent of U.S. market share (Makur-
vet, 2021). This is the case in part because, for small molecules, the cost of synthesis is
lower and the delivery of therapeutic material is simpler. Biologics have the advantages

117

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

118 New Directions for Chemical Engineering

FIGURE 5-1 Estimated total annual U.S. biotechnology revenues, 1980–2017. Bars show data,
while shaded areas are a numerical model based on annual growth rates. The inset shows annual
growth rates for subsectors (crops, biologics, industrial, aggregate) between 2000 and 2017.
SOURCE: Bioeconomy Capital (2018).

of offering highly selective targeting and, from an intellectual property perspective, being
more difficult to copy. However, they can be much less stable; are unable to cross many
biological barriers, including the blood–brain barrier; and are typically inactive when ad-
ministered orally. Both small (chemical) and biologic molecules are likely to continue
sharing the medical discovery and therapeutic markets, with the choice between the two
depending on the application and disease at hand.
This chapter describes the role of chemical and biochemical engineering in hu-
man health applications, with a particular focus on biomolecular engineering. Following
an overview of the role of biomolecular engineering in health and medicine, the chapter
describes opportunities for chemical engineers to advance personalized medicine; im-
prove therapeutics; understand the microbiome; design materials, devices, and delivery
mechanisms; and develop hygiene technologies. Finally, the chapter examines how chem-
ical engineers can contribute to addressing health disparities that result from societal in-
equity and reducing the costs of therapeutic treatments.

THE ROLE OF BIOMOLECULAR ENGINEERING


IN HEALTH AND MEDICINE

Biomolecular engineers drive the development of computational and experi-


mental techniques for identifying drug targets using multiscale modeling tools and data
science approaches (see Chapter 8), and then employ organ-on-a-chip and/or tissue model
development to replicate in vivo behavior. Small-molecule drug development and manu-
facturing still depend on chemical engineers who understand colloid science, particle dis-
solution, and multicomponent phase behavior to design formulations. For small-molecule

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Engineering Targeted and Accessible Medicine 119

discovery, automated, high-throughput screening of potential drugs will continue to inte-


grate chemical synthesis, modeling of binding surfaces, and automation. Scale-up and
manufacturing of both small molecules and biologics also rely on traditional process de-
velopment skills of biomolecular engineers in addition to novel approaches, such as the
integration of production and purification.
The origins of biomolecular engineering as applied to therapeutics can be traced
back centuries to the use of such organisms as baker’s yeast (Saccharomyces cerevisiae)
to ferment grape juice and leaven bread. The first formal application of engineering prin-
ciples to biological processes is often recognized as the production of penicillin in the
1930s and 1940s. The bacteriologist Sir Alexander Fleming is credited with discovering
in 1928 the ability of an accidental mold growth to inhibit the growth of staphylococcus
colonies, naming the active component of the mold penicillin. After significant work by
a collaborative team of scientists and engineers to produce the drug at scale, this seren-
dipitous discovery led to a major medical advance in the treatment of sepsis. Other similar
compounds from molds led to the field of antibiotics, which of course remains critical in
medicine to this day.
Despite the eventually profound impact of the discovery of penicillin, it was more
than a decade later (in the late 1930s, as the world was poised to enter World War II)
when Ernst Chain, Howard Florey, and Norman Heatley at Oxford identified the potential
agent from Penicillium notatum (Florey, 1949). They worked to produce it economically
by extracting it from the broth of the growing mold. Because of the intense bombings that
began in the United Kingdom, Florey and Heatley visited the United States to try to ad-
dress the needs for large-scale production; the collaborative transatlantic effort of bio-
chemists, bacteriologists, physicians, and chemical engineers (and more) made it possible
to produce penicillin biosynthetically, as chemical synthesis appeared to have little chance
of success. Even biological routes were difficult—hundreds of scientists and engineers
worked to create enough penicillin to treat all allied troops wounded in the D-Day inva-
sion of Europe (~250,000 patients/month). Notably, Margaret Hutchinson Rousseau—the
first woman to earn a PhD in chemical engineering and the first female member of the
American Institute of Chemical Engineers (AIChE)—designed a deep-tank fermentation
process for growing the surface-growing mold to enable penicillin production on a large
scale. Among other significant advances were media development, strain improvement,
and purification and recovery innovations that resulted in a 5,000-fold improvement in
yields to 50 g/L (ACS and RSC, 1999; Gaynes, 2017). These efforts illustrate how classic
chemical engineering concepts were applied to human health as early as the mid-20th
century, ultimately contributing to the broader convergence of human health and engi-
neering.
Since the first attempts to produce small molecules from biological organisms,
the development of biologically derived products has increased. As described in Box 2-2
in Chapter 2, biochemical and biomolecular engineering and production of medicines
have continued to grow, with major advances resulting from the development of recom-
binant DNA technology, the sequencing of genomes, the development of polymerase
chain reaction (PCR), the discovery of induced pluripotent stem cells, and the discovery

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

120 New Directions for Chemical Engineering

and implementation of CRISPR (clustered regularly interspaced short palindromic re-


peats) (see Figure 5-2). The first recombinant protein therapeutic approved by the FDA
was insulin (5.8 kDa) in 1982, derived via production in Escherichia coli (E. coli) with
subsequent purification and refolding (Carter, 2011; Leader et al., 2008). Since the 1980s,
many more proteins produced in E. coli have been approved, and they are still produced
in this simple bacterium. However, mammalian cells, particularly Chinese hamster ovary
(CHO) cells, have been used to produce around 70 percent of all recombinant therapeutics
since the late 2010s (Jayapal et al., 2007; Wells and Robinson, 2017).
Today, monoclonal antibodies (mAbs) represent the highest volume of sales and
a major focus of drug development. These therapeutics can bind to a specific antigen and
have been developed as highly specific treatments for such diseases as cancer, asthma,
arthritis, Crohn’s disease, migraines, and infectious diseases (Lu et al., 2020; Wells and
Robinson, 2017). The first therapeutic mAb was muromonab-CD3, approved by the FDA
in 1986, which was utilized in the treatment of transplant rejection until 2011. Humaniz-
ing antibodies, and later producing fully human mAbs as part of a library for protein en-
gineering, rapidly accelerated the development of therapeutic antibodies (Jones et al.,
1986; Tsurushita et al., 2005). Over the last 25 years, many antibodies have entered the
market (Figure 5-3), with 61 mAbs being approved by the FDA for clinical use at the end
of 2017 and an additional 18 new antibodies being approved from 2018 to 2019 (Lu et al.,
2020).

FIGURE 5-2 Biotechnology discoveries and notable drug approvals from 1973 to 2014. NOTE:
CHO = Chinese hamster ovary; CRISPR = clustered regularly interspaced short palindromic re-
peats; ESC = embryonic stem cell; FDA = U.S. Food and Drug Administration; MS = multiple
sclerosis; PCR = polymerase chain reaction; RNAi = RNA interference; TALENs = transcription
activator-like effector nucleases; ZFNs = zinc-finger nucleases. SOURCE: Wells and Robinson
(2017).

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Engineering Targeted and Accessible Medicine 121

FIGURE 5-3 Trademarked antibody therapies contribute to a large fraction of the drug market,
with major advances resulting from humanizing antibodies in the late 1990s. As indicated by the
arrows, the several anticancer antibodies introduced in the early 2000s represent 5 of the top 10
bestselling antibody drugs in 2018. NOTE: mAb = monoclonal antibody. SOURCE: Lu et al.
(2020).

Modern biomolecular engineering is very much at the intersection of chemical


engineering and molecular biology, biochemistry, materials science, and medicine. The
role of chemical engineers in addressing problems in health and medicine continues to lie
in process and scale-up, as well as in discovery and development of molecules and mate-
rials with pharmaceutical and medical applications. For example, the development of the
small-molecule therapeutics Remdesivir and Molnupiravir highlights the value of bio-
molecular engineering in combatting the COVID-19 pandemic. However, the repercus-
sions of the pandemic also serve as a reminder of the importance of the availability and
transport of raw materials, as well as the critical importance of scale-up of manufacturing
where medicines are needed and the fact that even in higher-income countries, health dis-
parities can result in needless deaths.
The next 20 years of chemical and biomolecular engineering will feature oppor-
tunities in personalized medicine; advances in the engineering of biologic molecules, in-
cluding proteins, nucleic acids, and such entities as viruses and cells; growth at the inter-
face between materials and devices and health; the use of tools from systems and synthetic
biology to understand biological networks and their intersections with data science and
machine learning; development of the next steps in manufacturing; and the use of engi-
neering approaches to address health equity and access to health care.

PERSONALIZED MEDICINE

An important area in which chemical engineering can contribute to the future of


medicine is personalized medicine, which denotes the tailoring of treatments for individ-

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

122 New Directions for Chemical Engineering

ual health, including cell and gene therapies for patient-specific disorders, as well as pre-
ventive care regimens based on an individual’s genetic-disease propensity or biomarker
presence. As more is understood about the role of genetics and environment in future
disease, a continuing emphasis on personalized medicine is likely, with a clear role for
chemical engineers in the development of appropriate models at scales ranging from
atomic to systems for target discovery and the design of medicines based on those targets.
The application of data science and modeling represents another opportunity for chemical
engineers to contribute in this space.

Small-Molecule Manufacturing

For small molecules that serve as active pharmaceutical ingredients (APIs), a


number of innovations could contribute to advances in applications during manufacturing.
A detailed discussion of the importance of small molecules and related research opportu-
nities for chemical engineers can be found in the recent report of the National Academies
Innovations in Pharmaceutical Manufacturing on the Horizon: Technical Challenges,
Regulatory Issues, and Recommendations (NASEM, 2021b). This section focuses on the
application of additive manufacturing applied to personalized medicine. One such appli-
cation is polypharmacy, which entails product formation through the use of 3-D printing.
This method is of growing interest because in the United States, more than 40 percent of
those over age 65 use five or more medicines, and 5 percent use more than ten (Kantor et
al., 2015), creating complexity and confusion for patients. Polypharmacy is also common
in Europe and Australia (Schöttker et al., 2017). Beyond the pharmaceutical implications
of understanding how these medicines may interact in treating disease, the ability to de-
liver medicines appropriately in a safe and efficacious way and to ensure that the dosage
and route of delivery are optimal is a need that chemical engineers and others can address
through both systems biology and advances in manufacturing. In addition, the ability to
combine multiple medicines into one delivery on a patient-demand basis would likely
improve health care outcomes.
The use of 3-D manufacturing for APIs poses many technical challenges, includ-
ing physics-based modeling of the fluid mechanics of drop formation and extrusion for
liquid and powder solidification, as well as challenges involved in the extrusion process
itself (NASEM, 2021b). Improving and simplifying the delivery of multiple drugs is a key
opportunity for chemical engineers.
Additional opportunities for improvements in small-molecule manufacturing lie
in predictive analytics and in the emerging field of translational medicine bioinformatics.
Pharmaceutical companies use predictive analytics to develop algorithms that predict op-
timal conditions for robust production by leveraging large historical datasets. For low-
dose or limited-duration drugs for which limited scale-up data are available, Bayesian
statistics are used to estimate acceptable parameter ranges and product specifications.
Translational medicine bioinformatics is an emerging field that draws on the fundamen-
tals of chemical engineering and has the potential to transform personalized medicine. By
analyzing raw data from DNA sequencing, patterns can be detected for specific mutations

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Engineering Targeted and Accessible Medicine 123

or genes to help predict which small-molecule drugs a patient might respond to. This field
holds the potential to turn biomarkers into clinical diagnoses.
An additional need is the implementation of plug-and-play analytics, allowing
rapid shifts from one production process to another (e.g., for small-batch production), and
more uniformity in data formats for analytical devices across vendors. As supply chains
continue to be stretched, the widespread adoption and implementation of digital twinning,
or simulated process models, to predict and plan for supply chain issues that affect down-
stream challenges (e.g., the availability of specific vials or packaging in a given number
of months) will also be beneficial for biotechnology and pharmaceutical production.

Cell-Based Therapies

Another aspect of personalized medicine is cell-based therapies—most important,


CAR (chimeric antigen receptor) T cell immunotherapy. Cancer cells typically carry a
ligand that blocks a specific receptor (PD-1) on the cell surface of T lymphocytes, render-
ing this protective part of the human immune system ineffective, and thereby enabling
cancer cells to evade the body’s natural defenses. CAR T cells are specially engineered
to express CAR with a high affinity for tumor antigens. Immunotherapy—often termed
immuno-oncology—has now become a mainstay of cancer treatment, with cell therapies
representing the largest growth area and immunomodulation a close second. To some de-
gree, the currently available therapies are influenced by what is known to be successful,
with 22 cell and gene therapies being approved by the FDA in 2021 (FDA, 2021) and
more than 460 targets in the current global pipeline (Xin Yu et al., 2019).
Therapeutic application of CAR T cells relies on retrieval of cells from a patient’s
blood, isolation of these cells followed by engineering in vitro, and then reintroduction to
the patient. This approach has been quite successful, particularly for leukemias; it is highly
individualized, allowing high specificity. In 2017, Kymriah® became the first FDA-
approved immunotherapy, wherein a patient’s blood is collected, and then individualized
therapy is created by introducing a virus to the T cells to allow them to express a 4-1BB
costimulatory domain to enhance cellular responses and express the CD-19 receptor.
However, the cost of CAR T therapy is high for single-patient treatments, and the prod-
uct’s stability and success can be patient specific. At present, a large number of immuno-
therapies are in clinical trials, and the annual growth rate is expected to be 15 percent for
cell therapies and nearly 30 percent for gene therapies (Mullin, 2021).
New directions for improving reliability and lowering costs are an ongoing focus
in the fields of biomolecular engineering and immunology. In particular, there is interest
in scaling down manufacturing practices that were developed for large-scale culture to
reduce the need for highly skilled workers (compared with the present process of collec-
tion of blood by clinical staff and its shipment to manufacturing sites). Also, to improve
the reliability and efficacy of treatment, key variables or biomarkers that lead to success-
fully engineered cells need to be identified, and simple analysis methods developed
(Wang et al., 2021b; Whitehead et al., 2020). The use of induced pluripotent stem cells
has the potential to reduce patient-specific cell collection, but also requires improvements

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

124 New Directions for Chemical Engineering

in both modulating and measuring differentiation and manufacturing of these surface-de-


pendent cells. Finally, T cell engineering requires viral vectors or particles that are man-
ufactured with consistent quality and functionality.

Computational Tools and Modeling to Improve Personalized Medicine

Systems biology applied to physiology is an additional avenue for chemical en-


gineers to contribute to personalized medicine. As early as the 1960s, Yeats and Urquhart
(1962) noted that biological systems and their responses to physical stimuli can be under-
stood using the engineering principles of process control. In fact, the impressive ability of
biological systems to show resilience in the face of perturbations is possible with biolog-
ical control systems that have components akin to their engineering counterparts, such as
sensors, controllers, and actuators. Diseases can be seen as a malfunction or failure of a
component in the control system, and quantitative approaches to understanding this be-
havior can benefit physicians’ ability to diagnose and treat these failures (Figure 5-4).
Chemical engineers have several opportunities to contribute to systems engineer-
ing of biological systems, such as sensor design and analysis, identification of fault de-
tection (using physiological, cellular, metabolic, or other data to identify changes in func-
tion), and process modeling to represent complex relationships for biological systems and
predict behavior (Ogunnaike, 2019). There are various approaches to developing these
models (Janes et al., 2017); as they mature, the models will play a role in the health care
industry in predicting and avoiding catastrophic health consequences for patients. Sys-
tems biology encompasses both the approach to framing questions about function and
understanding of feedback, adaptation, and dynamics to enable the development of meth-
ods for treating or resolving illness (Janes et al., 2017). A key feature of systems biology
is examination of the whole system, deconstruction or identification pathways and net-
works of biomolecules, and re-creation and integration of the information (Figure 5-5).
As one might expect, the use of large, accessible data repositories (so-called big data) is
a complementary approach to understanding biological behavior that can be used to shape
information gained from the models (see Chapter 8).

FIGURE 5-4 Feedback control paradigm for implementing personalized medicine. SOURCE:
Ogunnaike (2019).

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Engineering Targeted and Accessible Medicine 125

FIGURE 5-5 Engineering design approach applied to systems biology. Gray text provides an ex-
ample case study. SOURCE: Modified from Janes et al. (2017).

The continuous assessment of patient health—particularly for immunocompro-


mised individuals—would improve understanding of the spread of disease and enable
early-stage prevention and treatment. The potential ability to determine the status of a
patient with autoimmune or other health conditions is highly attractive, yet this is cur-
rently a daunting task that requires a simple and well-designed method for identifying
specific biomarkers, collecting leukocytes from the skin or blood, and evaluating the pres-
ence of inflammation or immune activation (Box 5-1). Other means of patient-state mon-
itoring might include the design of nanoparticle or microparticle probes targeting specific
cell types so they can be tracked and monitored over extended time periods.

ENGINEERING APPROACHES TO IMPROVING THERAPEUTICS

Vaccines as Biomolecular Therapeutic Agents

Although drug discovery is key to the development of new types of drugs, dis-
covery is only the beginning of the development phase; significant challenges arise in the
large-scale manufacture of drugs, including industrial-level product generation, purifica-
tion, and formulation. The COVID-19 pandemic has highlighted some of the challenges

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

126 New Directions for Chemical Engineering

BOX 5-1
Immune Engineering

Several opportunities exist for the integration of chemical engineering and immunology,
including cancer immunotherapies, vaccine design, and therapeutic treatments, particularly in
addressing challenges pertinent to infectious diseases and autoimmune disorders. Relevant
quantitative chemical engineering skills include

 the ability to design molecular and biomolecular agents that can stimulate or modulate
the immune system;
 the use of systems biology to predict the outcomes of manipulations of complex bio-
logical networks; and
 the ability to understand or control the flow and transport of key signaling molecules
involved in the immune response, from the innate response that yields the generation
of cytokines and interferons that are fairly universal and nonspecific, to the adaptive
immune response that is specific to a given target or set of target agents via engineered
T cells or NK cells.

Broad areas of interest include understanding of the physical and biological mechanisms
underlying how the immune system functions, applied virology, and efforts that leverage this
knowledge and engineering design to develop therapies and vaccines capable of being translated
to the clinic.
Chemical engineers are involved in the development of new vaccination concepts and the
molecular design of new delivery approaches for vaccines and biologic therapies such as anti-
bodies, including rapid drug development and commercial scaling of proteins or protein com-
ponents. Manufacturing methods that enable increased global availability or reliability of vac-
cines and therapies are vital for new technologies being introduced, from molecular to cellular
systems. Diagnostic approaches are critical to enable rapid and reliable detection of pathogens
and monitoring of biomarkers of immunity. Also of interest are systems biology, machine learn-
ing, physics-based modeling, and engineering approaches directed toward understanding innate
and adaptive immunity, with the goal of preventing and curing disease.

of bringing drugs to market. Key to the success of the rollout of COVID-19 vaccines,
messenger RNA (mRNA) vaccines from Pfizer-BioNTech and Moderna, was the gener-
ation of production lines to accommodate large-scale production and formulation. Fortu-
nately, Moderna had already gained experience in this area and built a manufacturing
facility for the development of other vaccines that were approaching early clinical trials.
Protein-based biologic drugs have been manufactured using cell-based bioreac-
tors to generate proteins for therapeutics, and a significant amount of biotechnology
know-how is based on the engineering of bioreactors that require management of titers
from microbes or mammalian cells, management of cell viability and oxygen and nutrient
levels, and extensive removal of cellular debris and waste products as part of the purifi-
cation process. Unlike these processes, the production of mRNA uses a cell-free enzy-
matic reaction; RNA polymerase is utilized in the presence of a DNA plasmid and an
RNA promoter that accelerates amplification of the DNA sequence. Additional enzymatic

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Engineering Targeted and Accessible Medicine 127

steps may also be used to modify the end groups of the mRNA. Because the process does
not use cells, the concentration of the product can be much higher than is possible in a
bioreactor for which cell density is limited and high concentrations of cells can lead to
cell death and toxicity. And because the reactions are enzymatic, the reaction rates, yield,
and purity are based on the nature of the promoter and the frequency of enzymatic errors
that lead to short oligonucleotide impurities, double-stranded RNA, and similar by-prod-
ucts that need to be removed to create a pristine product free from potentially dangerous
constituents. RNA self-replication machinery can be encoded in the sequence, thus in-
creasing rates of generation of mRNA. Purification steps have involved high-performance
liquid chromatography or other column-based chromatography. In general, the manufac-
ture of mRNA therapeutics is a very new area, with a very different set of reagents, kinet-
ics, products, and processes.
At the time that the SARS-CoV-2 virus appeared, Moderna had 10 other mRNA-
based drugs approved by the FDA as Investigational New Drugs or for clinical trials, and
had already launched a manufacturing facility for mRNA vaccines that were in various
phases of clinical trial. Because the manufacturing machinery was already in place, the
company was able to take full advantage of the highly versatile and modular nature of an
mRNA vaccine, along with a significant funding stimulus from the U.S. government.
Once the SARS-CoV-2 sequence had been revealed publicly, the U.S. National Institute
of Allergy and Infectious Diseases collaborated with Moderna to determine a most prob-
able target sequence from the spike protein, and Moderna generated DNA templates and
launched production of a new mRNA encoding for the COVID-19 antigen within days,
and using its platform, it took just 3 weeks to go from sequence to vaccine. By contrast,
the turnaround time for a standard vaccine—typically derived from a known protein anti-
gen or a deactivated form of the virus—can range from several months to years. The speed
at which the mRNA vaccine was developed, tested in laboratory animals, and produced
at levels sufficient for clinical trial was record breaking (Box 5-2).
Traditional vaccines use attenuated virus or virus capsids, but a great deal of work
is currently being done on subunit vaccines—vaccines that rely on the use of a protein
antigen that presents a key and broadly recognized epitope of the original infectious agent
that is able to help initiate an immune response. The many advantages of this method
include the ability to generate and mass produce a well-defined protein or polysaccharide
biomolecule product instead of dealing with isolated viruses, typically using classic bio-
reactor synthesis; increased safety of the resulting vaccine for patients, including those
who are immunocompromised; and the increased physical and thermal stability of the
vaccine. Critical to the success of such vaccines is the ability to identify key sequences
on the original infectious agent that can elicit the formation of neutralizing antibodies by
B cells and/or initiate a T cell–mediated response. The nature of this response and its
efficacy are highly dependent on the selection of the correct antigenic sequence or struc-
ture; ideally, this selected region would be a part of the virus that does not undergo many
genetic mutations, so as to ensure efficacy against disease even in the presence of variants.
Furthermore, subunit vaccines generally are not sufficiently immunogenic on their own

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

128 New Directions for Chemical Engineering

BOX 5-2
mRNA Vaccine Development

The discovery and development of disease treatments have been multidisciplinary enter-
prises for many years. One of the most important recent examples of the integration of engineer-
ing and biology to address medical needs is the rapid development of a COVID-19 vaccine. In
less than a year, vaccines were developed, tested in preclinical animal and clinical human trials,
and produced at large scale. This box summarizes a discussion with Moderna’s chairman of the
board, Noubar Afeyan, that addressed the vaccine technology; the role of chemical engineers in
advancing that technology; and the nature of drug discovery, formulation, and expedited bi-
omanufacturing.
Why mRNA? RNA is an information molecule that is modular and specific. In typical bio-
logics, each protein drug is a unique molecule that requires significant adaptation of formulation.
In contrast, messenger RNA (mRNA) is a coded molecule—when the drug is changed, only the
RNA sequence is changed, while the physical and chemical characteristics of the drug molecule
remain the same. Thus, mRNA is especially ideal for vaccines because the immune target is also
a code, and the goal is to replicate the antigenic molecular sequence—thus mRNA is a perfect
fit for vaccines. Additionally, because it is a platform technology, when the drug needs to be
adjusted or changed, the manufacturing process remains the same.
The role of chemical engineers. The idea to use mRNA as a vaccine molecule grew in part
from conversations between two chemical engineers—Bob Langer and Noubar Afeyan—who
launched Moderna to explore the possibility of delivering mRNA to enable cells within the body
to produce the protein antigens for vaccines. Chemical engineers are able to take scientific ad-
vances, such as the successful protecting groups that stabilize mRNA in the bloodstream, and
think more broadly about how to generate a therapeutic. Systems-level thinking aided in con-
sidering the factors needed to get from the molecular and cellular levels to the efficient delivery
and distribution of RNA macromolecules within the body.
Vaccine development. A key aspect of developing a viable mRNA vaccine is the ability to
systemically deliver mRNA, a highly negatively charged macromolecule, to cells that will gen-
erate proteins that are actively accessed by immune cells. The generation of an appropriate for-
mulation using cationic lipid nanoparticles was a problem suited for chemical engineers, and
finding formulations that were effective in enabling both mRNA delivery to the cell cytoplasm
and simultaneous upregulation of immune cells as an adjuvant was key to the formation of a
successful, effective, and commercially viable vaccine.
Frontiers in immune engineering. Chemical engineers can contribute to advances in such
frontiers as guide RNAs that can alter the expression of multiple genes at once, drugs generated
within red blood cells, and the engineering of biological systems. These kinds of developments
will ultimately lead to fewer small-molecule drugs—the body operates on proteins derived from
RNA, and relatively few of a cell’s most important regulating molecules are small molecules.
Ultimately, for many biomedical/pharmaceutical drugs, bioreactors may become irrelevant. It is
likely that in 50 years, all protein drugs will be mRNA drugs.
Clinical relevance of emerging technologies. CRISPR (clustered regularly interspaced short
palindromic repeats), synthetic biology, chimeric antigen receptor (CAR) T, and other cell ther-
apies are the future of medicine. The prospect of manufacturing these types of therapeutics is
challenging, but many similar challenges have proven to be solvable. There is no mystery here—
good engineers can solve even highly complex problems.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Engineering Targeted and Accessible Medicine 129

to induce a strong humoral or T cell response; therefore, for subunit vaccines that utilize
much smaller portions of the protein antigen, adjuvants need to be added to enhance the
body’s innate immune response and upregulate immune-cell activation. The future of vac-
cine development will include the use of genetic data mining and machine learning with
biomolecular computation to enable a predictive approach to vaccine subunit selection,
the ability to rapidly examine the immunogenic response of multiple protein subunits, and
the establishment of protein scaffolds that facilitate such rapid assay approaches.
Chemical engineers will play a significant role in the development of the protein
engineering and computational and biomolecular design space for the discovery of new
antigens. The design of antigens, including the selection of specific sequences, and the
use of different kinds of molecular adjuvants can affect the immunological response by
directing different cellular pathways. For example, for certain viral infections, such as
HIV, it is desirable to elicit a strong intracellular T cell response to prevent further viral
replication; in other cases, a humoral response with highly effective neutralizing antibod-
ies might be sought. Recent efforts have been focused on tissue-resident T cells and other
memory T cells that can play critical roles in advancing immunity. The selection of the
correct epitope to elicit a desired response and the engineering of the means of delivery
of a vaccine to maximize that response introduce additional complexities that chemical
engineering is well equipped to address. Aspects of vaccine delivery include the mode of
delivery, control and impact of pharmacokinetics of antigen exposure, route of delivery,
and degree and nature of uptake and presentation of antigens by antigen-presenting cells.
Such materials as degradable polymers or other forms of delivery devices, such as mi-
croneedles for transdermal delivery, may be used to help control these factors or influence
which cells are accessed upon release of the vaccine (Caudill et al., 2018; Prausnitz,
2017).
As important as the generation of antigenic molecules is, it is equally important
to design their solubility and biocompatibility for ease of manufacture and downstream
generation of purified vaccines. Because subunit vaccines are smaller biomolecular units,
they also provide a more accessible route to creating manufacturing platforms that allow
rapid vaccine development and scalable on-site manufacturing. An interesting biological
engineering consideration is the selection of biological hosts for the manufacturing pro-
cess. As an example, replacing mammalian CHO cells with rapidly dividing yeast cells
for the production of new subunit vaccines can cut by a factor of four the time for both
generation of clinical test compounds and, ultimately, manufacturing-scale generation
(Brady and Love, 2021). The ability to use stable, easily transfected, and rapidly dividing
cell lines can also enable the manufacture of key vaccines at low cost and high yield in
underserved parts of the world (Matthews et al., 2017). These kinds of innovations are
indicators of the potential impact of chemical engineers in enabling rapid response to dis-
eases and a much more accessible and equitable approach to human health.

Engineered Proteins as Therapeutics

Engineered antibodies are now perhaps the largest and most expansive area of
protein engineering. Since the first monoclonal antibody therapies were introduced in the

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

130 New Directions for Chemical Engineering

1980s, the domain of antibody therapies has exploded; by the end of December 2019,
nearly 80 antibody-based therapies were on the market, and eight of the top ten drugs on
the market were biologics (Lu et al., 2020). The design, development, and manufacturing
of antibodies have been critical in the development of treatments of several disorders.
Antibodies work via a highly specific binding and blocking mechanism that allows them
to be highly effective inhibitors of biomolecules and specific cell-membrane receptors,
and to bind to antigens to fight viral and other infectious agents.
Several challenges remain in the generation of antibodies on a commercial scale.
Chief among them is the time and scale needed for antibody production. The cellular ma-
chinery needed to make complex antibody molecules is available in mammalian cells such
as CHO; however, these kinds of cells typically exhibit low yields and slow rates of
growth, and they can require significant maintenance because of their sensitivity to tem-
perature and other environmental factors. Yeast and bacteria cells are much more produc-
tive, have rapid doubling times, and are readily manipulated to generate different protein
sequences. Unfortunately, however, these systems need to produce proteins intracellu-
larly, which requires cell lysis and complex separation steps. The design of the molecules
can impact bioavailability, solubility during manufacture, and blood stability. Chemical
engineers will play a key role in determining the best routes for antibody production.
The molecular design of antibodies is modular and provides a significant design
space; the Fc (fragment crystallizable) determines the cell effector function—the specific
cell type targeted and the nature of cellular impact. Recently, antibody fragments and
engineered fusion proteins have also entered the market, utilizing variants or portions of
the original antibody while introducing a range of advantages over full antibodies for a
number of applications. Each of the Fab (fragment antigen-binding) components contains
an inner binding region, and the single chain fragment (scFv) contains the recognition and
binding sequence from the Fab. These forms allow for more direct incorporation of a
much smaller binding molecule, thus providing molecules that do not require the complex
manufacturing steps needed to produce whole antibodies. Protein engineering of these
molecules has led to exciting therapeutic opportunities—the combination of Fab compo-
nents from two different antibodies can yield bispecific antibodies with dual-binding ca-
pability, and it is also possible to fuse an antibody with a targeting protein to create a
therapy that binds to specific cell types for targeted therapies. Furthermore, additional
modifications can be made in the degree or type of glycosylation, the presentation of
charge and charge heterogeneity, and control of the hydrophobic/hydrophilic nature of the
molecule. These options represent myriad possibilities for therapies that have not yet been
explored. Chemical engineers can manipulate these systems on the molecular level using
rapid assay methods and computation in combination with studies of manufacturing prop-
erties to achieve high yields and lower overall cost.
Biomolecular engineering has had an enormous impact on therapeutic designs. It
has come in a variety of modes, including computational peptide designs; display libraries
of phage, yeast, and bacteria for target discovery; and protein engineering of antibodies
and antibody fragments. The design of phage, bacteria, and yeast libraries has allowed the
generation of numerous diverse peptide motifs to identify target binding agents (Arnold,
2018; Wittrup, 2001). Identification of specific ligands has been a key driver of biologics

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Engineering Targeted and Accessible Medicine 131

discovery. Chemical engineers have made prominent contributions to this field through
the introduction of quantitative analyses and optimization. These display tools have been
extended to in vivo application, enabling direct in vivo identification of target-specific
binding sequences for organs, which can enable targeted drug delivery—the so-called
vascular zip code approach (Teesalu et al., 2012). Viral libraries have also been used to
identify target-specific gene therapies. For example, adeno-associated virus libraries can
be evolved to facilitate in vitro and in vivo gene delivery (Bartel et al., 2012). Such di-
rected evolution approaches have contributed to efficient optimization of agents in the
rather complex biological landscape.
Keeping proteins stable under physiological conditions for prolonged times is
also a significant hurdle. Engineering of agents that promote protein stability continues to
be an area of opportunity in chemical engineering. Biologics now represent a major class
of therapeutics, monoclonal antibodies have emerged as the dominant therapeutic modal-
ity, and nucleic acid–based therapeutics are poised to increase their role in the therapeutic
landscape. However, these delicate biologics require stringent storage conditions. Mainte-
nance of a “cold chain” for frozen formulations represents a serious challenge in the future
use of biologics, especially in resource-limited communities and geographic regions.
Strategies for biologics stabilization have their foundation in protein biophysics, an area
that has benefited from the active participation of chemical engineers. Fundamental stud-
ies founded in the thermodynamics of water structure, influenced heavily by chemical
engineers, have laid the foundation for strategies for protein stabilization. Solutions for
future challenges posed by the need for nucleic acid stability can potentially leverage the
knowledge generated about protein stability.

MODELING AND UNDERSTANDING THE MICROBIOME

The complexity of the microbiome—which includes multiple cell types involved


in both the microbial community of a given organ and the different host cells that support
and respond to that community—presents a challenge for modeling of host–microbiome
interactions to develop predictive solutions for human health (Box 5-3). The past 20 years
have seen several advances in the toolsets available to chemical and biomolecular engi-
neers for creating predictive models for cellular systems; these capabilities are enhanced
when combined with access to large banks of genomic data. Genetic information about a
microorganism, along with the location of key genes, can be determined, and this infor-
mation, in combination with understanding of the biochemical and metabolic activities of
the cell (effectively the reactome), makes it possible to determine the fully reconstructed
signaling networks of cells. This information can then be converted to a mathematical
representation and analysis so that computational approaches can be applied to enable
predictive capabilities of specific network signaling pathways.
These genome-scale metabolic models (GEMs) have progressed and expanded in
large part as a result of the constraints-based reconstruction and analysis (COBRA) ap-
proach (O’Brien et al., 2015). This approach narrows large numbers of possible pathways
in optimizations by applying key constraints, such as nutrient availability and rate of up-
take by cells. Experimental data can be used to enable cell-type and conditions-based

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

132 New Directions for Chemical Engineering

constraints that enable more specific models. More recent advances have made it possible
for computational approaches to take into account dynamic genomics due to evolution, as
well as changes in cell states. In the previous decade, the expression matrix (E-matrix)
was introduced for computationally predicting the expression of genes in an organism
through knowledge of its transcriptional and translational activity. Metabolic networks
and E-matrices can now be used together for modeling of cellular membrane composition
and thus of the impact of such environmental factors as reactive oxygen and hypoxic pH
shifts (Chowdhury and Fong, 2020).

BOX 5-3
The Human Microbiome and its Impact on Human Health

One of the areas of biomedicine that remains a relative frontier is the development of treat-
ments for human health based on the microbiome—the collection of naturally occurring mi-
crobes that reside in coexistence with the body, facilitating or complementing the biological
function of the human host. The most notable regions of localization of such microbiota include
the gastrointestinal (GI) tract, cervix, lung, ovary, skin, and oral mucosa, along with the mucosa
of other organs. Biologists have found a remarkable connection between the different microbes
present in the body and such factors as GI health and function, neurological processes, and even
the immune system. These commensal bacteria live and grow in the microenvironments created
within the body, utilizing nutrients generated by the body. They emit biomolecules that engage
signaling networks to modulate the behavior of other surrounding microbes and the human host.
Work over the past several decades has led to a better understanding of the role played by
these native bacteria in maintaining and enabling human health. Chemical engineers can make
a vital contribution toward understanding the complex signaling and resultant biological re-
sponses of the body, and in applying this knowledge toward human health treatments. Systems
biology approaches and the use of genomics can be deployed by engineers to describe more
completely the molecular pathways and genomic networks involved in regulation of the micro-
biome and human host. Ultimately, this kind of understanding will uncover the role of bacterial
communities in affecting health and indicate ways to translate that knowledge into therapies.
These treatments vary, and there are many possibilities in the design of therapies that transfer
or transplant healthy human microbiota to correct the existing bacterial systems and help restore
the original microbiota function. Furthermore, synthetic biology tools have made it possible to
directly modify bacteria to create new therapeutic approaches for addressing a range of chronic
disorders or enabling settings with greater resistance to infections or susceptibility to disease.

The advancement of these methods in metabolic engineering in general has


opened up many opportunities to apply them to understanding the microbiome. With the
toolsets provided by GEMs, it is possible to connect microbiome and host-cell interactions
with final biological function (Chowdhury and Fong, 2020). These efforts are greatly fa-
cilitated by significant cross-institutional efforts, including the National Institutes of
Health’s (NIH’s) Human Microbiome Project (HMP), a massive NIH Common Fund ef-
fort involving public and private institutions that led to identification and sequencing of
the microorganisms in multiple human organs and the generation of tools for connecting

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Engineering Targeted and Accessible Medicine 133

microorganism genomic data to disease state and function. A second phase of HMP—
HMP2—led to a series of publications on the role of the gut microbiome in inflammatory
bowel disease (IBD), preterm births, and diabetes (Nature, 2019). The gut microbiome is
the most studied microbiome of the human body, and significant progress has been made
toward building databases that will inform GEM models, as well as other computational
approaches, particularly machine learning and artificial intelligence methods that require
large datasets. An example of a comprehensive GEM model is AGORA (assembly of gut
organisms through reconstruction and analysis; Magnusdottir et al., 2017), an approach
that uses gut metabolic and genomic data to semiautomatically generate reconstructions
of metabolic pathways for 773 independent gut microorganisms. This accomplishment is
of great significance in illustrating the power of such models to assist in determining the
functional behavior of myriad combinations of cells in consortia, thus enabling such ca-
pabilities as determination of predictors of disease identification of targets for treating
disorders. In 2020, AGORA2 was introduced (Heinken et al., 2020). The number of mi-
crobes included increased 10-fold; additional factors, such as microbial drug degradation,
were introduced; and both singular and pairwise bacteria members are considered. These
kinds of models can lead to better understanding of disease (e.g., understanding the role
of microbial consortia as opposed to individual bacteria in regulating key processes) and
exhibit predictive capabilities (e.g., predicting patient susceptibility to IBD or Crohn’s
disease based on nutritive deficiencies).

Cell Transfer Approaches to Medical Applications for the Microbiome

Increased understanding that the gut microbiome of healthy humans can be ex-
tremely beneficial for the management of infection, inflammation, and metabolism has
led to the use of native human microbial consortia—the collections of different naturally
occurring bacteria species that reside within the body—as a means of regulating disease
(Colman and Rubin, 2014). The approach uses fecal matter transplants (FMTs) to transfer
the healthy and established microbiome from donor to patient. The efficacy of this and
related approaches, such as the use of oral pills formed from purified fecal matter, has
been demonstrated to show efficacy in restoring the ability to fight gut-related infections
(e.g., aggressive forms of colitis, severe sepsis, diarrhea) and are in clinical trials for many
related conditions (Wischmeyer et al., 2016)—more than 900 clinical trials for these ther-
apies were listed in clinicaltrials.gov in 2020 (Wang et al., 2021b). Chemical engineers
can help advance these kinds of native microbiota-based therapies to a new level by ad-
dressing their expansion, cost, and scalability—for example, by investigating isolation
and purification of native microbiota to enable more consistent therapeutic products and
approaches that can be readily replicated. Much is to be gained as well from a deeper
understanding of the impacts and effects of such therapies. Machine learning, in combi-
nation with information gained from genomics and systems-biology models, may enable
more effective predictive methods and more personalized approaches to treating patients
and pairing these therapies with given characterized FMT systems.
The use of native microbial consortia, or communities of microbes of different
types and function, demonstrates the power of the synergism that exists among the many

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

134 New Directions for Chemical Engineering

different cells within the human microbiome. Cellular communities consist of independ-
ent cells that work together to share the labor of generating different metabolites, main-
taining different roles in managing energy, producing molecules, and enabling breakdown
or processing of different nutrients. The various bacterial components create supportive
environments that enable the fully functional microbiome to provide dramatic gains in
immunological and anti-infective properties. Together, these communities are more re-
sistant to disease, and exhibit reduced metabolic burden and a broader range of function
than individual microbe types. Therefore, the next step in microbiome therapeutic ap-
proaches is the engineering of microbes and their combination to form similarly syner-
gized consortia with potentially greatly enhanced capabilities (McCarty and Ledesma-
Amaro, 2019). Given the complexity of the interacting cellular systems and their various
functions, a systems approach will be required to replicate important interdependencies
while incorporating new or enhanced functionalities.

Synthetic Biology as a Tool for Engineering Microbes


to Improve Microbiome Health

Bacteria have unique qualities that can be manipulated to enable the generation
of synthetic microbiota with programmed function. In conjunction with gene editing tools,
this work can yield a wide range of functions that depend on communication and interac-
tion with other cell types. These functions include intercellular signaling, among them
chemical, biophysical, and adhesion molecule–based interactions between cells, as well
as ionic and electrical channels. Quorum-sensing capabilities in bacteria populations,
based on the sensitivity of cells to an autoinducer molecule that increases concentration
as cell density increases, are an example of a microbial behavior system that can be ma-
nipulated to advantage in synthetic communities. It is possible to engineer quorum-sens-
ing mechanisms that are independent and thus completely orthogonal to native commen-
sal bacteria, thus enabling unique or independent “programs” instituted on the basis of
specific cell types that will be responsive to a given quorum-sensing signal. Other mech-
anisms of cell–cell communications include N-acyl-L-homoserine lactones (NHL) and
autoinducing peptides that can cause bacterial-cell circuits, programmed into given mi-
crobial species, to turn on or off based on their concentrations (Hwang et al., 2018). These
tools can be combined with additional means of triggering response; for example, bacteria
can be designed to respond to exogenous signals. As another example, inducer molecules
can be used to trigger the expression of transgenes that activate a promoter. Synthetic
biology enables the design of independent gene regulatory systems, and this orthogonality
makes it possible to generate a range of independent controls that can be used to manage
complex systems with multiple organisms.
Synthetic biology tools such as CRISPR and the assembly of DNA parts or “cir-
cuits” that can control genetic expression on demand can be used to modify a microor-
ganism by introducing the ability to generate a given molecular signal, gene regulator, or
indicator in the presence of a given disease condition. Examples of singular commensal
or symbiotic bacteria that have been engineered include

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Engineering Targeted and Accessible Medicine 135

 E. coli that can generate glycotransferases that yield molecular inhibitors to


dangerous enterotoxins that cause diarrhea,
 lactobacillus engineered to prevent infection through the generation of bacte-
ria-killing lactoferrins,
 cholera-controlling microbes that express a regulator of virulence genes, and
 E. coli microbes engineered to secrete antimicrobial peptides that can reduce
the effects of Salmonella (Tan et al., 2020b).

As shown in Figure 5-6, harmless commensal bacteria can be modified syntheti-


cally and then reintroduced to the microbiome in the gut using oral delivery or transplant
approaches. These newly introduced bacteria are then able to incorporate into the native
microbiota and generate molecules that can address infection or greatly lower the impact
of toxins released by undesired bacterial infections.
Along with modifying susceptibilities to infection, microbes can be engineered to
release regulators of the immune system. This ability is particularly relevant for the treat-
ment of inflammatory disorders such as IBD; Lactis bacteria (a probiotic used in produc-
tion of cheese and yogurt) can be modified to produce anti-inflammatory cytokines such
as IL-10 to treat IBD effectively. This and related approaches have generated great interest
in the use of engineered microbes for therapeutic applications for conditions ranging from
infection and inflammation to metabolic disorders and potentially even endocrine and au-
toimmune disorders, such as diabetes and obesity (Tan et al., 2020b). Given recent find-
ings regarding the gut microbiome–neurological connection and the likely advances in
biological understanding of both the gut and other human microbiome communities with
respect to disease and overall human health, this area of technology is likely to expand
significantly in the next decade.

FIGURE 5-6 Engineered commensal bacteria as living therapy in the gut. Commensal bacteria
can be engineered to (a) inhibit pathogenic bacteria or toxins or (b) reduce inflammation through
secretion of anti-inflammatory proteins. SOURCE: Tan et al. (2020b).

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

136 New Directions for Chemical Engineering

Activating molecules can include toxins, antibiotics, or specifically determined


inducer molecules. Syntrophic dependencies, in which a molecule produced by one mi-
croorganism synergizes the activity or metabolism of another, further increase the com-
plexity of these systems and enable the kinds of metabolic interdependencies found in
nature (Harimoto and Danino, 2019; Hwang et al., 2018). It is clear that using these kinds
of engineering tools to better design therapies based on engineered consortia will require
applying understanding of complex signaling behaviors, modeling, and the manipulation
of synthetic biology tools.
Chemical engineering, which has had a traditional focus on metabolic engineer-
ing, is a discipline particularly well poised to enable advances in these areas. In the late
20th century, the field of biochemical engineering brought forth the concept of the cell as
a reactor and the use of cellular pathways to create molecules of interest with a high and
controlled yield. Early developments in this field led to work on metabolic manipulation
of cells to direct and control certain cellular reaction pathways over others. The microbi-
ome is in some sense a highly tuned and sensitive bioreactor or series of bioreactors within
the human body. The challenges for chemical engineering lie in designing cells for ma-
nipulating not only the generation of desired compounds but also the rate of generation,
the selectivity for a given target, and the longer-term stability of the reacting system.
Chemical engineers have embraced synthetic biology as a critical tool in addressing such
problems and are well equipped to advance this work.
The unique skillsets of chemical engineers can contribute to gaining knowledge,
enabling discovery, addressing disease, and enhancing human health through understand-
ing and regulation of the human microbiome. Key challenges and opportunities include
further advancement of synthetic biology tools that incorporate environmental and condi-
tional responses, regulation of the reactome across multiple species, and engineering of
cellular consortia to achieve patient-specific outcomes. The expansion of engineered ther-
apies to other microbiomes will result in the need to determine effective modes of delivery
to specific organs, which may require the engineering of new biomaterials that support
the transfer of microbiota to different parts of the body. Along with advances in GEM
methodologies, additional data science and computational approaches will likely become
more important in predicting and regulating metabolism. Increased availability of data on
the gut, skin, vaginal, oral, and other organ microbiomes will enable the use of artificial
intelligence and machine learning algorithms, with anticipated increased use of models
and computation to direct the early detection of potential disease and the application of
preventive health measures to avoid disease.

DESIGN OF MATERIALS, DEVICES, AND DELIVERY MECHANISMS

Devices play an important role in the landscape of biomedical technologies. De-


vices that control drug administration are of special interest to chemical engineers and an
area in which they have made significant strides. Such devices include external infusion
pumps, implantable pumps, self-regulated delivery devices, and transdermal patches,
among others (Anselmo and Mitragotri, 2014; Vargason et al., 2021). Implantable pumps

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Engineering Targeted and Accessible Medicine 137

offer a patient-compliant means for long-term disease management and have had a trans-
formative impact on diabetes, pain management, and neurological disorders.
Patients with type 1 diabetes are on lifelong treatment with insulin, and the need
for frequent insulin injections represents a significant limitation in diabetes management.
Implantable pumps offer an excellent alternative, especially for the delivery of basal in-
sulin (Cescon et al., 2021; Shi et al., 2019; Wolkowicz et al., 2021). Several challenges,
including those related to insulin stability and device biocompatibility, had to be over-
come for these devices to reach the clinic. Another major challenge has been managing
the compatibility of implanted pumps with the human body (Kleiner et al., 2014; Park and
Park, 1996). Foreign devices, including implanted pumps, evoke a foreign-body response
that includes the arrival of immune cells, leading ultimately to the formation of fibrous
capsules, whose properties play an important role in determining the durability of the de-
vice. Chemical engineers have made important contributions to this field through under-
standing of the foreign-body response and the development of coatings to minimize fi-
brous capsule formation.
Control strategies are especially relevant for insulin pumps, for which an active
control of insulin release is necessary to maintain euglycemia. A healthy pancreas adopts
a complex control algorithm to control insulin release such that glucose levels remain
within a tight window. Accordingly, chemical engineers have led extensive efforts to de-
velop model-based and adaptive control strategies that enable communication between
glucose sensors and insulin pumps. As glucose sensors have advanced to the point of
providing continuous measurement over several days in patients, effective control algo-
rithms have supported the development of closed-loop insulin delivery devices—the so-
called artificial pancreas (Doyle et al., 2014). The paradigm of self-regulated pumps can
be extended to other indications for which continuous and regulated delivery is essential.
The development of such devices requires novel formulation strategies, pump designs,
control algorithms, and continuous sensors, challenges for which chemical engineers are
well suited.
Another exciting frontier for chemical engineering is the development of devices
for completely noninvasive methods of drug delivery. Such devices can be placed on the
skin, where they can offer a needle-free method of continuous and regulated drug admin-
istration. These transdermal patches provide a natural means for the sustained release of
drugs (Prausnitz and Langer, 2008), providing the key benefit of active patient control
and ease of termination when needed. Several transdermal patches are currently available,
allowing easy delivery of nicotine and fentanyl, for example. However, these patches are
limited to the delivery of small-molecule drugs because of the transport limitations of
human skin, leaving many emerging drugs beyond their scope.
Several advances have been made in improving drug delivery across the skin to
expand the use of transdermal patches to biologics. One such advance, and an area in
which chemical engineers have already made an impact (Hao et al., 2017; Prausnitz,
2017), is microneedles, which penetrate the skin to a depth sufficient for delivering drugs
but not for inducing pain. Microneedles are showing high potential, particularly for deliv-
ering vaccines, skin being an excellent organ for vaccine delivery because of the presence
of immune cells. Further, microneedles carry vaccines in solid form, thus improving their

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

138 New Directions for Chemical Engineering

storage stability. Such solid, stable, self-administered vaccines could be particularly suit-
able for use during the current COVID-19 pandemic. Strategies have also been developed
for enhancing noninvasive delivery of biologics through oral or inhalation routes, with
enabling contributions from chemical engineers (Brown et al., 2020; Matthews et al.,
2020; Morishita and Peppas, 2006).
The human body is a connected system of various organs, each designed for a
unique function. While the macroscopic behavior of each organ has historically been stud-
ied in medicine, quantitative understanding and control of transport properties in these
systems are limited. In the absence of this control, access to many organs is severely lim-
ited. At a fundamental level, transport is limited by the body’s natural metabolic processes
and transport barriers. These biological barriers, while serving the important purpose of
regulating the body’s metabolic functions, also limit how drugs can be delivered in the
body (Figure 5-7). Accordingly, many potential drugs fail to reach their destination in the
body, and thus most molecules never become drugs. This limitation ultimately reflects the
high cost and lengthy timeframes of drug development.

FIGURE 5-7 Examples of biomaterials and their routes of administration for in vivo use. In addi-
tion to pills and injections, biomaterials have been developed to administer drugs successfully in a
variety of other ways. SOURCE: Fenton et al. (2018).

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Engineering Targeted and Accessible Medicine 139

The early contributions of chemical engineers to small-molecule drug delivery


were focused on developing encapsulation strategies to control the release of drugs. Mac-
roscopic and microscopic depots have been designed to encapsulate small and large drugs.
The most commonly used microsphere-based drug delivery systems consist of a suspen-
sion of polymer microspheres that can be delivered by subcutaneous injections. Funda-
mental studies describing the diffusion and degradation kinetics and mechanisms of dif-
fusion in the polymer matrix, led by chemical engineers, have played an important role in
the establishment of these systems (Ritger and Peppas, 1987). Triggered release of drugs
from these polymeric depots, mediated by such external triggers as light, ultrasound, and
magnetic fields, has allowed better control over the drug release (Sun et al., 2020). Chem-
ical engineers have numerous opportunities to advance sustained-release depots, espe-
cially with respect to extending the release duration, which can further improve patient
compliance. Such advances are especially relevant for such applications as ocular deliv-
ery, given the strong motivation to minimize the number of injections in the eye. Long-
term compatibility of depot systems with the immune system also represents a continued
challenge.
At a nanoscale level, the same polymeric systems have provided benefit for tar-
geted drug delivery, a concept first introduced for the delivery of chemotherapeutic
agents. Numerous efforts have been made to improve the targeting ability of nanoparticles
by virtue of their surface modification of polyethylene glycol (PEG), which reduces sur-
face protein adsorption and subsequent hepatic sequestration. Nanoparticles can be further
modified by targeting ligands, including peptides, aptamers, and antibodies, to improve
target accumulation (biomaterials are discussed in more detail in Chapter 7). A number
of nanoparticles, including liposomes, polymeric nanoparticles, and inorganic nanoparti-
cles, are already commercially available, and several are available in the clinic (Karabasz
et al., 2020). Lipid nanoparticles have played a central role in packaging of mRNA to
enable its stabilization and intracellular delivery (Mitchell et al., 2021). While a large
number of materials, targeting moieties, and design strategies have been explored, achiev-
ing exquisite targeting remains an unmet need.
Moving forward, opportunities for chemical engineers in the field of drug deliv-
ery include developing a better understanding of transport processes and leveraging this
understanding to accomplish better targeting. One of the key challenges is the limited
spatial and temporal resolution of drug imaging in the body. Model systems (e.g., organs-
on-chips) can provide a deeper understanding of such transport processes (Bhatia and
Ingber, 2014; Ghaemmaghami et al., 2012). Drug delivery methods also need to be pa-
tient-centric; that is, they need to improve patient compliance, including such considera-
tions as cost, simplicity, and convenience.

Building Sensors and Diagnostic Tools

Rapid, frequent, and inexpensive diagnosis is the foundation of successful health


care. Blood-based diagnostic methods, including those in the routine medical checkup and

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

140 New Directions for Chemical Engineering

in focused disease diagnosis, have been the cornerstone of past and even current diagnos-
tic infrastructure. With perhaps the notable exception of glucose monitoring, blood-based
diagnosis is typically done in the hospital setting, requiring clinical supervision.
Extensive efforts have recently been made to shift away from this inconvenient,
discrete operation of blood sampling to a new paradigm focused on continuous measure-
ments of physiological markers. Most such advances have been made in glucose monitor-
ing for diabetes (Lipani et al., 2018). Continuous, wearable devices now available for
patient use measure glucose for days through a single wearable device. In addition to
providing multiple measurements, continuous sensors offer an educational tool enabling
patients to understand how their glucose levels respond to various metabolic and external
factors (e.g., Brown et al., 2019). Efforts have also been made to measure glucose con-
centrations in a completely noninvasive manner, making use of spectroscopic methods,
noninvasive collection of tissue fluid through the skin, or the use of sweat-based sensors.
The latter devices measure various analytes present in the sweat and use these measure-
ments to derive a physiological assessment (Chung et al., 2019). A variety of analytes are
present in sweat, including small molecules, such as glucose and hormones, and large
molecules, such as proteins. Improvements to sensitivity, accuracy, and modeling based
on these measurements are needed if these devices are to achieve broad health care appli-
cation.
In addition to molecule-based diagnosis, wearable sensors can perform a variety
of other diagnostic measurements, including measurements of temperature, blood oxy-
genation, pulse, and even biopsy alternatives (Gao et al., 2019; Kim et al., 2019). The
ability to derive these key measurements through a wearable sensor has already trans-
formed the collection of human physiology data. Traditionally, these measurements were
possible only using large, bulky sensors, often connected to stationary electronic proces-
sors and displays. However, the ability to perform these measurements in a truly nonintru-
sive manner has enabled the collection of information about human responses in real-life
situations.
Extrapolating these trends into the future, wearable sensors are expected to collect
massive amounts of information about human behavior in healthy and diseased condi-
tions. Some of the technical challenges in this field include the development of sensors to
accurately collect and measure the small amounts of analytes that are available in sweat.
Another challenge lies in accurately correlating these measurements with blood concen-
trations, a challenge that pertains specifically to the transport of analytes from the blood
into the interstitial fluid, and relating this information to individual patient health. Can
models be developed using these data to predict catastrophic physiological events through
such analysis? Can drug reactions be better understood or predicted through analysis of
such massive amounts of data? Opportunities exist for chemical engineers to develop
large-scale network models with which to understand the dynamics and connectivity of
adverse events. By virtue of their training in understanding and appreciating multiscale
dynamics, chemical engineers are especially well suited to undertaking this challenge.
Personalization of drug therapies is an exciting opportunity to reduce adverse events with-
out compromising therapeutic efficacy. However, data need to be available to support

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Engineering Targeted and Accessible Medicine 141

such personalization at the design stage and reduction of adverse events at the follow-up
stage.

Cell-, Organ-, Organism-on-a-Chip to Model Biological Functions

Drug development is typically a lengthy and expensive process, with 10–15 years
and approximately $2.6 billion dollars typically being required for development of a new
drug (DiMasi et al., 2016). A large fraction of this time and expense is associated with
late preclinical and clinical development. Acceleration of lead identification enabled by
high-throughput screening, rational drug design, or computational platforms has greatly
enhanced the availability of early lead candidates. However, the transition of these leads
to preclinical and clinical programs is limited by challenges associated with the unknowns
involved in evaluating the interactions of the leads in the body. In fact, the overall proba-
bility that a drug entering clinical trials will be approved by the FDA is about 12 percent
(DiMasi et al., 2016). Currently, despite advances in computational systems biology and
in vitro models, more than 60 percent of these failures are due to lack of efficacy and
another 30 percent to toxicity (Waring et al., 2015).
Another key limitation is the limited relevance of animal to human pathophysiol-
ogy. Small animals, especially mice, are commonly used in preclinical studies, their use
being driven largely by the existence of disease models in these rodent systems. Specifi-
cally, several genetically engineered murine models exist with which to assess some of
the key biological aspects of a drug, including phenotypic disease manifestation, target
specificity, and improved survival. However, translation of these benefits to the engineer-
ing aspects of drug development is severely limited. For example, some of the major bi-
ological barriers to drug transport in the body, such as skin and mucous membranes, are
substantially different in mice and humans. Some key aspects of these barriers that dictate
diffusion differ greatly between mice and humans, in some cases primarily because of the
differences in body mass and in others because of the innate differences between these
species. Beyond mice, the universe of large-animal models is highly fragmented, and their
use depends on the disease in question. Ultimately, the nonhuman primate model, which
is often a key preclinical model, is ethically, logistically, and financially challenging. In
addition to limiting the speed of drug development, these hurdles bias the landscape of
drug developers because the required resources and time are often a luxury available only
to large companies. Tools to minimize the burden of preclinical drug development will
not only reduce the cost and time of development and provide preclinical information that
is more relevant to clinical programs, but also level the playing field for drug developers.
Several advances have been and continue to be achieved in the development of
scaled-down microphysiological systems (the so-called organ-on-a-chip or human-body-
on-a-chip) that will provide an output at least as predictive as animal testing (Low et al.,
2021). This field was pioneered by the cosmetic industry with the goal of eliminating
animal testing for its products. That strategic decision led to the development of in vitro
human epidermis models that can provide meaningful information about the safety of
cosmetic products (Faller et al., 2002). The last two decades have seen similar devices
able to mimic the function of other vital organs, including the liver, brain, and lungs,

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

142 New Directions for Chemical Engineering

among others (Low et al., 2021). Such systems require considerations not only of the
biology—that is, incorporation of different cells (e.g., lung epithelial cells, astrocytes,
hepatocytes) in the system—but also of the engineering aspects of flow and interfaces.
For example, the liver is the most well-perfused organ in the body, and its function de-
pends on that. Hence, a liver-on-a-chip would need to account for the intricacies of vas-
cular flow (Schepers et al., 2016). Such devices can play a role in assessing drug toxicity.
Hepatotoxicity is a key safety concern for many drugs, and a means of screening for it at
an earlier stage could substantially accelerate drug development. Liver chips could also
help screen nanomedicines, which are actively cleared by the liver after their administra-
tion to the body. Similarly, a lung-on-a-chip would need to incorporate the key features
of an air–water interface in the presence of lung surfactant and mucus. Such lung chips
could aid in assessing the interactions of drugs with the lung microenvironment, a topic
that has become critically important during the current COVID-19 pandemic (Saygili et
al., 2021).
Systems that capture the key elements of the immune system in vitro are also
expected to support the design of future therapeutic products. The immune system is cen-
tral to many major health concerns, including infections, cancer, and autoimmune dis-
eases. The immune system functions through extensive orchestration of many cell types,
including dendritic cells, T cells, and others, to stimulate a holistic response to drugs and
vaccines. Systems that capture the key events of such orchestration in vitro (e.g., lym-
phatics-on-a-chip) would provide insights into the interaction of therapeutics with the im-
mune system.
Moving forward, such organs-on-a-chip will require active incorporation of bio-
logical and engineering aspects into the design (see Figure 5-8). From the biological per-
spective, incorporating all essential cell types in the system is critical. More is not neces-
sarily better because the model systems need to capture the essential complexity of the
organ in the human body while being as simple as possible. Critical as well will be main-
taining the cells in a correct phenotypic state, and ensuring cellular communication in
these systems is also critical. Such key attributes include the barrier function—for exam-
ple, diffusive properties of tissues and permeation across tight junctions or dynamics aris-
ing from vascular flow. Such systems need to leverage advances in microfabrication to
capture key structural complexities in organs and advances in biological tools, such as
gene editing, to control cells, and they require the means to incorporate and address com-
plexity in a tissue microenvironment.
From an engineering perspective, most nonorganoid-tissue chips have very low
throughput. Thus only a few replicates can be performed at any one time, which limits the
ability to screen thousands of potential hits during drug discovery. More automated and
miniaturized systems will be needed for commercial use. In addition, most systems are
fabricated in-house in academic laboratories, so reliability and reproducibility become
limiting. Clear approaches to quality control are needed, including physiological valida-
tion (sensitivity, specificity, and precision). Particular opportunities of interest to chemi-
cal engineers include understanding the connectivity of various organs-on-chips and as-
sessing the role of flow in cells and organs. Both development and validation of these
systems are research areas that chemical engineers are well suited to address.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Engineering Targeted and Accessible Medicine 143

FIGURE 5-8 Advances in microphysiological systems rely on the development of three main
components: the cell source, chip technology, and biomarker discovery. SOURCE: Ramadan and
Zourob (2020).

HYGIENE AND THE ROLE OF CHEMICAL ENGINEERING

Historically, engineers have played a major role in the development of hygiene


and the sanitary infrastructure, developments that have led to an increase in life expec-
tancy in much of the world. Indeed, diarrhea (endemic cholera) is no longer among the
leading causes of death in world statistics (see Figure 4-8 in Chapter 4). Today in the era
of COVID-19, chemical engineers, especially those collaborating with other scientists and
engineers in environmental sciences and technology, have opportunities to contribute to
societal health and well-being and help narrow the disparities between low- and middle
income countries and higher-income countries.
Recent years have seen great strides in better understanding the link between in-
door air quality and health,1 an area of growing concern in which chemical engineers can
be expected to make major contributions. Within months of the onset of the COVID-19
pandemic, for example, long-lived aerosols (airborne particulates or droplets) were iden-
tified as the primary source of human-to-human transmission of the virus (Edwards et al.,
2021; Prather et al., 2020) and a focal point for mitigation strategies, including the use of
social distancing, face masks, and high-efficiency air filters. Figure 5-9 depicts the mech-
anistic framework for such superspreading events as the choir practice of the Skagit Val-
ley Chorale, showing an aerosol size distribution from Bazant and Bush (2021). From
their fundamental understanding of fluid mechanics, chemical engineers can readily con-
firm the time scale (many hours) for airborne suspension of aerosols in the 1-micron- to
submicron-length scales in the figure, which indeed is the time scale described in the early
COVID-19 literature (Prather et al., 2020).

1
The National Academies’ Committee on Emerging Science on Indoor Chemistry is currently
examining the state of science regarding chemicals in indoor air.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

144 New Directions for Chemical Engineering

FIGURE 5-9 Model predictions for the steady-state, droplet radius–resolved aerosol volume frac-
tion (linear scale—top panel; logarithmic scale—bottom panel) produced by a single infectious
person in a well-mixed room. The model accounts for the effects of ventilation, pathogen deacti-
vation, and droplet settling for several different types of respiration (singing, singing softly, speak-
ing, whispering, mouth breathing, nose breathing) in the absence of face masks. The ambient con-
ditions are taken to be those of the Skagit Valley Chorale superspreading incident. SOURCE:
Bazant and Bush (2021).

Chemical engineering and aerosol experts Edwards and colleagues (2021) took
the analysis of aerosol size distribution in another useful direction—variation within the
population as a function of several factors, including COVID-19 infection, age, and body
mass index. They showed that all three of these factors can influence aerosol sizes by
three orders of magnitude, and thus serve as a useful starting point for estimating of
spreading distances and infection transmission rates. Ultimately, such insights on trans-
mission rates can help guide policies (Samet et al., 2021) and mitigation efforts, such as
the use of masks and air filters.
Hand sanitizers and related consumer/cleanser products are another aspect of hy-
giene that has risen to prominence during the COVID-19 pandemic, and an area in which
chemical engineering has an important role to play, especially with respect to the balance

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Engineering Targeted and Accessible Medicine 145

between antimicrobial/antiviral efficacy and product safety and environmental impact.2


The unprecedented high volume of use of hand sanitizers to combat COVID-19 has raised
concerns regarding potential toxicity and/or negative impact on the environment
(Mahmood et al., 2020) and rekindled interest in the development of new materials and
formulations that can address these concerns.
Historically, the chemical engineer’s understanding of particulate and aerosol fil-
tration has been based on mathematical models for capture by a single fiber (Spielman,
1977), with additional enhancements to build in geometric considerations, such as tortu-
osity and porosity. Such modeling approaches are common to other areas of chemical
engineering that also analyze flow through porous media (e.g., modeling of oil reservoirs
and mass transport in porous catalysts). Recent advances in coating materials, additive
manufacturing, and systems engineering (Bezek et al., 2021; Christopherson et al., 2020)
highlight further opportunities for filtration technologies for face masks and indoor filters.
As attention moves past the current pandemic, and other potential challenges to indoor air
quality and possible perturbants are better understood, chemical engineering will have
growing opportunities to help develop materials that can change color or otherwise indi-
cate health challenges so they can be addressed by physical barriers or treatment.

ENGINEERING SOLUTIONS FOR


ACCESSIBILITY AND EQUITY IN HEALTH CARE

NIH defines health disparities as differences among specific population groups in


the attainment of full health potential that can be measured by differences in incidence,
prevalence, mortality, burden of disease, and other adverse health indicators. While the
term “disparities” is often used or interpreted to reflect differences among racial or ethnic
groups, disparities can exist across many other dimensions as well, such as gender, sexual
orientation, age, disability status, socioeconomic status, and geographic location
(NASEM, 2017). Despite overall improvements in population health over time, many
such disparities have persisted, and in some cases, widened.
There is a persistent lack of awareness of and engagement with health disparities
within science and engineering in both research and educational activities (Vazquez,
2018); indeed, one barrier to the greater involvement of engineers in health disparities
research is their lack of knowledge in this area (McCullough and Williams, 2018). Alt-
hough many health disparities result from systemic issues that can be addressed only
through larger social changes, chemical engineers can still play a role in helping to resolve
these issues. If they are to do so, however, these issues need to be introduced in the class-
room and the workplace, imparting an understanding of the complexities and implications
of health disparities, including the associated public health concerns and the social context
of differential medical treatment based on race, gender, sexual orientation, age, disability

2
One aspect of this balance can be seen on the FDA webpage “FDA updates on hand sanitizers
that consumers should not use,” with the following hazards (including product recalls): contains
methanol or 1-propanol; contains microbial contamination; is subpotent, containing insufficient
levels of ethyl alcohol, isopropyl alcohol, or benzalkonium chloride; is packaged in a container
that resembles a food/beverage container, increasing risk of accidental ingestion.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

146 New Directions for Chemical Engineering

status, socioeconomic status, and geographic location (Barabino, 2021). Addressing these
disparities in chemical engineering education can even help attract students from diverse
backgrounds (the importance of which is discussed in Chapter 9), students who are more
likely to engage in exploring these issues as they go forward in the field (Thoman et al.,
2015).
One opportunity for applying engineering solutions to reducing disparities is in
low-resource settings. With respect to COVID-19, for example, chemical engineers can
play a role in developing affordable vaccines for low-income populations who may not
have access to traditional vaccine supply chains, distribution facilities, or cold-storage
options (e.g., Frueh, 2020). Generally, creating appropriate technologies for low-resource
settings requires that engineers consider not only the scientific rigor and effectiveness of
the technologies but also their adaptability to local needs and ability to be maintained
using resources the community can afford (TARSC, 2015). Today, while information is
easily available to anyone with internet access, scientific tools and health care devices still
prove to be expensive and inaccessible for many communities. As an example of address-
ing this problem, chemical engineers played a role in creating an appropriate and low-cost
technology for diagnosing sickle cell disease in rural sub-Saharan Africa. What followed
was a variety of new, low-cost, point-of-care testing devices that enable individuals to
seek out sickle cell treatment in time (McGann and Hoppe, 2017; Nnodu et al., 2019;
Oluwole et al., 2020).
Ethics, empathy, and attitudes are interrelated and important for eliminating dis-
parities and building pathways to health equity. Conversations surrounding the ethics of
engineering typically emphasize the integrity of the procedural steps of the scientific pro-
cess. When creating and implementing new technologies, chemical engineers need to in-
corporate a broader set of ethical considerations and cultural competencies. Essential eth-
ical considerations include the impact of new technologies or processes on low-resource
communities and marginalized populations who experience greater health disparities, in-
cluding how new treatments or technologies will be received among different cultures and
populations and the impact on the environment. Chemical engineers have an opportunity
to help reduce health disparities when they explicitly incorporate human-centered design
into technologies to make them more accessible, equitable, and culturally sensitive. Col-
laboration across disciplines—engineers joining with clinicians, social scientists, policy
makers, and members of the communities being served—will help address multilevel de-
terminants of disparities and lead to better and more equitable interventions.
Another aspect of increasing accessibility to reduce disparities is lowering the
cost of therapeutic interventions. The U.S. demand for monoclonal antibodies (mAbs) and
a number of other biological compounds continues to grow, in part because of the increas-
ing average age of people in the United States (Figure 5-10) and the diseases associated
with aging, including cancer and cardiovascular and respiratory diseases. The success of
mAbs in addressing those diseases drives demand. Unfortunately, the cost of producing
biologics and the subsequent costs to consumers create pressure to improve flexibility and
reduce costs so as to increase health care equity while maintaining reliability and stability
during manufacturing and distribution.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Engineering Targeted and Accessible Medicine 147

FIGURE 5-10 Projected population share of children under 18 and adults aged 65+ up to 2060.
Adults aged 65+ are expected to outnumber children under 18 by 2034. SOURCE: U.S. Census
Bureau (2018).

Approaches to reducing costs include streamlining workflows and increasing


plant size, as well as developing strategies for producing higher protein yields per unit
volume both in bioreactors and during purification and storage. One of the most obvious
ways to produce high protein yields is to use continuous bioreactors, or perfusion culture.
As in a traditional continuous reactor, this approach entails keeping catalyst (in this case
the cells) within the reactor while adding fresh reactants (media) and removing product
and spent media. Industry has long avoided this approach, as product yields have reached
2–10 g/L in fed-batch culture through media and process-based optimization. However,
increased application of single-use technologies has opened up the possibility of broader
application of perfusion technology, which can lead to three-fold increases in volumetric
productivity (Bausch et al., 2019). In addition, expanding manufacturing space quickly is
much more challenging with traditional plant designs, with new facilities costing more
than $400 million and taking 5 years to build (Jagschies, 2020).
Some avenues toward improving perfusion culture processes include better
scaled-down models to enable early development and screening of cells and optimization
of media, which during traditional industrial development can take 3–6 months. Blending
of concentrated media stocks by in-line dilution is also needed; even four-fold concen-
trates can save time and resources. Needed as well is the ability to combine these blending
and perfusion runs with adaptive control technology driven by models of nutrient con-
sumption, as well as improved cell-line development to ensure that the cells provide ro-
bust product (yield and quality) over the length of time expected for growth (2–3 months).
In addition, supply chain issues, such as quality and consistency of raw material, become
even more important with a perfusion approach to minimize product variance. Regulatory
issues become more significant as well in that the product may change composition with

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

148 New Directions for Chemical Engineering

time (Allison et al., 2015); thus, improved control technology and defined product batches
could lead to better acceptance of this approach.
One of the most important directions for reducing costs is moving beyond tradi-
tional model organisms and cells, particularly mammalian cells used for the production
of therapeutic antibodies. Alternative cells to meet the catalytic needs of metabolic engi-
neering for greener chemistries, wastewater treatment, and biomass conversion are the
focus of other chapters of this report (Chapters 3, 4, and 6, respectively). For therapeutic
applications, cell-free systems may provide opportunities to address on-demand therapeu-
tics and orphan disease treatments (e.g., Swartz, 2012). As an alternative to mammalian
cells, microbial hosts provide the advantages of reduced production costs and shorter pro-
cess times in both development and production.
For antibodies, the addition of complex sugars, or glycosylation, particularly of a
type that addresses efficacy and lack of immunogenicity, has limited the broader applica-
bility of nonmammalian systems. Alternative cell types need to allow rapid growth; au-
thentic protein processing and secretion; and the ability to engineer the cell readily, in-
cluding with the use of genomic tools; these cell types also need to lack mammalian viral
infectivity (Matthews et al., 2017). A number of chemical engineers have already made
compelling cases for and shown successes in developing and implementing genetic tools
for the use of yeasts, from Pichia pastoris to K. lactis, Y. lipolytica, and K. phaffi (Ham-
ilton et al., 2006; Jiang et al., 2019; Miller and Alper, 2019; Panuwatsuk and Da Silva,
2003); S. cervisiae and P. pastoris are already FDA-approved production organisms for
vaccines and cytokines. To enable widespread application, continued improvements in
volumetric productivities are needed, as are improved genetic tools and models for me-
tabolism, as well as a better understanding of why some products are not well expressed.
More broadly, developing technologies or tools that replicate the capabilities of
their expensive counterparts—so-called frugal science—enables out-of-the-box solutions
for global science and health care. For example, considering the different parts of key
devices and creating alternative low-cost parts resulted in the creation of a $1 hearing aid
(Sinha et al., 2020) and a hand-powered paper centrifuge (<$0.25) that enables blood sep-
aration in resource-poor settings (Bhamla et al., 2017).

CHALLENGES AND OPPORTUNITIES

Modern biomolecular engineering is very much at the intersection of chemical


engineering and molecular biology, biochemistry, materials science, and medicine. Cur-
rent challenges for applications in health and medicine include advancing personalized
medicine and the engineering of biological molecules, including proteins, nucleic acids,
and other entities such as viruses and cells; bridging the interface between materials and
devices and health; improving the use of tools from systems and synthetic biology to un-
derstand biological networks and the intersections with data science and machine learn-
ing; developing the next steps in manufacturing; and using engineering approaches to ad-
dress equity and access to health care. All of these challenges present opportunities for
chemical engineers to apply systems-level approaches and their ability to work across

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Engineering Targeted and Accessible Medicine 149

disciplines. Specific opportunities include the application of systems engineering to bio-


logical systems in such areas as

 sensor design and analysis;


 identification of fault detection (using physiological, cellular, metabolic, or
other data to identify changes to function);
 process modeling to represent complex relationships for biological systems
and predict behavior; and
 understanding and modification of molecular pathways and genomic net-
works involved in the regulation of normal physiology, as well as disease.

Opportunities to apply quantitative chemical engineering skills to immunology


include cancer immunotherapies, vaccine design, and therapeutic treatments for infectious
diseases and autoimmune disorders. The development of completely noninvasive meth-
ods for drug delivery represents an exciting frontier of device- and materials-based strat-
egies. Chemical engineers are also well positioned to advance work on sustained-release
depots and targeted delivery of therapeutics.
The demand for mAbs, therapeutic proteins, and mRNA therapeutics continues
to grow, in part because of the increasing average age of people in the United States.
Unfortunately, the costs to produce biologics and the subsequent costs to the consumer
create pressure to improve flexibility, reduce costs, and increase health care equity while
maintaining reliability and stability during manufacturing and distribution. This challenge
provides an opportunity for chemical engineers to develop novel bioprocess and cell-
based improvements through collaborations with biologists and biochemists.
Across all these areas, interdisciplinary and cross-sector collaborations will be
critical, as will coordination across the federal agencies with responsibilities in these ar-
eas. An additional major challenge for chemical engineers is their lack of awareness of
health disparities. Introducing these issues in the classroom will enable future chemical
engineers to play an active role in reducing health inequities. Furthermore, increased di-
versity of both students and instructors in the classroom will provide a broader perspective
on the challenges requiring engineering solutions.

Recommendation 5-1: Federal research investments in biomolecular engineering


should be directed to fundamental research to

 advance personalized medicine and the engineering of biological mole-


cules, including proteins, nucleic acids, and other entities such as viruses
and cells;
 bridge the interface between materials and devices and health;
 improve the use of tools from systems and synthetic biology to under-
stand biological networks and the intersections with data science and
computational approaches; and
 develop engineering approaches to reduce costs and improve equity and
access to health care.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

150 New Directions for Chemical Engineering

Recommendation 5-2: Researchers in academic and government laboratories and


industry practitioners should form interdisciplinary, cross-sector collaborations to
develop pilot- and demonstration-scale projects in advanced pharmaceutical manu-
facturing processes.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

6
Flexible Manufacturing and the Circular Economy

 Chemical engineering as a discipline was founded in the need to deal with hetero-
geneous feedstocks, especially petroleum, and this need will be amplified in the
transition to more sustainable feedstocks.
 The production and manufacturing of useful materials and molecules enabled by
chemical engineers are now creating problems that must be solved at scale.
 Chemical engineers play a critical role in manufacturing and can thus contribute to
more sustainable manufacturing through efficiency, nimbleness, and process in-
tensification.
 The principles of green chemistry and green engineering will be important in the
shift from molecular to larger system scales, and to more sustainable manufactur-
ing and a circular economy.

The chemical engineering discipline is broadly concerned with enabling realistic,


cost-effective, efficient, and safe physical and chemical transformations of matter into
more useful molecules or materials. In the last century, the discipline of chemical engi-
neering enabled transformations of the entire landscape of both modern society and the
planet. This chapter focuses on some key examples of the challenges and opportunities
for chemical engineering in moving toward more flexible manufacturing and a circular
economy.
Manufacturing is generally defined as the synthesis and formulation of useful
products. In the last century, chemical engineers revolutionized manufacturing across all
sectors of the economy, including agrochemicals and fertilizers, cement, consumer goods,
flavors and fragrances, food and feed, fuels, paints and coatings, paper and pulp, pharma-
ceuticals and biologics, polymers, semiconductors, and many others. To give some idea
of scale, for plastics alone—a prominent example discussed later in this chapter—the
mass of these synthetic materials manufactured in just the last 70 years now exceeds the
mass of all living animals on the planet (Elhacham et al., 2020). Additionally, the chemi-
cals and materials manufactured at scale today contain substantial embodied energy and
produce significant greenhouse gas (GHG) emissions (Figure 6-1). These two metrics are
important benchmarks, along with issues of environmental and social justice and supply
chain resilience, for calibrating new technology development that can mitigate anthropo-
genic climate change.
Given their critical role in manufacturing, chemical engineers have many oppor-
tunities to increase its environmental sustainability. This chapter provides an overview of
the intersection of manufacturing and chemical engineering, followed by a discussion of
feedstock flexibility, distributed manufacturing and process intensification, and the im-
portance of transitioning from a linear to a circular economy.

151

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

152 New Directions for Chemical Engineering

FIGURE 6-1 CO2 emissions, in kg CO2/kg material, and embodied energy, in MJ/kg material, for
various materials. SOURCE: Gutowski et al. (2013).

INTERSECTION OF MANUFACTURING AND CHEMICAL ENGINEERING

The intersection of manufacturing and chemical engineering is founded on the


systems-level, quantitative approach intrinsic to the profession (Peters et al., 2002; Turton
et al., 2012). This approach includes the ability to conduct rigorous mass and energy bal-
ances, coupled with appropriate technoeconomic assessment (TEA) and life-cycle assess-
ment (LCA). The tools of TEA and LCA enable detailed analysis of developed processes
to identify potential efficiency gains from reducing cost, energy use, or material inputs.
These efficiency gains are accomplished by an improved ability to recycle materials or by
the transition from batch to continuous processes, among other changes. TEA and LCA
tools are also critical in manufacturing to identify “leapfrog” processes that can displace
current methodologies (TEA and LCA are described in more detail in Chapter 8). More
recently, consideration of environmental and social justice has become increasingly im-
portant for chemical engineers in these analyses.
In addition to TEA and LCA, the overall principles of green chemistry (Anastas
and Warner, 1998) and green engineering (Anastas and Zimmerman, 2003), highlighted
in Box 6-1, provide qualitative guidelines that chemical engineers can place in a quanti-
tative, objective context with systems-level approaches. This capability ultimately enables
more efficient and responsible manufacturing practices and better decision making re-
garding trade-offs.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Flexible Manufacturing and the Circular Economy 153

BOX 6-1
Principles of Green Chemistry and Green Engineering

Prevention—it is better to prevent waste than to treat or clean up waste after it has been created.
Atom Economy—synthetic methods should be designed to maximize the incorporation of all
materials used in the process into the final product.
Less Hazardous Chemical Syntheses—wherever practicable, synthetic methods should be de-
signed to use and generate substances that possess little or no toxicity to human health and the
environment.
Design Safer Chemicals—chemical products should be designed to effect their desired function
while minimizing their toxicity.
Safer Solvents and Auxiliaries—the use of auxiliary substances (e.g., solvents, separation
agents) should be made unnecessary wherever possible and innocuous when used.
Design for Energy Efficiency—energy requirements of chemical processes should be recog-
nized for their environmental and economic impacts and should be minimized. If possible, syn-
thetic methods should be conducted at ambient temperature and pressure.
Use of Renewable Feedstocks—a raw material or feedstock should be renewable rather than
depleting whenever technically and economically practicable.
Reduce Derivatives—unnecessary derivatization (use of blocking groups, protection/deprotec-
tion, temporary modification of physical/chemical processes) should be minimized or avoided
if possible, because such steps require additional reagents and can generate waste.
Catalysis—catalytic reagents (as selective as possible) are superior to stoichiometric reagents.
Design for Degradation—chemical products should be designed so that at the end of their
function they break down into innocuous degradation products and do not persist in the envi-
ronment.
Real-time Analysis for Pollution Prevention—analytical methodologies need to be further
developed to allow for real-time, in-process monitoring and control prior to the formation of
hazardous substances.
Inherently Safer Chemistry for Accident Prevention—substances and the form of a sub-
stance used in a chemical process should be chosen to minimize the potential for chemical ac-
cidents, including releases, explosions, and fires.
Inherent Rather than Circumstantial—designers need to strive to ensure that all materials
and energy inputs and outputs are as inherently nonhazardous as possible.
Design for Separation—separation and purification operations should be designed to minimize
energy consumption and materials use.
Maximize Efficiency—products, processes, and systems should be designed to maximize mass,
energy, space, and time efficiency.
Output-Pulled Versus Input-Pushed Products—process and systems should be “output
pulled” rather than “input pushed” through the use of energy and materials.
Conserve Complexity—embedded entropy and complexity must be viewed as an investment
when making design choices on recycle, reuse, or beneficial disposition.
continued

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

154 New Directions for Chemical Engineering

BOX 6-1 continued

Durability Rather than Immortality—targeted durability, not immortality, should be a design


goal.
Meet Need, Minimize Excess—design for unnecessary capacity or capability (e.g., “one-size-
fits-all” solutions) should be considered a design flaw.
Minimize Material Diversity—material diversity in multicomponent products should be min-
imized to promote disassembly and value retention.
Integrate Material and Energy Flows—design of products, processes, and systems must in-
clude integration and interconnectivity with available energy and material flows.
Design for Commercial Afterlife—products, processes, and systems should be designed for
performance in a commercial “afterlife.”
Renewable Rather Than Depleting—material and energy inputs should be renewable rather
than depleting.

SOURCES: Green Chemistry (Anastas and Warner, 1998); Green Engineering (Anastas and
Zimmerman, 2003).

Safety and safe operations are the most critical responsibility of the chemical en-
gineering field. Safety is more important than reaching the goals of improving efficiency,
increasing cost-effectiveness, and lowering environmental impacts of manufacturing pro-
cesses. Simply put, many of the centralized industrial manufacturing facilities that chem-
ical engineers have enabled over the last century handle gases, liquids, and solids at such
scales and under such operating conditions that an accident can harm people and damage
local and regional environments. Strict adherence to safety, including its inclusion in
chemical engineering education, is essential to the discipline and needs to be rigorously
maintained from the laboratory to the refinery.
Most conventional manufacturing processes in a chemical engineering context
are operated at extremely large scale and in capital- and operating-intensive centralized
facilities to harness economies of scale. However, advances in the valorization of flexible
feedstocks, electrification of the manufacturing sector (Schiffer and Manthiram, 2017),
and the concepts of scale-out and distributed manufacturing will play a key role for chem-
ical engineering going forward. Some of these concepts are likely to play major roles in
the deployment of manufacturing to low- and middle-income countries and to economi-
cally depressed regions of higher-income countries, as well as in various efforts at reshor-
ing of manufacturing through new technologies. Indeed, industrial manufacturing has the
potential in this century to at least partially transform physically from the scale of the
petrochemical complexes studied by today’s undergraduate chemical engineers to more
heterogeneous intensified and distributed manufacturing sites, including those with elec-
trically driven power sources. Notably, flexible and distributed manufacturing have al-

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Flexible Manufacturing and the Circular Economy 155

ready been a major focus of research and development among the international commu-
nity, but substantial opportunities remain for the U.S. chemical engineering community
to contribute meaningfully in these areas.
Lastly, the concept of the circular economy is commonplace in today’s vernacu-
lar, including across many STEM (science, technology, engineering, and mathematics)
professions. The Ellen MacArthur Foundation defines the circular economy as a systemic
approach to economic development designed to benefit businesses, society, and the envi-
ronment; in contrast to the “take-make-waste” linear model, a circular economy is regen-
erative by design and aims to gradually decouple growth from the consumption of finite
resources (EMF, 2021). This broad concept has a figurative home squarely in the chemical
engineering approach to manufacturing and green engineering principles because the
ideas of materials recycling and, more generally, the efficient use of matter and energy
are central to the field’s systems-level thinking. Transitioning manufacturing from a linear
to a circular economy is a key opportunity for chemical engineers.

FEEDSTOCK FLEXIBILITY FOR MANUFACTURING


OF EXISTING AND ADVANTAGED PRODUCTS

The chemical engineering profession emerged in large part to confront the urgent
challenges faced more than a century ago in the then-burgeoning petroleum refining in-
dustry. Petroleum is a highly heterogeneous organic resource. The global-scale petro-
chemical refining industry, in concert with the chemical engineering profession, was thus
born out of the ability to characterize, fractionate, and ultimately valorize feed streams
that are highly diverse in chemistry and that change as a function of time and source.
Indeed, many of the original fractionated streams that could be derived from petroleum
processing as a consequence of producing the original target fuels were considered waste.
The ingenuity of chemists and chemical engineers, however, gave rise to uses for these
waste compounds, including such diverse applications as asphalt, building blocks for syn-
thetic polymers, and such formulated products as lubricants and processing fluids. These
and myriad other high-value chemical applications today form the highly profitable chem-
icals backbone of the petrochemical business. The scale of the petrochemical industry
worldwide is staggering: in 2019 its annual production volumes were 5–5.5 billion metric
tons (100.69 million barrels per day) of crude oil, 4.1 trillion cubic meters (3.6 billion tons
of oil equivalent) of natural gas, and 7.96 billion metric tons of coal (IEA, 2021c,d; EIA,
2021d).
The continued drive toward more efficient, environmentally friendly, and cost-
effective manufacturing processes will likely benefit greatly from a much wider range of
available feedstocks for use as building blocks to produce the chemicals and materials
demanded by modern society (Figure 6-2). This concept of feedstock flexibility can be
broadly defined as the input of diverse feedstocks into a transformation process that is
able to process various starting compounds or mixtures of compounds to produce the tar-
get product. Petrochemical refineries already do this today. Importantly, today’s petro-
leum feedstock requires oxidative chemistry to produce oxygenated molecules from hy-
drocarbons. For a different feedstock that is already oxygenated, such as lignocellulosic

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

156 New Directions for Chemical Engineering

biomass, reductive chemistry would be needed to manufacture such products as hydro-


carbons.
The use of a flexible feedstock also influences ideas around distributed manufac-
turing, as discussed below. For example, the availability of stranded natural gas resources,
along with the potential harm of leakage of GHGs from those resources, makes conversion
of these streams via chemical, biological, electrochemical, or other means a key oppor-
tunity for chemical engineers. The use of nonthermal approaches requiring minimal utility
infrastructure may be critical for the ultimate feasibility of small-scale and distributed
harnessing of such feedstocks as stranded natural gas; industrial waste gases; and indus-
trial, commercial, or municipal wastewater (Khalilpour and Karimi, 2012; Tuck et al.,
2012). This work also includes, as shown in Figure 6-2, the use of selective transfor-
mations afforded by biological and chemocatalytic transformations. Indeed, biological
transformations of conventional feedstocks, combined with knowledge from the environ-
mental bioremediation community, can show how bioprocessing can be incorporated into
the petrochemical industry.
Flexible feedstock sources beyond those derived from fossil fuels include large
amounts of available biogenic and waste carbon inputs, such as municipal solid waste
(MSW). The goal of integrating biobased pyrolysis oil made from lignocellulosic biomass
or MSW into a petroleum refinery has been pursued for several decades (Chen et al., 2014;
Talmadge et al., 2014), despite challenges with inorganic foulants and catalyst poisons
common in biogenic and waste carbon. The chemical engineering community has a key
opportunity to understand how biogenic and waste-based substrates affect current manu-
facturing infrastructure. And much more room is available to explore the concepts of “re-
finery integration” of stabilized, biogenic and/or waste-based, carbon-rich streams into
the existing, mostly amortized petrochemical complex.

FIGURE 6-2 Possible treatment and transformation pathways for more flexible feedstock sources
and feedstocks. NOTE: F&O = fats & oils.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Flexible Manufacturing and the Circular Economy 157

More broadly, existing waste streams of biogenic and waste-based carbon could
serve as useful feed streams for leapfrog technologies. Chemical engineers are already
playing critical roles in the harvesting, densification, conversion, and scale-up of innova-
tive processes for leveraging these biogenic and waste-based feedstocks. As discussed in
Chapter 3, by 2030 there will be an estimated 1 billion dry tons of lignocellulosic biomass
annually in the United States alone that can potentially be sustainably harvested (DOE,
2016). This potentially large feedstock, along with large amounts of other available wet
waste (e.g., food waste, manure, sludge, fats, oils, and greases total are ~700 million tons
per year; Milbrandt et al., 2018), is distributed across the United States. Not only does its
use offer the potential to produce meaningful amounts of transportation biofuels to offset
substantial GHG emissions (Chapter 3), but it also could serve as a key feedstock for
chemical manufacturing. Products could include both direct-replacement biochemicals
(Nikolau et al., 2008) and biochemicals that do not resemble their fossil-based counter-
parts but offer a performance advantage (Cywar et al., 2021; DOE, 2018a). MSW also
offers a substantial and important feedstock, which again is highly heterogeneous.
The manufacturing of direct-replacement chemicals offers substantial opportuni-
ties for chemical engineers to develop scaled-out, distributed manufacturing systems and
innovative, large-scale processes that can compete with the conversion of fossil resources.
Conversely, performance-advantaged bioproducts could also serve as economic incen-
tives to invest in new capital infrastructure at scale to displace the fossil carbon-based
feed streams that dominate chemicals and materials production at scale today. These tar-
get bioproducts are an opportunity for chemical engineers to develop fully integrated,
novel processes for transforming typically highly oxygenated feedstocks (e.g., sugars, ar-
omatics derived from lignin, algae biomass) into novel molecules and new materials for
which the properties often are not known a priori. Indeed, performance-advantaged bi-
oproducts can offer potential benefits along the entire value chain (from feedstock to man-
ufacturing, use, and end of life) relative to incumbent chemicals and materials, thus
providing product design, economic, and environmental benefits (Cywar et al., 2021;
DOE, 2018a). Systems-level approaches will be necessary to understand the ultimate po-
tential of a given process concept or early demonstrations to reach scalability for manu-
facturing processes (Cywar et al., 2021).
From a process perspective, the conversion of heterogeneous feedstocks of essen-
tially any type into valorized end products can be broadly categorized into the “fraction-
ate-first” approach that the petrochemical complex has successfully adopted or a “one-
pot” approach that attempts to convert all feedstocks simultaneously. The latter includes
such conversion approaches as hydrothermal liquefaction, pyrolysis, and gasification.
While TEA and LCA, along with the demonstrable technical feasibility of a given process,
will ultimately and quantitatively inform how various processes are adopted, scaled, and
enabled, many opportunities exist to define new flowsheets using emerging tools in elec-
trochemistry, photochemistry, synthetic biology, integrated separations and catalysis, and
many other tools that are familiar to chemical engineers. These approaches, and combi-
nations thereof, present opportunities to significantly change the manufacturing land-
scape.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

158 New Directions for Chemical Engineering

From a chemistry perspective, new feedstocks—especially those related to ligno-


cellulosic and algal biomass, wet organic waste, and CO2 and other waste gases—are of-
ten highly oxidized relative to conventional fossil feedstocks. Many common existing
transformations add heteroatoms (e.g., nitrogen and oxygen) to hydrocarbon feedstocks
to manufacture products. The use of new feedstocks that are more oxygen rich (and po-
tentially contain other heteroatoms) offers the opportunity to develop new and novel trans-
formation processes. These transformations could also take place in a condensed rather
than a gas phase, the latter of which is typical of most conventional hydrocarbon pro-
cessing. New separation technologies will be critical to realizing these transformations,
as will catalysts that can enable the necessary reductive chemistry while remaining stable
in aqueous environments.
Beyond innovative process and chemistry developments, there are fundamental
research problems for chemical engineers to solve in the feedstock flexibility arena. A
prominent example that hinders systems-level work is the lack of robust thermodynamic
data in common process-modeling packages for the molecules found in biobased or waste-
based feedstocks. Such data are plentiful for the hydrocarbon-rich feedstocks of relevance
to petrochemical refining. However, the ability to model important thermodynamic prop-
erties of other molecules, which are often richer in oxygen and other heteroatoms, poses
a substantial challenge, as does their incorporation into process simulators.

PROCESS INTENSIFICATION AND DISTRIBUTED MANUFACTURING

Many of chemical engineering’s historical successes involve the efficient produc-


tion of chemicals on a large and therefore economical scale. A world-scale ammonia plant,
for example, generates 1,000 metric tons of ammonia each day. The design tools needed
to scale up individual unit operations in these processes are well established and are a key
focus of undergraduate chemical engineering education. Many opportunities exist to de-
velop novel processes, which can potentially improve performance but have not always
been used in commodity-scale plants.
Process intensification (PI) is designed to create improved chemical processes by
moving beyond the idea that a single piece of equipment performs a single unit operation.
In a traditional chemical plant, for example, a reactor and a distillation column might be
used separately for a reaction and a separation step in an overall process. Reactive distil-
lation (e.g., Taylor and Krishna, 2000) and membrane reactors (e.g., Iulianelli et al., 2016)
are two alternative PI strategies that combine these two steps into a single process. PI also
encompasses efforts to use nontraditional driving forces to accomplish unit operations—
for example, microwave heating in reactions (e.g., Goyal et al., 2019) or the use of struc-
tured contactors in adsorption-based separations (e.g., Koros and Lively, 2012). Chemical
engineers have numerous opportunities to use these strategies to develop innovative new
processes, as well as to remove bottlenecks from existing large-scale processes. (Exam-
ples relevant to electronic-materials manufacturing are highlighted in Box 6-2). While PI
presents many opportunities for innovation, it is also necessary to acknowledge that there
are often trade-offs between PI and process flexibility, which may be more important for
the flexible feedstocks discussed elsewhere in this chapter.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Flexible Manufacturing and the Circular Economy 159

BOX 6-2
Process Intensification in the Electronic Materials Industry

Process intensification (PI) has been a major technology focus since the 1990s (Creative
Energy, 2007). PI reduces the size and cost of chemical production modules. The short time
scale for the commercialization of electronic materials and the rapid ramp-up of the demand
cycle require manufacturing solutions that can be scaled up quickly to meet the demand. Prep-
ositioning of modular assets becomes economical when the process equipment is compact, self-
contained, and standardized (Bielenberg and Bryner, 2018; Stankiewicz and Moulijn, 2000).
An important solvent used in the electronic industry—propylene glycol methyl ether acetate
(PGMEA)—has been the subject of numerous applications of PI to reduce waste, improve qual-
ity, reduce impurities, and lower costs, including catalytic distillation using a fixed-bed catalyst
integrated into the distillation packing, which reacts propylene glycol monomethyl ether with
methyl acetate (Hsieh et al., 2006; see Figure 6-2-1).

FIGURE 6-2-1 Reactive distillation supports a novel intensified process for the manufacture of
a key solvent used in the electronics industry. SOURCE: Hussain et al. (2019).

Another example of PI in manufacturing of electronic materials is the use of external accel-


eration to eliminate the limitation of the gravity driving force in phase separation. Extraction
and distillation can be compressed in small spaces using high-gravitational rotating packed beds.
Nanoparticles are useful in many electronic-material processing steps, and monodisperse CaCO3
nanoparticles can be made in high-gravity rotating packed-bed reactors. Breakthroughs are often
achieved by applying existing technologies from different chemical and materials production
industries to a new area (Cortes Garcia et al., 2017; Kang et al., 2018; Rao, 2015).
Nitrogen trifluoride is used as a cleaning and key etchant gas. The first large-scale industrial
process used a two-step synthesis in which hydrofluoric acid was first fed to a traditional elec-
trolysis cell to produce fluorine gas. The purified fluorine was then fed simultaneously with
ammonia into a continuous stirred tank reactor, which produces nitrogen trifluoride and a solid
waste stream of ammonium fluoride. In an intensified process, ammonia and hydrofluoric acid
are added directly to a modified fluorine electrochemical cell to produce nitrogen trifluoride
without a solid waste stream (Coronell et al., 1997; Hart et al., 2015; Krouse et al., 2016). This
combination of reaction steps and the attendant simplification of process equipment reduce
waste and lower capital cost.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

160 New Directions for Chemical Engineering

Four principles for PI design have guided thinking about how chemical processes
are developed (Harmsen, 2007; Tian et al., 2018):

 Maximize the effectiveness of intra- and intermolecular events. Improving


process kinetics is a major principle for obtaining higher process perfor-
mance, as it is usually the underlying limiting factor for low conversion and
selectivity.
 Give each molecule the same processing experience, which results in prod-
ucts with uniform properties. Uniform product distributions facilitate waste
reduction, which in turn reduces the efforts required for product separation.
 Optimize the driving forces at every scale, and maximize the specific surface
area to which these forces apply. Doing so results in more efficient processes
that use lower amounts of enabling materials, which then leads to reductions
in equipment sizes.
 Maximize the synergistic effects of partial processes that enable multitasking.
By combining several processing tasks, higher process efficiencies can be
achieved compared with their stand-alone counterparts.

The general categories of applicable technologies include structured devices (e.g.,


structured catalyst-based reactors, microreactors, nonselective membrane reactors), hy-
brid processes (e.g., extractive crystallization, heat-integrated distillation, reactive distil-
lation, selective/catalytic membrane reactors), energy transfer processes (e.g., rotating
packed beds, sonochemical reactors, microwave-enhanced operations), dynamic pro-
cesses (e.g., oscillatory baffled reactors, reverse flow reactor operation), and others (e.g.,
supercritical reactions, cryogenic separations; Harmsen, 2007). Many catalytic processes
are also good candidates for PI technologies (e.g., Boger et al., 2003; Broekhuis et al.,
2001; Cybulski and Moulijn, 1994; Kapteijn and Moulijn, 2020; Machado and Broekhuis,
2003; Machado et al., 2005; Nordquist et al., 2002; Welp et al., 2006, 2009). For example,
slurry catalytic processes in the specialty chemical and pharmaceutical industry with gas
and one or two liquid phases can be intensified by replacing the slurry catalyst with a
fixed monolith catalytic reactor. These reactors can be installed in a pump-around loop to
existing classic stirred-tank reactors to allow modularization and the ability to operate a
single stirred tank with multiple beds using a monolith-loop reactor arrangement. Finally,
microreactors are a more recent trend in chemical reactor synthesis. Their small channels
allow for extremely accurate temperature control and high mixing intensity for single-
and multiphase reaction processes. Scale is achieved by increasing the numbers of micro-
reactor systems. The continuous flow reaction and separation networks allow for adequate
production to meet most demand.
Traditional large processes increase product volume by scaling up. An alternative
strategy achieves product volume by scaling out the deployment of many compact pro-
cessing units in parallel. This is the key aim of modular manufacturing. Scaling out has
benefits over scaling up in such processes as

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Flexible Manufacturing and the Circular Economy 161

 water treatment, both in municipal settings and for produced water from oil
or gas wells;
 upgrading of natural gas from remote wells where pipeline infrastructure for
gas transport is problematic;
 processing of bioproducts where transportation costs play a significant role
in net GHG emissions;
 pyrolysis of waste polyolefins where the size of the pyrolysis reactors is lim-
ited by heat transfer considerations; and
 industrial sectors, such as pharmaceuticals or electronic materials, where
highly valuable products are often produced in small quantities.

Although examples in pharmaceutical processes typically use highly controlled


feedstocks, the other examples are cases in which modular processes need to function
despite significant variations in the availability and location of the process feedstock. Ad-
ditionally, scaling out is sometimes necessary because of technical realities, while in other
cases it offers an economic advantage. Pyrolysis of waste polyolefins is operated in
scaled-out plants because of the heat transfer considerations; however, the price of the
pyrolysis oil increases significantly in this context. On the other hand, and in the same
example, collecting plastic waste over a large area and transporting it to a pyrolysis plant
add cost that the benefit of a scaled-up plant might not be able to counterbalance. In this
latter case, a distributed network of pyrolysis plants makes economic sense.
PI and modular manufacturing are important areas in which the chemical engi-
neering research community can provide intellectual leadership. A challenge for the aca-
demic community is that the successes (or failures) of work in either area are inherently
determined at a process scale. Both demonstrating that a process is possible at “lab scale”
and combining such work with process modeling and/or LCA are needed to support large
investments. Also critical is the development of new materials and processes that can be
deployed readily at the requisite scale and cost for use in the target processes. To give just
one example, individual membrane modules used in current water treatment applications
have surface areas measured in hundreds or thousands of square meters. Any putative new
membrane that cannot be produced readily and economically at this scale will simply have
no meaningful impact in water treatment, regardless of how superlative its performance
may be (Koros and Zhang, 2017). This observation highlights the importance of pursuing
research focused on the manufacturability of modular components, not just on the devel-
opment of new high-performance materials. This is also an area in which close working
relationships between academic researchers and industrial practitioners can be fruitful.
Rapid advances in additive manufacturing (e.g., 3-D printing; see Box 6-3) have
also opened up a wide range of possibilities for generating new devices used in chemical
processes. These possibilities are perhaps especially rich in the realm of PI and modular
manufacturing. Because much work in additive manufacturing is taking place in the me-
chanical engineering and materials science communities, chemical engineers have numer-
ous opportunities to combine expertise from these adjacent fields with application-spe-
cific needs in chemical processing.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

162 New Directions for Chemical Engineering

BOX 6-3
Additive Manufacturing

Additive manufacturing (AM), typically accomplished by a form of patterned layer depo-


sition, is increasingly used to create a range of products, from dentures to mobile homes. These
methods (e.g., vat polymerization, material extrusion, material or binder jetting, powder bed
fusion) provide versatility and customization beyond what has historically been achieved with
conventional manufacturing. The ease, flexibility, and distributed access of AM enabled support
for the health care community during the early months of the COVID-19 pandemic, particularly
in the production of nasal swabs and other devices, such as face shields, that were not otherwise
able to replace standard medical equipment (America Makes, AMT, and Deloitte, 2021).
Improvements in vat polymerization stereolithography in particular have enabled a contin-
uous form of AM, offering an increase in speed and quality over layer-based approaches, albeit
with the technology’s own size constraints (e.g., de Beer et al., 2019; Tumbleston et al., 2015).
This technology depends on the existence of an inhibition volume in which there is no polymer-
ization between the product and the cross-linking light source to replenish the source polymer.
Chemical engineers are particularly well suited to addressing the fluid transport challenges of
this medium.
AM or 3-D printing has opened up new avenues for chemical engineers and materials sci-
entists to create novel structures as research prototypes or products. A key enabling feature of
3-D printing has been the ability to customize the structure of the material in question and go
from the design phase directly to manufacturing without having to go through the prototype
phase. The last decade has witnessed particular growth in this field, with 3-D printing evolving
from a “boutique” tool to one that can serve machine shop and manufacturing needs. Techno-
logical advances over the last decade have improved the speed and scale of 3-D–printed objects
using various polymeric materials and metals. This capability is particularly important for print-
ing custom objects from individualized specifications. The ability to transform digital structures
into their physical form makes 3-D printing especially indispensable to 21st-century research
and manufacturing.
3-D printing has also been actively pursued for printing organs. Organs, by nature, are char-
acterized by 3-D architectures needed for their respective biological functions (e.g., chambers
and valves for the heart, nephron tubes for kidneys). Current approaches for organ engineering
rely on the use of appropriate templates that provide initial structural support for cell growth,
but then rely on the natural ability of cells to self-organize into the right tissue architecture. 3-D
printing provides an additional layer of architectural control. Combinations of materials and
cells have been 3-D printed to produce organoids that capture some of the essential biological
and structural features of their natural counterparts.
Opportunities for chemical engineers in AM broadly include supporting advances for faster
printing with higher resolution, as well as use of multiple or more advanced and sustainable
materials with attention to end of life. These efforts are best undertaken in partnership with
mechanical and software engineers, collaboratively improving the technologies that scale these
processes for widespread use.

Lastly, many modular processes (e.g., the treatment of stranded natural gas) are
likely to occur in remote locations where monitoring or access by highly trained operators

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Flexible Manufacturing and the Circular Economy 163

is limited. This fact, together with the issue of feedstock flexibility discussed above, high-
lights the need to develop modular processes that are inherently robust to process upsets
and that take full advantage of advanced instrumentation and control strategies. Modular
processes could leverage electrochemical power generated at small scale and on-site
and/or biological manufacturing, which can often require lower heat and power inputs.

THE CIRCULAR ECONOMY AND DESIGN FOR END OF LIFE

The Industrial Revolution, beginning with the invention of the steam engine in
the 17th century, enabled the use of raw materials and energy—which seemed to be infi-
nitely available at the time—to make and eventually to mass produce products. The re-
sulting economy was thus primarily linear (i.e., take → make → dispose), with resources
being extracted or harvested, energy being used to make products, and products being
disposed of at their end of life (Collias et al., 2021). The world’s linear economy annually
generated about 110 million metric tons of MSW in 1900, more than 1 billion metric tons
in 2000 (including about 80 percent of consumer products, excluding packaging, disposed
of after a single use), and about 2 billion metric tons in 2016, and this number is forecast
to grow to 3.4 billion metric tons by 2050 (Hoornweg et al., 2013; Kaza et al., 2018). The
consequences of the growth of the current economy are unsustainable.

The Circular Economy

A sustainable future requires a transition to a circular economy. In that model,


resources are managed differently than in the linear economy, the way products are made
and used changes, and new consideration is given to the fate of products and materials at
their end of life. The circular economy uses waste streams as sources of secondary re-
sources, and it incorporates the principles of green chemistry and engineering (Box 6-1;
Anastas and Warner, 1998; Anastas and Zimmerman, 2003; Collias et al., 2021).
In the circular economy model, materials and products are made efficiently and
reused, thus preventing waste. If new raw materials are needed, they are produced sus-
tainably. This model represents a paradigm shift that is consistent with emerging con-
sumer preferences for recycling and reducing nonrecyclable waste (Nielsen, 2015), as
well as new government restrictions on pollution and waste. A circular economy can be
facilitated by emerging digital technologies and product designs that track materials and
products and extend their lives. Decoupled from the consumption of finite resources, it
can drive economic growth with business, societal, and environmental benefits. In the
long term, the circular economy will achieve cost savings in materials that result in cost
savings in products and packaging. For consumers, the benefits of the circular economy
will come from new developments and trends in such areas as urbanization, technologies
(e.g., anaerobic digestion), information technology capabilities, online retail sales, busi-
ness models, and packaging technologies (EMF, 2014).

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

164 New Directions for Chemical Engineering

The circular economy model encompasses two cycles: biological and technical
(Figure 6-3). In the biological cycle, food and biological materials (e.g., wood) feed back
into the system through recycling processes, such as composting and anaerobic digestion,
that regenerate living systems, such as the soil. In the technical cycle, such strategies as
recycling, reuse, repair, and remanufacturing allow the recovery and restoration of prod-
ucts, components, and materials (EMF, 2013a,b, 2014).
The circular economy is based on three strategies (EMF, 2017a,b) that are con-
sistent with the principles of green chemistry and engineering:

 Strategy 1: Design to avoid pollution and waste—Reduce or eliminate GHG


emissions and hazardous substances and the resulting pollution of air, land,
and water. Also limit, if not eliminate, waste of materials during the manu-
facturing of products and packaging, as well as during the discarding of prod-
ucts and packages at their end of life.
 Strategy 2: Extend useful life—Design for durability, reuse, remanufacturing,
and recycling to keep materials, products, and packaging circulating in the
economy as long as possible, thus preserving energy and materials.
 Strategy 3: Regenerate natural systems—There is synergy between a circular
economy and a biobased economy. In the biobased economy, biobased ma-
terials are made from biobased resources and/or with the use of biobased en-
ergy, and these biobased materials cycle between the economy and natural
systems (to become biobased resources). Ideally, the circular economy uses
renewable and avoids the use of nonrenewable resources.

FIGURE 6-3 A circular economy system diagram showing the flow of materials, nutrients, com-
ponents, and products (biobased materials are shown in green). SOURCE: EMF (2019).

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Flexible Manufacturing and the Circular Economy 165

Each of these strategies is described further below, with a focus on plastics as an


example of an area in which chemical engineers will play a key role. The essential prin-
ciples of the circular economy, however, apply beyond the manufacture of synthetic pol-
ymers.
Chemical engineers are uniquely positioned to solve problems associated with the
three strategies of the circular economy. Their contributions could include redesigning
processes and products to reduce or eliminate pollution, developing new ways to reduce
and utilize waste, designing products to be used longer, and designing processes and prod-
ucts using sustainable feedstocks. Besides providing technical expertise and leadership in
various areas of the circular economy, chemical engineers have opportunities to address
challenges in the following four more specific technology areas:

 purification of materials with a large volume and a high rate of collection,


such as paper, cardboard, polyethylene terephthalate (PET), glass, and steel;
 recycling of polymers with a large volume and a low rate of collection;
 utilization of by-products of manufacturing processes, such as used concrete,
CO2, and food waste; and
 development of materials with high value that currently have a small volume
and a low rate of collection, such as 3-D printing materials and biobased ma-
terials.

Design to Avoid Pollution and Waste

Pollution and waste are associated with the production of all materials, products,
and packaging. To show how concepts associated with the circular economy can be ap-
plied to reducing energy consumption and GHG emissions, it is useful to consider the
opportunities for plastics.1 More than 90 percent of the feedstock for the plastics industry
is petroleum or natural gas. About 6 percent of the world’s production of petroleum and
natural gas liquids (NGLs) is used to produce plastics, with about half used as feedstock
and the other half as fuel in the production processes (EMF, 2016). Another significant
percentage of natural gas production is used as feedstock and fuel for plastics manufac-
turing.
Since the early 1950s, plastics have dramatically changed people’s way of life,
but at the same time, they pose an environmental challenge because of the means used for
their disposal in various parts of the world. Since 1950, 8,300 million metric tons of syn-
thetic polymers has been produced, and 4,900 million metric tons has been discarded
(Figure 6-4). Resistance to degradation—one of the most important properties of many

1
Here, for the purposes of simplicity, the terms “plastics” and “polymers” are used interchange-
ably. IUPAC (the International Union of Pure and Applied Chemistry) defines plastics as “poly-
meric materials that may contain other substances to improve performance and/or reduce costs.”
Plastics are manmade and include both thermoplastics and thermosets, whereas polymers can be
manmade or occur naturally.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

166 New Directions for Chemical Engineering

plastics—is the main cause of their persistence in the environment. Thus, a circular econ-
omy for plastics provides the best opportunity to continue enjoying the benefits of plastics
in everyday life while reducing the environmental impact of their mismanaged disposal.
This circular economy of plastics addresses both aspects of the first strategy listed above
(i.e., both pollution and waste reduction).
The primary goal of the circular economy of plastics is that plastics never become
waste, and instead reenter the economy. There are two main strategies for achieving this
goal: create an effective after-use plastics economy, and drastically reduce the leakage of
plastics into natural systems and other negative externalities (EMF, 2016). Steps to reduce
the pollution and waste of plastics include selecting from and executing various options
that are depicted in typical waste hierarchies (Billiet and Trenor, 2020). The most pre-
ferred option is rethinking and redesigning the product and package. If that option is not
possible, other options, in order of preference, are as follows:

 Reduce the amount of plastic used in the product and package.


 Reuse the product and package.
 Recycle the product and package.
 Use the plastic in the product and/or package as fuel.
 Dispose of the plastic in the product and/or package in a managed landfill.

FIGURE 6-4 Global production, use, and fate of polymer resins, synthetic fibers, and additives,
1950–2019, in millions of metric tons. SOURCE: Geyer (2021).

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Flexible Manufacturing and the Circular Economy 167

The least desirable outcome is disposal of the plastic in a mismanaged way. Re-
placing fossil-derived plastics with biobased plastics (totally or partly) and using renew-
able energy in the production of fossil-derived plastics are some alternative options for
reducing the carbon footprint of plastics. Chemical engineers have an opportunity to apply
quantitative, systems-level thinking to this problem through the application of TEA and
LCA to determine which options optimize emissions reductions while considering other
trade-offs, such as water consumption, cost, and environmental justice. Chemical engi-
neers also have an important role to play in increasing and improving the recycling of
plastics. Opportunities exist in all recycling technologies, such as mechanical recycling;
dissolution recycling; and advanced recycling methods, such as depolymerization, pyrol-
ysis, and gasification (Figure 6-5).

FIGURE 6-5 Processes (mechanical, dissolution, and chemical or advanced recycling processes)
for recycling of plastics, recycling loops, and plastic feedstocks in each recycling process. NOTE:
EoL = end-of-life; PE = polyethylene; PET = polyethylene terephthalate; PP = polypropylene; PS
= polystyrene; WtE = waste-to-energy. SOURCE: Collias et al. (2021).

There are numerous opportunities to improve the quality of plastics produced and
increase the range of plastics that can be mechanically recycled. Technologies that clean
waste plastics and remove surface and bulk contamination are also the focus of chemical
engineering work. Mechanical recycling of mixed polymers leads to the formation of pol-
ymer blends, and the associated technologies for compatibilization are important. Plastics
present an additional challenge in that, unlike metals or glass in recycling, polymers
change their structure. Advanced recycling will help recycle the polymers that mechanical
recycling cannot, and could provide infinite recycling loops. An example of an LCA of
mechanical recycling processes is described by Franklin Associates (2018).
Scaling up of dissolution recycling processes that produce food-grade plastics
from waste plastics presents another opportunity for chemical engineers. Examples of
such processes include Newcycling® by APK AG, PureCycle Technologies (Layman,

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

168 New Directions for Chemical Engineering

2019a,b), and CreaSolv® by CreaCycle GmbH (Maeurer et al., 2012). The main unit op-
erations that chemical engineers can further develop, optimize, scale up, and commercial-
ize are polymer dissolution, extraction, processing, layer separation, filtration, contami-
nant migration, and solvent diffusion and recycling (Pappa et al., 2001; Walker et al.,
2020; Zhao et al., 2018).
Depolymerization of condensation polymers is applicable to polyesters (e.g., PET
and polylactic acid [PLA]) and polyamides (e.g., nylon), among others. Hydrolysis, meth-
anolysis, glycolysis, ammonolysis, aminolysis, and hydrogenation are typical chemical
processes that, depending on the chemical used for the PET chain scission, degrade the
starting polymers to either their respective monomers or oligomers. The product of depol-
ymerization is purified to remove contamination and colorants, and then used to form new
polymers. Catalyst or enzyme development, process optimization and scale-up, purifica-
tion, and polymer processing are key areas with an intense chemical engineering focus.
Examples of LCA for PET recycling are found in Shen et al. (2010) and Singh et al.
(2021).
Pyrolysis is used for polyolefins, multilayer packaging, fiber-reinforced compo-
sites, polyurethanes, and other polymers that are difficult to depolymerize. Pyrolysis takes
place at moderate to high temperatures and produces various hydrocarbons. Chemical en-
gineers play a central role in the design of the reactor and its modeling. Catalytic pyrolysis
(Cocchi et al., 2020; Miandad et al., 2016; Ratnasari et al., 2017; Zero Waste Scotland,
2013) is another technology that presents chemical engineering challenges and opportu-
nities. The objective of this technology is to achieve C–C bond breaking at lower temper-
atures and reaction times relative to thermal pyrolysis, and to produce a higher-volume
and higher-quality liquid fraction. Catalysts that have been explored include FCC (fluid
catalytic cracking) catalyst, spent FCC catalyst, ZSM-5 (Zeolite Socony Mobil-5),
HZSM-5, Cu-Al2O3, zeolites, Fe2O3, MCM-41 (Mobil Composition of Matter No. 41),
coal fly ash, and mixtures of the above (Miandad et al., 2016; Ratnasari et al., 2017). In
addition to thermal and catalytic pyrolysis technology vectors, plasma pyrolysis, micro-
wave-assisted pyrolysis, and hydrocracking are other pathways for converting feedstock
to pyrolysis oil for further conversion. A chemical engineering challenge in pyrolysis is
to achieve process scale-up rather than process scale-out (i.e., scale via parallel units/
modular design), which is common today, including with the advent of electrochemically
driven reactors. This scale-up challenge results from the difficulty of achieving adequate
heat transfer to the plastic (which acts as an insulator) as the volume of the reactor in-
creases and its surface area does not increase proportionally. As a result, the typical annual
capacity of pyrolysis reactors/plants is 5–10 thousand tons, and larger-capacity plants are
constructed with many reactors in parallel rather than the more economical option of using
larger reactors.
Gasification converts plastics to a gaseous mixture of CO and H2 (syngas), CO2,
CH4, and other light hydrocarbons via partial oxidation in the presence of steam and ox-
ygen or air at less than the stoichiometric ratio (Higman and van der Burgt, 2008;
Rezaiyan and Cheremisinoff, 2005).
Incineration can be useful for end-of-life plastics and other waste. In waste-to-
energy (WtE) processes, for example, MSW is combusted, and the heat produced is used

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Flexible Manufacturing and the Circular Economy 169

to make steam for generating electricity or to heat buildings. Besides producing electric-
ity, WtE contributes to reducing the amount of material that would otherwise go to land-
fills and the resulting landfill emissions. In 2018 in the United States, about 12 percent of
the 292 million tons of MSW was burned in WtE plants, generating 13 billion kWh of
electricity in 2019 (EIA, 2020c). The percentage of MSW that feeds WtE installations
ranges from 12 percent in the United States to 74 percent in Japan. Many large landfills
generate electricity from the methane gas that is produced from the decomposition of bi-
omass. Incineration and WtE processes can also potentially release hazardous chemicals
and particles. These chemicals can affect air quality, affecting the neurological, respira-
tory, and reproductive systems of the human body and damaging the environment. Re-
moving these chemicals from the incineration process is an important chemical engineer-
ing challenge.
Other potential solutions to the plastics disposal problem beyond the recycling
technologies discussed above include the following:

 Use of biodegradable plastics and various enzymes and biodegrading organ-


isms for plastics—LCAs are necessary to determine whether in some envi-
ronments, the negative effects of the uncontrolled release of the biodegrada-
tion products into the atmosphere outweigh the benefits of using
biodegradable plastics.
 Closed-loop recycling of polymers synthesized with in-chain functional
groups that act as break points—For example, Häuβler and colleagues (2021)
prepared a redesigned version of polyethylene with renewable polycarbonate
and polyester in-chain functional groups that can be recycled chemically by
solvolysis with a recovery rate exceeding 96 percent.
 Composting and anaerobic digestion—Compositing is a managed process of
controlled decomposition of organic material by aerobic microorganisms pro-
ducing compost (also called humus), CO2, water, and heat. Anaerobic diges-
tion is used primarily to process wet waste material with microorganisms that
break down organic material in the absence of oxygen and produce biogas.

Extension of Useful Life

In addition to reducing environmental impact, there is also an economic incentive


for keeping materials, products, and packaging in use as long as possible. The economic
value of plastic packaging that is lost after a single use is estimated to be between $80–
120 billion, and the environmental cost of plastic packaging is estimated to be $40 bil-
lion—more than the industry’s total profits (McKinsey Sustainability, 2016). In designing
products for longevity, the entire life cycle of the product and the value chain need to be
managed. For example, if a physical unit is needed, design thinking might suggest making
it more durable or making it easy to maintain by using designs that allow critical compo-
nents to be replaced when worn out.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

170 New Directions for Chemical Engineering

Apart from designing for easy disassembly, use of single-material packaging is at


the top of the list of desirable options for a product at the end of life. Multimaterial pack-
aging does not allow for mechanical and other types of recycling because the various
materials cannot easily be separated. Such multimaterial packaging might include blends
of different plastics or multilayer films, with each layer made of a different polymer. Sep-
aration technologies for multimaterial films (e.g., biomimicry-based adhesives between
the materials or layers) and the development of monomaterial solutions that achieve the
same performance as multimaterial solutions are two key areas in which chemical engi-
neers will continue to contribute.

Regeneration of Natural Systems

A biobased economy includes both biobased materials and biobased fuels (fuel
uses are discussed in Chapter 3). Sugars (e.g., glucose) are the typical feedstocks for a
biobased economy, with biomass being the preferred primary source. Biomass is typically
classified as generation 1 (e.g., sugarcane, corn, sugar beet, potato) or generation 2 (e.g.,
crop residues, grass, straw, wood, MSW). Because generation 1 biomass as a sugar source
competes with food uses, generation 2 biomass is a better long-term source of sugars.
Generation 2 biomass contains three natural polymers: cellulose (35–50 percent
of the biomass), hemicellulose (25–30 percent), and lignin (15–30 percent). Cellulose is
a linear polymer composed of (1,4)-d-glucopyranose units linked by β-1,4-glycosidic
bonds. Hemicellulose is a branched heteropolymer composed of hexoses, pentoses, and
uronic acids linked by different bonds. Lignin is a high-molecular-weight, amorphous
polymer composed of aromatic rings of phenyl propane. Various processes are required
to separate biomass into its three main components, and then convert cellulose and hemi-
cellulose to sugars, and lignin to other smaller chemical units. The challenges and oppor-
tunities for chemical engineering in the use of biomass are significant. Technically and
economically successful hydrolysis of biomass and the production of low-cost sugars are
key to enabling the biobased economy. Chemical and enzymatic hydrolytic processes
need to be advanced in this space. Also, valorization of lignin is key to making overall
biomass utilization successful.
Use of biomass for fuels, products, and power has been recognized by the U.S.
Department of Energy (DOE) as a critical component of reducing dependence on volatile
supplies and prices of imported oil (Biddy, 2016). The importance of chemicals in the
economy is disproportionate to the amount of oil used to produce them. For example, only
about 3–4 percent of petroleum is used to make chemicals, whereas the pretax value of
the petrochemical products (e.g., plastics, detergents, paints, adhesives, cosmetics, pesti-
cides) was about $375 billion in 2005—about the same as the pretax value of transporta-
tion fuels, which use about 71 percent of petroleum (Marshall, 2007). The production cost
of biobased chemicals and its comparison with that of the corresponding petroleum-de-
rived chemicals depend on the prices of biomass, petroleum, and natural gas, as well as
production capacity, technology maturity, and other factors. As the current price of natural

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Flexible Manufacturing and the Circular Economy 171

gas is relatively low (compared with the price of petroleum on an energy basis), the cur-
rent production cost of the incumbent chemicals is relatively low relative to that of the
biobased chemicals.
Over the past two decades, biorefineries have been proposed as a way to achieve
favorable economics compared with typical petroleum refineries and downstream pro-
cessing plants. The products of biorefineries can be fuels, chemicals, or both. The Inter-
national Energy Agency (IEA, 2020b; IEA, ICCA, and DECHEMA, 2013) presented a
vision of biorefineries based on eight feedstock platforms:

 Syngas—Biomass gasification yields syngas, which is converted into prod-


ucts using fermentation or other chemical processes.
 Biogas—Anaerobic digestion of high-moisture-content biomass (such as ma-
nure and waste streams from food processing plants) yields primarily me-
thane, which can be used as a chemical feedstock.
 C6 and C5 sugars—C6 sugar streams and, to a lesser extent, C5/C6 mixed
sugar streams can be used in fermentation processes to produce various chem-
icals.
 Plant-based oil—The basis of the oleochemical industry, with a by-product
of glycerin, plant-based oil has been explored as feedstock for the production
of chemicals.
 Algae oil—Algae has a higher productivity than plants because the entire bi-
omass can be used to produce high-value products, such as high-protein food
or feed, pigments, antioxidants, vitamins, and sterols.
 Organic solutions—When fresh wet biomass (such as grass or clover) is de-
watered, an organic solution (nutrient-rich press juice) and a press cake (fi-
ber-rich) are produced, and both can be used to make other chemicals.
 Lignin—Lignin’s native structure suggests that it can be used to produce ar-
omatic molecules; however, only a limited number of products from its de-
rivatives, lignosulfonates and kraft lignin, have been produced so far.
 Pyrolysis oil—Biomass pyrolysis produces pyrolysis oil, which can then be
upgraded to different chemicals.

Biobased plastics are also part of the circular economy of plastics. Currently,
there is a biobased plastic-alternative material (either biobased, biodegradable, or both)
for many fossil-derived conventional plastics, at least at laboratory or pilot scale (Siracusa
and Blanco, 2020). There are three main groups of biobased plastics (Figure 6-6):

 biobased or partially biobased and nonbiodegradable plastics, such as bi-


obased polyethylene (PE; e.g., Braskem’s I’m green™), polypropylene, and
PET and biobased polymers (e.g., polytrimethylene terephthalate [PTT]) (top
left quadrant of Figure 6-6);
 biobased and biodegradable plastics, such as PLA, polyhydroxy alkanoate
(PHA), polybutylene succinate (PBS), and starch-based blends (top right
quadrant of Figure 6-6); and

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

172 New Directions for Chemical Engineering

 biodegradable plastics from fossil resources, such as polybutyrate adipate ter-


ephthalate (PBAT), polyvinyl alcohol (PVOH), and polycaprolactone (PCL)
(bottom right quadrant of Figure 6-6).

Compared with the conventional fossil-derived plastics (bottom left quadrant of


Figure 6-6), biobased plastics can potentially have a lower carbon footprint and additional
end-of-life options, such as composting. The global production of biobased plastics was
2.1 million metric tons in 2020, representing less than 1 percent of total plastics produc-
tion. Biodegradable plastics made up about 58 percent of the biobased plastics volume,
with the remainder comprising biobased and nonbiodegradable plastics (European
Bioplastics, 2021). LCA can be a valuable tool for evaluating the trade-offs between con-
ventional fossil-derived and biobased plastics.

FIGURE 6-6 Qualitative plot of polymers based on their biodegradability (horizontal scale) and
whether they are bio- or fossil-based (vertical axis). NOTE: PA = polyamide; PBAT = polybutyrate
adipate terephthalate; PBS = polybutylene succinate; PCL = polycaprolactone; PE = polyethylene;
PET = polyethylene terephthalate; PHA = polyhydroxy alkanoate; PLA = polylactic acid; PP =
polypropylene; PTT = polytrimethylene terephthalate. SOURCE: European Bioplastics (2021).

The economics of the processes used to make biobased plastics depend heavily
on the stoichiometry of the feedstock conversion to the product, and particularly on the
oxygen content of the feedstock. For example, the theoretical mass yield of PE from sugar
is about 30 percent, and the rest (70 percent) of the sugar mass is lost as water and CO2
(see the oxygen chart in Figure 3-10 in Chapter 3). This process starts with sugar fermen-
tation to ethanol, dehydration of ethanol to ethylene, and polymerization of ethylene to
PE. This bio-PE cannot compete in price with fossil-derived PE because of the stoichi-
ometry of the conversion reaction from carbohydrate to hydrocarbon and the economy of

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Flexible Manufacturing and the Circular Economy 173

scale present in the production of fossil-derived PE. A similar argument holds for the
production of biofuels from sugar or biomass. Note that the capacity of a typical petro-
leum refinery is 10–130 million metric tons per year (measured as the amount of petro-
leum that can be distilled in the atmospheric distillation units), whereas the capacity of an
average ethanol fermentation plant is about 0.25 million metric tons, and that for ethanol
dehydration to ethylene is smaller.
The biobased chemicals that are economically advantageous to produce from sug-
ars are those that have an oxygen content similar to that of sugar or biomass, such as
monoethylene glycol (MEG) and furan-2,5-dicarboxylic acid (FDCA), with mass theo-
retical yields of about 100 percent; acrylic acid (AA), with mass theoretical yield of 80
percent; and polyethylene furanoate (PEF), purified terephthalic acid (PTA), and PET,
with theoretical mass yields exceeding 50 percent. Use of such feedstocks as lignin, CO,
CO2, and vegetable oils will be advantageous to produce chemicals with similar oxygen
contents. Various LCA studies on biobased plastics have been reviewed (Walker and
Rothman, 2020).
It is important for chemical engineers to be involved in the development of the
biobased economy. More specifically, challenges and opportunities for chemical engi-
neers in this domain include

 scalable and economical processes for producing biobased plastics (e.g., PEF
and PBS) and biobased monomers (e.g., biobased terephthalic acid; Collias
et al., 2014);
 new biodegradable plastics for the marine environment;
 scalable and economical technologies for deconstructing lignin to molecules
that can be used as feedstocks for various chemicals, such as
- bacterial, enzymatic, fungal, photocatalytic, hydrogenolysis, pyrolysis,
oxidation via ionic liquids, and hydrolysis deconstruction technologies
(Cao et al., 2018a; Glasser, 2019; Kellett and Collias, 2016; Ragauskas
et al., 2014; Xu et al., 2019; Zakzeski et al., 2010); and
- engineering of lignin to make it more amenable to low-energy chemical
deconstruction, such as by using monolignol ferulate transferase to intro-
duce chemically labile ester linkages into the lignin in poplar trees
(Wilkerson et al., 2014), and the proposed work on use of CRISPR (clus-
tered regularly interspaced short palindromic repeats)–based genome en-
gineering to edit multiple genes simultaneously to optimize biomass pro-
cessing (Chanoca et al., 2019);
 technologies that use CO2 as a feedstock for various chemicals;
 reduction technologies for carbohydrate feedstocks to produce less oxygen-
ated chemicals and hydrocarbons (e.g., dehydration of lactic acid to acrylic
acid [Collias et al., 2018; Godlewski et al., 2014]; see Bozell and Petersen
[2010] for the top 10 chemicals of the bioeconomy and the National Renew-
able Energy Laboratory [Biddy, 2016] for the 12 chemicals with prospects
for near-term deployment);

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

174 New Directions for Chemical Engineering

 technologies that avoid energy-intensive processes, such as steam cracking;


and
 technologies for converting waste plastics to higher-value chemicals or struc-
tures (upcycling), such as PE waste upcycling to long-chain alkylaromatics
using tandem catalytic conversion by Pt supported on γ-alumina (Zhang et
al., 2020b), lubricant and waxes (Celik et al., 2019; Rorrer et al., 2021), diac-
ids (e.g. succinic acid, adipic acid) via oxidation, hydrogen (Uekert et al.,
2020), and lithium-ion battery anodes (Villagomez-Salas et al., 2018).

CHALLENGES AND OPPORTUNITIES

A sustainable future will require a shift to a circular economy in which the end of
life of products is accounted for, incorporating green chemistry and engineering. The
chemical engineering profession emerged in large part to confront the urgent challenges
faced more than a century ago in the then-burgeoning petroleum-refining industry. Petro-
leum is a highly heterogeneous organic resource. The continued drive toward more effi-
cient, environmentally friendly, and cost-effective manufacturing processes will benefit
from a much wider range of available feedstocks for use as building blocks to produce
chemicals and materials. The challenge of feedstock flexibility offers chemical engineers
an opportunity to develop advances in reductive chemistry and processes that will allow
the use of oxygenated feedstocks, such as lignocellulosic biomass. Chemical engineers
also have substantial opportunities develop scaled-out, distributed manufacturing systems
and innovative, large-scale processes that can compete with the conversion of fossil re-
sources. Fundamental research opportunities include the collection of robust thermody-
namic data to improve the modeling of feedstock molecules that include oxygen and other
heteroatoms.
Current challenges in process design include the need for improvements in dis-
tributed manufacturing and process intensification—areas in which the chemical engi-
neering research community can provide intellectual leadership. Collaborations between
academic researchers and industrial practitioners will be important for demonstration at
process scale. In the transition from a linear to a circular economy, specific opportunities
for chemical engineers include redesigning processes and products to reduce or eliminate
pollution (e.g., Shi et al., 2021), developing new ways to reduce and utilize waste, design-
ing products to be used longer, and designing processes and products using sustainable
feedstocks.

Recommendation 6-1: Federal research funding should be directed to both basic and
applied research to advance distributed manufacturing and process intensification,
as well as the innovative technologies, including improved product designs and re-
cycling processes, necessary to transition to a circular economy.

Recommendation 6-2: Researchers in academic and government laboratories and


industry practitioners should form interdisciplinary, cross-sector collaborations fo-

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Flexible Manufacturing and the Circular Economy 175

cused on pilot- and demonstration-scale projects in advanced manufacturing, in-


cluding scaled-down and scaled-out processes; process intensification; and the tran-
sition from fossil-based organic feedstocks and virgin-extracted inorganic feedstocks
to new, more sustainable feedstocks for chemical and materials manufacturing.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

7
Novel and Improved Materials for the 21st Century

 Chemical engineers have a critical role to play in the development of new ma-
terials and materials processes from the molecular to the macroscopic scale.
 Chemical engineers’ integration of theory, modeling, simulation, experiment,
and machine learning is accelerating the discovery, design, and innovation of
new materials and new materials processes.

Chemical engineers are deeply involved in the design, synthesis, processing,


manufacturing, and ultimately disposal of materials of all kinds, and the connections be-
tween the fields of materials science and chemical engineering are numerous and diffuse.
A taxonomy might suggest that materials scientists and engineers are concerned mainly
with materials structure, properties, and performance, while a chemical engineer would
likely focus less on performance and more on materials processing, but there are as many
exceptions to that generalization as there are examples. In industry, the distinction is usu-
ally unmeasurable. Full coverage of the myriad dimensions of chemical engineering’s role
in the materials economy would require its own report, but this chapter highlights several
important aspects and key opportunities for the future. For example, materials research in
academic chemical engineering departments will include work in polymer science, rheol-
ogy, catalysis, biomaterials, nanomaterials, electronic materials, self-assembly, and soft
matter; several of these subjects have been discussed in earlier, application-focused chap-
ters of this report.
Chemical engineers have been responsible for many advances in materials design
and development. An example is the reverse osmosis membrane for water desalination
made of cellulose acetate and developed by chemical engineers in the 1950s (Cohen and
Glater, 2010). Likewise, many of the polymeric materials used in regenerative engineer-
ing and drug delivery have emerged over decades from the laboratories of chemical engi-
neers. Other contributions include new catalysts and zeolites (Ahmed, 2007), Gore-Tex®,
and the automobile catalytic convertor.
Chemical engineers have played a key role in materials processing and system
design that enable plants to make useful amounts of materials safely. They also are devel-
oping solutions for the consequences of decades of plastic generation, which has left the
world awash in plastic waste. And they bring rigorous life-cycle assessment (LCA) to the
processes that produce materials, as examined in greater detail in Chapter 6.
The economic and environmental burden of electronic waste rivals that of plas-
tics—indeed, the fastest-growing segment of the global solid waste stream is electronics
and electrical waste (Kaya, 2016)—and poses a challenge even more complicated to ad-
dress. Consumer products such as mobile phones and personal computers (PCs) contain

176

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Novel and Improved Materials for the 21st Century 177

more than 1,000 different chemical products, and more than 260 million PCs were sold in
2019 alone (Gartner, 2020). Gold, silver, copper, and palladium are the most valuable
metals found in the waste stream, but they are challenging to recover. To gain perspective
on the magnitude of the problem, consider that an economically viable gold mine yields
5 g of gold per ton of ore, while 1 ton of discarded cell phones can yield as much as 150
g of gold, 100 kg of copper, and 3 kg of silver (Nimpuno and Scruggs, 2011).
This chapter explores four areas of materials research, design, and production in
which chemical engineers are particularly active: polymer science and engineering, com-
plex fluids and soft matter, biomaterials, and electronic materials. These four areas are
but a subset of materials work done by chemical engineers, and a deeper survey would
yield its own report. Other areas that the committee did not explore include advanced
(non–petroleum-based) performance fluids, such as lubricants or high-temperature heat
transfer fluids; mixture formulation in general (beyond complex fluids); materials for sep-
arations beyond water purification (discussed in Chapter 4); and construction materials
such as concrete and asphalt,1 an area that could lead to new collaborations with civil
engineers. In addition, there may be a role for chemical engineering in quantum materials
and quantum information technologies.2 Catalytic materials and applications in the energy
transition are discussed in Chapter 3.

POLYMER SCIENCE AND ENGINEERING

In the 100 years since the publication of Hermann Staudinger’s landmark paper
Über Polymerisation (1920)—which marked the beginning of an ability to design plastics
with infinitely tunable properties, moldability/processability, and remarkably low ex-
pense—polymers have infiltrated every aspect of people’s lives. The purification of water,
the preservation of food, the clothing people wear, the diapers on babies, the vehicles used
for transport, the components of computers, and the delivery of medicines to the body all
contribute to reliance on a global polymer industry whose products have grown in volume
to more than 350 million metric tons per year since 1950 (Figure 7-1). Chemical engineers
have been central to the development of this industry from its inception because of their
integrated understanding of chemical synthesis and catalysis, thermodynamics,
transport/rheology, and process/systems design. Indeed, the understanding of polymer
rheology and design of processing equipment to handle highly viscous polymer melts
represents a triumph of chemical engineering. Over this 100-year timeframe, the field of
polymer science, including the contributions of chemical engineering, has become in-
creasingly molecular in focus, leading to the ability to create new materials that integrate
new polymer chemistries, topologies, and assemblies.

1
The National Academies’ Committee on Repurposing Plastics Waste in Infrastructure will
explore the effectiveness and utility of plastic waste in asphalt mixes and other materials used in
infrastructure, which may present opportunities for chemical engineers to collaborate with other
disciplines.
2
The National Academies’ Committee on Identifying Opportunities at the Interface of Chem-
istry and Quantum Information Science is currently examining the research needed to make pro-
gress in this area.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

178 New Directions for Chemical Engineering

FIGURE 7-1 Increase in primary plastic production between 1950 and 2015. The figure shows
end uses including textiles, industrial machinery, consumer and institutional products, electri-
cal/electronic, building and construction, transportation, and packaging. SOURCE: Geyer et al.
(2017).

Chemical engineers’ backgrounds in polymer chemistry, thermodynamics, and


kinetics have been especially well suited to designing and understanding the interplay
between thermodynamic and kinetic driving forces acting through the atomic- (mono-
mer), nanoscopic- (sequence, chain shape), and mesoscopic-length scales that lead to self-
assembly—ultimately determining the macroscopic functional properties of a polymeric
material. Block copolymers are one example of such hierarchically assembled materials.
The ability to control the properties of the individual blocks to induce hierarchical order
and control polymer macroscopic properties has enabled such applications as thermo-
plastic elastomers, semiconductors, nanolithographic masks and patterns, and ion-con-
ducting solid-state electrolytes. Directed self-assembly of block copolymers led to the
ability to make nanometer-scale patterns with high fidelity across macroscopic areas to
rival stereolithography. New opportunities arise as one considers new assemblies in which
the building blocks are more complex, are potentially dynamic, and change under external
stimuli. In this way, both the molecular-scale structure and the macroscopic properties
could adapt with time and situation to lead to truly active and responsive materials. Chem-
ical engineers are well positioned to address the molecular engineering of hierarchical
assembled polymers, which will rely on the design of chain conformation, mesoscopic
structure, and macroscopic function in an integrated manner. For example, noncovalent
interactions (e.g., metal-ligand, electrostatic, hydrophobic, hydrogen-bonding interac-
tions) lead to dynamic, transient bonding with multiple handles to control multiple length
and time scales simultaneously.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Novel and Improved Materials for the 21st Century 179

As laid out in the National Science Foundation workshop report Frontiers of


Polymer Science and Engineering, making the next leap forward in the design of func-
tional materials will require understanding, predicting, and utilizing molecular building
blocks, sequence, conformation, and chirality to direct mesoscopic structure on desired
time and energy scales to control macroscopic polymer properties (Bates et al., 2017).
Further, while the ability to predict structure across length scales of six orders of magni-
tude has improved dramatically, it is now necessary to learn how to predict properties and
behavior in order to guide material design. Within the next 20 years, the capability to
design polymers from the bottom up for target applications will likely advance to a point
at which the desired properties and behavior of a polymer can be specified and, using a
combination of multiscale simulation and artificial intelligence (AI), the building blocks
and the processing strategy can be designed to make it.
The ability to rationally embed information and function in synthetic macromol-
ecules by controlling monomer sequence opens the door to polymers with the sophistica-
tion of biological molecules (Figure 7-2) and potentially with the ability to meet “the
grand challenge of precision control over polymer structure and function” (Bates et al.,
2017). Recently, strategies for controlling monomer sequence have evolved to enable syn-
thesis of sequence-controlled synthetic polymer chains at gram scale (Lutz et al., 2013),
allowing for molecular weight, architecture, and chain end control through highly effi-
cient coupling reactions and step-wise automated synthesis and purification. Two funda-
mental questions that need to be answered if the potential of these advances is to be real-
ized: What monomeric sequence should be incorporated? and What level of control is
necessary to yield new structures and properties? (Bates et al., 2017). Now that almost
anything can be made, what should be made? Even a single 80-mer chain composed of
two different repeat units permits 6 x 1023 distinct sequences. In this Avogadro-scale de-
sign space, predictive tools that allow selection of specific targets of assembly become
essential, as a design driven by intuition is no longer useful.

FIGURE 7-2 Increasing complexity in the synthesis of polymers ranging from copolymers to bi-
ological polymers. SOURCE: Modified from Rosales et al. (2013).

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

180 New Directions for Chemical Engineering

While the future of designed, molecularly engineered polymers-on-demand is ex-


citing, responsible care of the environment demands a paradigm shift to include life-cycle
considerations in the design imperatives. The combination of ultralow cost, scalability,
and utility has led to the widespread use of plastics in every aspect of modern life. As a
result, society has accumulated a huge debt of polymers that will never degrade. Global
production of plastics has risen exponentially in the 70 years since their first mass pro-
duction, from less than 2 Mt in 1950 to 359 Mt in 2018 (PlasticsEurope, 2019), and pro-
duction is projected to double again in the decade ahead (Jambeck et al., 2015). Plastics
are found in industries as diverse as food, medical devices, electronics, transportation, and
construction. Packaging is one of the most common applications for plastics, and most
packages are used once and then discarded after a life cycle of much less than 1 year
(Teuten et al., 2009).
Mismanagement of plastic waste has contributed to persistent, visible environ-
mental pollution (Rahimi and García, 2017), the formation and widespread dispersal of
microplastic fragments, and leaching of contaminants that impact the health of ecosystems
(UNEP, 2021). Only about 9 percent of plastic waste is recycled, and almost all is con-
verted via mechanical recycling to lower-value materials (downcycling; Geyer et al.,
2017). The combination of widespread use of plastics and the failure to deal with their
end-of-life phase has brought the plastic-waste problem to the forefront of the world’s
attention. The extreme aspect ratio and entanglement of polymer molecules give plastics
the unique properties that led to their widespread use over the last century, but these same
features—combined with the breadth of designer chemistries (and designer properties) of
commodity plastics—present significant challenges to the recycling or upcycling of plas-
tics to monomers or high-value products (see Chapter 6).
The development of feasible depolymerization processes requires fundamental
understanding of catalysis, polymer chemistry, and the rapidly changing melt rheology
during processing, and all of this knowledge is in the domain of chemical engineering.
Similarly, chemical engineers are well poised to develop alternative, scalable plastics with
properties equal to or better than those of current commodity plastics and with a greener
life cycle.

COMPLEX FLUIDS AND SOFT MATTER

Traditionally, chemical engineers have designed and formulated functional fluids


that are exploited in fields ranging from food, personal care, and pharmaceuticals to active
braking fluids and bulletproof vests. These so-called complex fluids often include func-
tional structures that form spontaneously by self-assembly or under external fields by di-
rected assembly. The structures can represent equilibrium states, dynamical steady states,
or kinetically trapped states. Such formulations have been developed to sequester spar-
ingly soluble or delicate molecules, to modify rheology, and to control phase behavior
(Larson, 1999). In the past two decades, chemical engineers have been at the forefront of
developing scalable strategies for advanced functional material assembly in complex flu-
ids and soft matter. The advent of nanotechnology lent urgency to this area, as the promise
of advanced functionality relied on the development of efficient and scalable schemes for

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Novel and Improved Materials for the 21st Century 181

incorporating microscopic and colloidal building blocks into larger structures (Box 7-1).
Although this area of research is the modern version of colloid science and chemistry, its
academic center in the United States now rests firmly in chemical engineering depart-
ments.

BOX 7-1
Smart Materials

Passive daytime radiative cooling (PDRC) surfaces (see Figure 7-1-1) have been proposed
as coatings for buildings to reduce the indoor temperature to below ambient, thereby reducing
dependence on air conditioning. PDRC surfaces rely on reflection from structures embedded in
the coating, and radiation in the long-wave infrared window of the spectrum to send radiated
energy out of the Earth’s atmosphere (Catalanotti et al., 1975). These surfaces typically have
complex designs incorporating elements that can include alternating layers of materials (Raman
et al., 2014; Rephaeli et al., 2013), emitter particles, and carefully placed metal mirrors (Chae
et al., 2021; Zhai et al., 2017). However recent advances have revealed a new path to their de-
sign: an effective PDRC surface was generated from a porous polymer film (polymethyl meth-
acrylate, poly[vinylidenefluoride-co-hexafluoropropene]) that exploited broadband scattering
from the bubbles for reflection of sunlight and the emission of the polymer itself for radiation
(Mandal et al., 2018; Wang et al., 2021c). The design of such scalable functional structures is
an area in which chemical engineers can contribute in new ways to functional materials for
energy management.

FIGURE 7-1-1 Passive daytime radiative cooling surfaces. (a) High reflection and emission
enable a net radiative loss from the surface to generate cooling below ambient temperatures. (b)
A scanning electron microscope image of a photonic radiative cooler material shows its seven
layers. (c) A schematic of hybrid material that absorbs all incident infrared irradiance. (d) Mi-
crographs show the top and cross-section of a porous film. SOURCES: Mandal et al. (2018);
Raman et al. (2014); Zhai et al. (2017).

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

182 New Directions for Chemical Engineering

Surfactants

Surface active agents (surfactants) are found in nearly every household product
and pharmaceutical formulation and are used industrially in processes ranging from emul-
sion polymerization to enhanced oil recovery. Historically, surfactants have been synthe-
sized from petroleum feedstocks, although considerable progress has been made in the
synthesis of surfactants from plant-based feedstocks. For example, nonionic surfactants
made with sugar-based glucoside hydrophilic moieties instead of petroleum-sourced eth-
ylene oxide groups have the advantage of coming from a renewable resource. In the same
spirit, plant-based oils can be used to synthesize the hydrophobic portions of surfactant
molecules. Catalytic conversion of biomass can create surfactant molecules with proper-
ties (e.g., tolerance in hard water) superior to those of conventional materials (Park et al.,
2016).
Surfactants can be used at low concentrations to modify and control the surface
tension and other properties of interfaces. Chemical engineers have made leading contri-
butions to understanding of the statics and dynamics of far-from-equilibrium soft materi-
als, including multiphase fluid systems that contain dispersed droplets or bubbles, surfac-
tants, and other adsorbed materials. Examples include droplets in emulsions, bubbles in
foams, and dispersions of one polymer in another (a polymer composite). These systems
are typically dominated by complex dynamics at the interfaces. For example, surface-
active molecules or materials self-assemble by adsorption on the surfaces of droplets and
bubbles. These surfactant monolayers generate rich stress conditions that alter the effec-
tive stress of composite systems and can determine the stability of the dispersions to a
dynamic perturbation. This perturbation can occur under processing or aging, and can
determine the behavior of emulsions, foams, and composites commonly exploited in
fields as diverse as personal care, pharmaceuticals, foods, and the design of tires.
The response of the system to a perturbation depends on the composition of the
interface, which evolves over time. Understanding of the dynamics of surfactants and ad-
sorbing species is built on insights from chemical engineering into mass transport and
thermodynamics. This transport often occurs in the presence of flow and deformation of
the interface and can depend strongly on flow conditions. Understanding of these stress
conditions comes from chemical engineering’s long history of studying the fluid mechan-
ics of multiphase flows with nonideal complex stresses (Scriven, 1960). Chemical engi-
neers have addressed these issues by understanding droplet and bubble dynamics and have
developed paradigms with which to understand the nonlinear dependence of interface mo-
bility and surfactant concentration (Stebe et al., 1991) and to reveal the modes of droplet
extension, deformation, and breakup (Stone and Leal, 1989). In extensional flows, drops
break up to form tiny droplets with diameters far smaller than that of the parent drop
(Taylor, 1934). Chemical engineers have revealed how the presence and distribution of
surfactant dictate this occurrence (Anna et al., 2003; Eggleton et al., 2001; Pozrikidis,
1997). These are classic examples in which far-from-equilibrium effects occurring in self-
assembled structures determine the dynamics, stability, and processing conditions of soft-
matter systems.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Novel and Improved Materials for the 21st Century 183

The self- and directed assembly of materials on interfaces remains an exciting


field. Interfaces are inherently open systems in contact with two bulk phases, and are sites
for accumulation of (denatured) proteins, macromolecules, lipids, and particles, all of
which can self-assemble, interact, form structures, and influence dispersion behavior.
Chemical engineers continue to design interfacial probes to understand interface dynam-
ics (Crocker et al., 2000), including those that exploit Brownian motion in the interface
(Squires and Mason, 2009) and those driven externally (Reynaert et al., 2008). Chemical
engineers design formulations for delivery to interfaces, with broad impacts that include
the design of lung surfactant replacements for premature infants (Alonso et al., 2004) and
of surfactant formulations for oil spills (Owoseni et al., 2014). Many of the concepts de-
veloped in these inquiries are now used to study materials assembly at complex fluid in-
terfaces (Kaz et al., 2012) for functional materials and sensing applications (Sivakumar et
al., 2009).
At higher concentrations in aqueous solutions, surfactant molecules self-assemble
into a variety of micellar aggregates, and at even higher concentrations form various liquid
crystal phases. The hydrophobic core of a micelle can be used to solubilize a hydrophobic
molecule, and micellar solutions are the base of many formulations for water-insoluble
drugs. Under some conditions, micelles can grow into long, flexible cylinders that entan-
gle and give rise to highly viscous and viscoelastic solutions that are useful, for example,
as fracking fluids for oil fields (Samuel et al., 2000).
The combination of a hydrophilic material (water) and a hydrophobic material
(oil) with a surfactant in most cases creates an emulsion that is thermodynamically unsta-
ble. Appropriate mixing and shear, together with use of the right kind and concentration
of surfactant, can produce emulsions that are long-lived or easily reemulsified by the con-
sumer. Such emulsions, in which the dispersed drops are micron sized or larger, are com-
mon in consumer products and the food industry. Their structure may be water droplets
in oil (W/O) or oil in water (O/W). It is possible as well to produce multiple emulsions,
such as a water drop in an oil drop within another water drop (W/O/W). Additional pro-
cessing can drive the droplet size smaller and create nanoemulsions, in which droplets are
as small as 10 nm (Helgeson, 2016). Multiple nanoemulsions can also be made, and
nanoemulsions are very attractive as drug delivery vehicles (Zhang et al., 2018a). Finally,
under some circumstances, water, oil, and a surfactant can form a thermodynamically
stable solution called a microemulsion, which contains microdomains of oil and water. It
is an unfortunate nomenclature, but the characteristic domain size in a microemulsion is
smaller than that of a nanoemulsion.
The formulation and optimization of complex fluids draws on core chemical en-
gineering concepts from thermodynamics, transport, and kinetics while recognizing the
key role of molecular forces. Chemical engineers are likely to continue to be at the fore-
front of this work.

Nanoparticles

Among the most exciting developments in soft matter in the last 20 years is the
design, synthesis, and assembly of nanoparticles—colloidal particles ranging in size from

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

184 New Directions for Chemical Engineering

one to a few thousand nanometers. Just as whole new classes of polymer architectures can
now be designed and synthesized through molecular engineering principles and made into
complex functional materials, nanoparticles represent a new class of building block with
tremendous potential for functional materials. The size of nanoparticles means they are
controlled by the same laws of statistical thermodynamics that control polymers and com-
plex fluids. Beyond the micron-sized polystyrene (PS), polymethylmethacrylate
(PMMA), or silica colloidal particles long studied by chemical engineers, today’s nano-
particles can be made from a variety of materials (e.g., gold, silver, CdTe/S/Se, PbS/Se).
Nanoparticles such as these are grown in solution, often as faceted nanocrystals with pol-
yhedral shape. The size and shape of nanoparticles can be controlled using various tech-
niques, yielding hundreds of different possible shapes with nearly monodisperse size dis-
tributions. Lithographic techniques now make possible essentially any particle shape. By
combining shape with anisotropic interparticle interactions, practically any nanoparticle
building block is possible. Spheroidal particles of PS or PMMA can be coated anisotrop-
ically with gold or platinum, or made of multiple materials, to form patchy particles called
Janus particles. A Janus particle—where the two halves of the particle have different in-
teractions with the solution—is the particle equivalent of a block copolymer. Particle an-
alogs of triblock copolymers are also possible, as are particle analogs of star and other
copolymers. These patchy particles are conferred valency by the surface pattern resulting
from the synthesis and subsequent coating or functionalization (Glotzer and Solomon,
2007). This valency, combined with particle rigidity, gives rise to unexpected self-assem-
bled structures, many of which are isostructural to atomic and molecular crystals but on
larger length scale (Figure 7-3).

FIGURE 7-3 Examples of patchy particles made possible by combining nanoparticle shape with
interaction anisotropy through surface patterning obtained via, e.g., grafted ligands, complemen-
tary DNA, or gold coating. Each row represents a homologous series of patchy particles resulting
from changing only one alchemical variable (from top to bottom: Janus balance, aspect ratio, fac-
eting, pattern discreteness, branching). SOURCE: Adapted from Glotzer and Solomon (2007).

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Novel and Improved Materials for the 21st Century 185

In the near future, it will be possible to make nanoparticles of any size and shape,
out of almost any combination of atomic elements or in a core/shell structure, and func-
tionalized with any type of surface ligands—including alkanes, dendrimers, DNA, and
proteins—into any arbitrary surface pattern. These patchy particles are next-generation
building blocks for self-assembly into complex and functional structures. For example,
DNA-functionalized noble metal nanoparticles are programmable atom elements that pro-
vide a powerful path to a wide range of self-assembled colloidal crystal structures (Kim
et al., 2006; Laramy et al., 2019; Tian et al., 2020b). Mixtures of particles functionalized
with complementary DNA linkers can form colloidal cocrystals, in principle, of arbitrary
complexity limited only by the shapes of the particles and human imagination. Even one
type of particle with self-complementary linkers can, because of its anisotropic shape,
self-assemble into complex structures, such as a colloidal clathrate crystal with more than
100 particles in its unit cell (Lin et al., 2017). The interplay between enthalpy and entropy
and between thermodynamics and kinetics is central to assembly engineering of new ma-
terials from nanoparticles and falls squarely within the expertise of chemical engineering.
The same question posed above for block copolymers—Now that we can make anything,
what should we make?—applies here as well. Theory and computer simulation are essen-
tial for finding the “sweet spots” in the vast design spaces available for patchy particles.
Data science, and in particular deep learning with neural networks, has an important role
to play in nanoparticle design for self-assembly. Alchemical potentials (van Anders et al.,
2015) and other methods (Sherman et al., 2020) will enable inverse design of nanoparti-
cles for targeted applications.
Beyond the self-assembly of complex structures, the next decade will see in-
creased effort to understand, design, and engineer the thermodynamics and kinetics of
assembly processes. Engineering assembly processes, or “assembly engineering,” will
make possible functional materials, reconfigurable materials, materials that rely on meta-
stable structures, and soft metamaterials. Assembly engineering will also make possible
materials inversely designed to have unique combinations of properties not typically
found together, enabling wholly new classes of materials for a wide range of applications.
Assembly engineering may also involve directed assembly—for example, using
external electromagnetic fields that dynamically rearrange suspended colloids to modify
suspension rheology (as used in active brakes) and optics (as used in electronic paper e-
ink). Sometimes, the field is one that comes from the complex fluid itself. The most ob-
vious of such fields is surface tension; capillarity lies at the heart of some of the most
widely adopted schemes for organizing colloidal building blocks. Nanoparticles are often
spread and compressed on air–water interfaces of Langmuir troughs to form self-assem-
bled structures trapped at the air–water interface. Nano- and microparticles are commonly
deposited on solid surfaces using capillary assembly methods that rely on the collection
and deposition of the building blocks near the three-phase contact line. So great are sur-
face tension and the change in energy when particles adsorb that adsorbed particles can
become trapped on the interface. Pickering emulsions are droplets stabilized by kinetically
trapped monolayers of particles at their interfaces. Bijels are their bicontinuous counter-
parts, stabilized by jammed particle layers. The structures formed using surface tension

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

186 New Directions for Chemical Engineering

can be exploited directly, can be polymerized, or can serve as templates for other ad-
vanced materials. For example, polymerized bijels have been explored as hollow fiber
membranes, and metallized bagels have been explored as high-surface-area electrodes.
Janus and other patchy particles can be made into “active” particles that are self-
propelled. One approach is to coat half the particle with a material (e.g., palladium) so
that the particle is propelled via catalytic activity when the material encounters a “fuel”
(e.g., hydrogen peroxide). Another approach is to drive nanoparticles with external mag-
netic or electric fields (Han et al., 2018). Collections of active particles can produce com-
plex emergent behavior far from equilibrium, including swarming and motility-induced
phase separation, even in the absence of interparticle attractive interactions (Marchetti et
al., 2013; Takatori and Brady, 2015). The field of active matter is a burgeoning one that,
as it relies on nonequilibrium thermodynamics, falls naturally within the portfolio of
chemical engineering research. By combining assembly engineering with active nanopar-
ticles, chemical engineers are well positioned to create novel materials and material ma-
chines with robotic function at the nanoscale.

BIOMATERIALS

The design of biomaterials has seen strong growth in chemical engineering as a


large number of advances have led to clinical translation that leverages degradable and
biologically derived polymer systems. Materials design has had an influence in areas
ranging from controlled localized release systems to small interfering RNA (siRNA) de-
livery, and a significant impact on the advancement of medicine and improvement in pa-
tient health. Key areas for advances in the biomaterials field include regenerative engi-
neering, wound healing, systemic and localized delivery of nucleic acids, and the delivery
to and detection and imaging of regions of the body that present unique barriers or oppor-
tunities.
The biomaterials field can also provide solutions in areas beyond human health.
Applications with unique promise include areas that support plant and animal science, as
well as the design of plant- or organic-based materials systems that could enable biologi-
cally derived or biomimetic polymers. This section focuses primarily on health-related
applications, with a brief discussion of other areas in which biomaterials will be important.
In addition the recent National Academies report Frontiers of Materials Research: A De-
cadal Survey (NASEM, 2019d) includes an extensive discussion of research opportunities
in biomaterials.

Materials for Regenerative Engineering

The need for methods for generation or repair of tissues and organs is compelling.
The ability to create materials systems that can be tuned to mimic cellular environments,
including the extracellular matrix of tissues and organs, and the biological cues they pre-
sent opens up the pathway to deriving organs and repairing tissues. Wound healing; the
regeneration of bone and cartilage; the replacement of key organs, such as skin or liver;
and the repair of tissues, such as cardiac tissue, are just some of the applications enabled

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Novel and Improved Materials for the 21st Century 187

by the generation of materials matrices and components that can be combined to support
appropriate living cell types. Key to the function of biomaterials for these applications is
the ability to provide settings with the appropriate physical, chemical, and mechanical
cues to enable single- or multicellular systems to arrange, order, and structure into orga-
nized tissues. The array of chemistries available for generating tissue scaffolds is vast;
however, there are overarching requirements for low material toxicity, minimal immuno-
genicity, and the ability to match mechanical properties of the original organs or tissues.
In many cases, moreover, the biomaterial needs to be capable of undergoing controlled
degradation as native tissue is evolved to replace the synthetic material scaffold. The com-
plexity of the biological tissues that require replication introduces additional constraints
on the materials systems that can be used for these applications.
Polymeric materials, including a number of polymer systems that form hydrogels,
have played a critical role in the design and development of biomaterials. These advances
have been made possible by the range of synthetic backbones available; the ability to
“design in” degradability based on hydrolytic susceptibility, pH, or the presence of prote-
ase-susceptible bonds; and the ease of functionalization of polymeric constructs—all of
which make them excellent scaffolds for presenting a range of different proteins, peptides,
and other molecular cues on their surfaces. Along with polymers, a range of mineral com-
posite materials—including hydroxyapatite; porous silica; and other engineered materials,
such as carbon nanomaterials and inorganic or hybrid materials systems—have provided
biophysical cues, such as mechanical stiffness, and sources for biomineralization of hard
tissues, such as bone.
Key advances include the development of dynamic synthetic hydrogels whose
chemistry can be reversibly activated using light or physical or chemical stimuli to induce
changes from one state to another. The more traditional chemically cross-linked hydrogel
network is static, which prevents these networks from undergoing remodeling or support-
ing different tissue development or growth stages. However, more recent capabilities have
allowed the incorporation of photoresponsive groups that enable 3-D patterning of surface
functional groups. Such groups can act as cellular cues, develop reversible cross-linking
to modify mechanical stiffness, or enable materials that can reversibly exhibit different
mechanical states (Rosales and Anseth, 2016). The use of dynamic chemical bonds and
secondary interactions could enable such changes in biological state as hypoxia, cell state,
or the presence of certain signaling molecules to trigger changes in a way that enables
cycling (Figure 7-4).
More recent efforts have also been directed toward engineering materials that can
incorporate and support multicellular systems—essentially allowing for the generation of
2-D and, more particularly, 3-D cellular cocultures supported by materials that enable
cell–cell interactions like those in biological systems. To a large extent, work has focused
on direct replacement or replication of tissues or organs; however, there has also been an
interest in engineered multicellular living systems (M-CELS; Kamm et al., 2018), which
include in vitro cellular cultures that can exhibit additional, unique, or modified functions,
such as on-demand production of a biochemical upon sensing of a trigger molecule, or
actuation of cellular strips upon a given electrical or chemical stimulus. These concepts

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

188 New Directions for Chemical Engineering

require the design of materials systems that can provide not only physical support but also
a matrix that will enable cell–cell configurations that best enable function.
Demands for multicellular systems, whether for synthetic organs or living ma-
chines, include the need to achieve larger and more complex materials systems that can
support angiogenesis and the development of a vascular system that can provide oxygen
and nutrients to cells while in culture. For these reasons, some of the key challenges in
this area include supporting endothelial and smooth muscle cells while promoting the
sprouting and growth of stable vessels; degradable materials systems more readily accom-
modate a growing vascular system. Organization of cells of different types is also required
for sufficient function; biomaterials that can be patterned, aligned, or otherwise manipu-
lated to organize cells selectively are of great interest for these applications.
Additive manufacturing, or 3-D printing, is an important tool for biomaterials
synthesis. Biological inks, or bio-inks, which involve the integration of cells into biolog-
ically compatible materials using water-based solutions or suspensions, can be used with
additive manufacturing methods to create complex patterned structures containing spe-
cific cells in precise arrangements, as with organs and functional tissues (Murphy and
Atala, 2014). Because different tissues exhibit quite different mechanical and chemical
properties, a range of different kinds of bio-inks are needed to address these various needs.
Furthermore, needs for additional materials include ways to generate cellular suspensions
in droplet form while protecting cells from extreme shear forces during printing. The ma-
terials used need to have sufficient cohesion to maintain shape and enable adhesion to the
scaffold. Ultimately, the softer scaffolds typically used may also suffer from mechanical
failure during assembly because of the weight of the printed scaffold, and this balance
between mechanical support and cell compatibility sometimes leads to trade-offs (De-
cante et al., 2021). It is anticipated that the development of new and more readily manip-
ulated materials systems for bio-inks that address these issues will be a focus of chemical
and materials engineers.

FIGURE 7-4 Biological extracellular matrix and synthetic strategies involving reversible chemis-
tries. (a) The native extracellular matrix (ECM) is a heterogeneous fibrillar network that provides
biochemical and physical cues to cells. (b) Synthetic hydrogels are traditionally static, polymeric
networks (middle panel). Dynamic hydrogels can capture aspects of the native ECM by temporally
controlling ligand presentation (left panel) or reversibly cycling through changes in mechanics
(right panel). SOURCE: Rosales and Anseth (2016).

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Novel and Improved Materials for the 21st Century 189

Additional needed developments include the ability to incorporate vasculature


into these systems—a key issue requiring more than precision cellular placement. While
cellular systems can be printed in minutes, it can take days for precursor or stem cells to
develop appropriate stable vasculature. It is therefore difficult to incorporate blood vessels
directly into such tissues, and patterning of endothelial cells is insufficient for creating
functional systems. Finally, a greater ability to mimic the native extracellular matrix en-
vironment is necessary to achieve systems that can develop and evolve to generate inte-
grated organs. Although it is possible to imbed growth factors and other proteins into
existing matrices, it would be preferable to pattern them by printing at higher resolutions
(Ng et al., 2019).
It is important to note the importance of the above techniques not only for tissue
regeneration but also for organ-on-chip applications, which replicate organ function using
microfluidic devices. Such applications are discussed in Chapter 5.

Materials for Drug Delivery

A second large and significant area of growth for biomaterials is the generation
of materials systems that can deliver drugs to the body with control over the timing of
release; the release location; and the targeting of specific tissues, cells, or organs (see also
Chapter 5). The past few decades have led to a detailed understanding of the chemistries
of biodegradable materials capable of achieving desired time-release profiles and of the
thermodynamics that control self-assembled materials, such as block copolymers and lip-
osomes. A range of accomplishments in both more traditional and new materials systems
enable drugs to be encapsulated effectively and to retain their efficacy while decreasing
undesirable side effects and cellular or organ-level toxicity.
A large array of functional polymers, lipids, silica, and inorganic hybrid systems
have been developed that enable different modes of preprogrammed delivery or delivery
on demand in response to specific stimuli. Recent developments have led to therapies that
involve the administration of biologic molecules, ranging from proteins and peptides to
nucleic acids, such as siRNA and messenger RNA (mRNA). Although many advances
have been achieved in the field of controlled drug delivery, several key challenges and
opportunities remain following the discovery of new therapeutic approaches and the in-
troduction of new experimental and computational tools to aid in the understanding of
these materials.

Systemic Delivery and Nanomedicine

The development of nanoparticle systems—which include polymeric and liposo-


mal nanoparticles, as well as inorganic and hybrid nanomaterials—offers another oppor-
tunity for chemical engineers to make an impact in biomaterials. Applications of nano-
particles are promising, particularly for cancer treatment, for which nanoparticles are
believed to have an advantage in sequestering toxic drugs from the rest of the body and
targeting tumors. Unfortunately, although safety and efficacy have improved and cancer

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

190 New Directions for Chemical Engineering

nanomedicines are currently in clinical trials, the delivery of nanoparticles in patients is


complex and presents several challenges.
The surface chemistry, shape, size, and mechanical properties of nanomaterials
affect their circulation half-life within the blood stream, their ability to access tumors by
crossing the endothelial barriers that make up the vascular wall, and their downstream
release characteristics as they progress from circulation to accumulation in the tissue of
the target tumor. Nanomaterials can be designed to present a range of different surface
ligands, including synthetic and native biomolecules, peptides, and proteins, such as an-
tibodies that exhibit strong affinities for surface-bound proteins or other biomarkers on
the surfaces of target cells, such as tumor cells. With more detailed investigations of the
impact of physical and chemical features and nanoparticle composition, the field is rec-
ognizing that these characteristics can yield significant differences in nanoparticle
transport and distribution in the body. This increased understanding will guide how
nanocarriers are deployed against a targeted disease such as cancer.
Nascent opportunities in the cancer field include nanoparticles that can contain
combination therapies addressing two or more vulnerabilities found to be synergistic in
each tumor type. Of particular importance is the ability not only to encapsulate dual ther-
apies in a singular particle system, which can present unique materials challenges, but
also to control when these therapeutics are released, as in some cases, synergy depends
on the timing and staging of release. An advantage of nanoparticle systems is the ability
to tune these release profiles and ensure that target cells receive a sufficient dose of each
drug.

Nanomedicine beyond Cancer

Although cancer is by far the disease setting most studied in nanomedicine,


nanocarriers have also been used in several other biomedical applications. Nanoparticle
systems show potential for the targeted delivery of antibiotic or antiviral therapeutics to
infected regions of the body (Yeh et al., 2020), particularly in the case of inflamed and
infected regions such as lung or cardiac tissue. Additionally, nanoparticles can be de-
signed to “home” to immune cells in circulation or in the lymph nodes based on nanopar-
ticle ligands designed to bind to cellular surface markers. In this case, nanoparticles are
fundamentally interesting for the delivery of vaccines. Several organs present unique chal-
lenges for delivery that have been addressed using nanomedicine. For example, nanopar-
ticles with dense, highly hydrated polyethylene glycol (PEG) brushes have been devel-
oped to penetrate the unique barrier of corneal tissue for eye treatments, leading to
commercial products for addressing macular degeneration. As another example, because
nanoparticles are naturally filtered through the liver, they also prove to be useful systems
for delivery of drugs to the liver. As nanomaterials are found to be highly versatile and
broad in material properties for adapting to challenging organs, their use is anticipated in
several additional areas, including avascular cartilage tissues and the ultimately challeng-
ing blood–brain barrier, an area that will be extremely impactful for the delivery of treat-
ments for neurological disorders.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Novel and Improved Materials for the 21st Century 191

High-Throughput and Data Science Approaches to Discovery

A more systematic approach to nanoparticle design can ensure that the field ad-
vances in a manner that contributes meaningfully to patient care. High-throughput meth-
ods have been devised for generating large numbers of nanoparticles with differing char-
acteristics (e.g., Dahlman et al., 2017). When such approaches are coupled with the ability
to perform large-scale in vitro screens, it is possible to discern nanocarrier candidates that
exhibit increased efficacy for delivery. The potential to generate significant databases that
include interactions of nanocarriers with a range of tumor-associated cells, including pa-
tient-derived cells, can also provide the basis for machine learning approaches for under-
standing and guiding the design of nanocarriers that target tumor or tumor-associated
cells, such as stromal or immune cells. These studies need to be coupled with in vivo
studies to understand the ultimate transport and trafficking properties and targeting effi-
cacy; such studies may also be designed in a high-throughput fashion using DNA barcod-
ing or chemical barcoding methods to label nanoparticles, in conjunction with state-of-
the-art imaging and detection methods coupled with appropriate cell-sorting algorithms.
It is also possible to consider these methods as a means of determining possible bi-
omarkers for patients whose tumors readily take up nanoparticles or whose tumor vascu-
lature may exhibit the enhanced permeation and retention effect, thus making it possible
to identify candidates who will benefit most from nanoparticle therapies. These ap-
proaches represent a renaissance in understanding of and much more rapid evolution of
nanoparticle design. Future work will extend recent accomplishments to a much broader
set of nanocarrier compositions and structures, allowing a greater amount of nanomateri-
als discovery toward tailored nanoparticle function.

Nucleic Acid Delivery

Perhaps the most important and impactful recent advance in drug delivery is the
ability to deliver and transfect nucleic acids, including mRNA, siRNA, and DNA. The
ability to directly encode cells to produce specific proteins or silence genes is powerful
because it enables the direct programming of cells, including the use of CRISPR (clus-
tered regularly interspaced short palindromic repeats) components. Although viral vectors
can be very effective in transfecting genetic material, they also pose health and safety
risks, as well as issues of scale for manufacturing, reproducibility, and tunability. Chal-
lenges to the delivery of nucleic acids are multifold; they include the need to compact a
very large and highly negatively charged macromolecular species to a size compatible
with cell uptake and the need to create a protective coating upon encapsulation of the
nucleic acid to protect against exposure to proteases in the bloodstream. It is also im-
portant for systemic delivery approaches that the encapsulating materials system have a
neutral or negative charge because of the cytotoxicity and protein opsonization resulting
from positively charged nanomaterials. Ultimately, key susceptibilities for delivery of
these biologic systems have been degradation of the RNA/DNA molecules themselves
due to the presence of RNAse or DNAse enzymes in the body. Recent advances in func-

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

192 New Directions for Chemical Engineering

tionalization of the nucleic acids can render siRNA cargoes fairly inert to breakdown un-
der physiological conditions; however, although mRNA can be protected with functional
groups, it remains more susceptible to breakdown and requires packaging that protects
against degradation. Additional issues requiring attention are the targeting of nanocarriers
to desired organ or cell types and the ability to direct trafficking of the nanoparticle to the
cytosol or nucleus, where the cargo can be available for direct transfection or translation.
Cationic lipid nanoparticles (LNPs) have been found to exhibit the above proper-
ties while generating stable complexes that protect nontarget cells and tissues from expo-
sure to the nucleic acid cargo during distribution throughout the body. The positive charge
of the lipids interacts with negatively charged RNA backbones to generate uniform nano-
particles, usually with a net positive charge. If the cationic lipids are combined with
pegylated lipid molecules, the PEG can act to dilute and shield positive charge, making
the resulting structures increasingly stable while avoiding interactions with serum proteins
and circulating cells, such as monocytes, that lead to clearance and destabilization. Key
to the successful delivery process is the need for the cargo to exit the endosome as it
buffers to lower pH. Cationic LNPs are the basis for effective endosomal escape, enabling
release of RNA or DNA to the cytosol; it is thought that the lipids aid in compromising
the endosomal membrane and lead to its disruption and release. The first LNP therapeutics
were introduced by Alnylam as an siRNA therapeutic and involved a cationic lipid com-
plexed with a chemically stabilized RNA interference (RNAi) molecule. Significant ad-
vances in the composition and structure of the lipids used for LNPs enabled the develop-
ment of the Moderna and Pfizer mRNA vaccines for COVID-19 and form the basis of a
powerful platform for RNA medicine. Novel and improved formulations and production
of LNPs will be important for future vaccine manufacturing.
There are some limitations to the use of cationic LNPs. Lipids tend to traffic to
specific regions of the body, including the liver, because of the apolipoprotein E (ApoE)
receptor present on liver cells, and it can be challenging to design lipids with appropriate
biodistributions for targeting other organs following intravenous injection. Lipids also
provide fewer chemical handles for introducing smart or responsive chemistries, and they
do not allow for staged or more complex release systems. Polymeric systems excel in
providing a large design space while enabling the presence of high-charge-density seg-
ments by using block or segmented block copolymers or polyelectrolytes. Although some
polymers have been found to be highly effective transfection agents, a trade-off between
efficacy and toxicity has resulted from the impact of their high positive charge. One of
the important biomaterials challenges in the upcoming decades will be the discovery and
design of synthetic vectors that can rival the transfection efficiencies of viral delivery
while remaining highly safe.

ELECTRONIC MATERIALS

Chemical engineers have played a central role in the discovery, design, and pro-
duction of the materials (e.g., polymers, semiconductors, glasses) needed to create elec-
tronic devices. The growth in demand for these devices, and thus these materials, shows
no signs of slowing. The development of new materials for these applications will require

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Novel and Improved Materials for the 21st Century 193

sophisticated research and development, as discussed here, and will likely require ele-
ments of AI and data analytics for the materials development (see Chapter 8).

Semiconductor Manufacturing

The common process steps in manufacturing semiconductor devices are insulator


deposition, lithography, etching, wet cleaning, metallization, and chemical–mechanical
planarization, although not necessarily in that order (Table 7-1). Common challenges for
these steps include ever-increasing purity requirements in response to increased up-
stream/downstream process sensitivities, environmental considerations in response to
global requirements and compatibility, means of maintaining consistent manufacturing
and packaging processes, and approaches for managing and improving global distribution.
There can also be challenges specific to individual molecule or formulation types, such
as the stability of the formulation and its sensitivity to air and the safe handling of flam-
mable or toxic materials.
The critical processes for materials manufacturing include synthesis, purification
(including distillation and filtration), formulation blending, and packaging and materials
compatibility. Each of these processes may be considered during the discovery phase of
a new material but are fully accounted for during the transition to mass production.
Several trends are expected to continue. First, the number of new materials being
introduced to the semiconductor and adjacent electronic industries will continue to in-
crease, and more rapid molecule or formulation identification will be required to meet the
industry cadence. In addition, some historically important processes may become obsolete
because of environmental restrictions and a lack of supply due to supply chain interrup-
tions (especially in the case of rare earth elements) or general shortages (as in the case of
helium).
Second, materials will need to be increasingly pure, with the purity levels of met-
als reaching the low parts-per-trillion range, and greater synthetic process control will be
needed to reduce already low-level organic and inorganic impurities. The enhanced purity
requirements will drive improvements to in-line analytical process monitoring and real-
time process control, as well as other advances in manufacturing. Chemical engineers can
play a role in each of these areas. Finally, the markets for most major materials used for
semiconductor manufacturing are expected to grow (Tremblay, 2018), driven by increas-
ing use of materials in manufacturing of both logic and memory devices, as well as overall
industry growth.

Challenges in Electronic Materials Discovery

Research and development of electronic materials occur in sequence but have


different inputs, outputs, timelines, and risks. For suppliers of materials to the electronics
industry, the input to the research process is an established material need from device
makers, and the output is enabling knowledge, ultimately in the form of a product concept
to satisfy that need. The product concept is the input for the development process, which
determines how this concept can be realized as a salable product manufactured economi-
cally and reliably at scale.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering
TABLE 7-1 Primary Semiconductor Manufacturing Processes, Common Materials Used in Each Process, and Chemical Engineering

194
Processes Required
Manufacturing Current Material Related Chemical
Process Step Type of Processing General Material Classes Challenges Engineering Processes
Deposition  Plasma-enhanced  Organosilane  Safe handling  Synthesis
 Chemical vapor  Silicon-containing polymers  Environmental and purity  Purification
Copyright National Academy of Sciences. All rights reserved.

 Atomic layer  Organometallics  Packaging


 Spin-on  Metal-containing formulations  Chemical
 Electroplating distribution
 Physical vapor

Etching and dopant gases  Plasma-assisted etching  Inert and reactive gases  Safe handling  Synthesis
 Halogenated gases  Environmental and purity  Purification
 Mixed specialty gas blends  Packaging technology  Packaging

Lithography  Spin coating  Formulated-polymer blends  Environmental and purity  Polymer synthesis
 Solvents  Distillation
 Metal-containing polymeric  Purification
blends

Wet cleaning  Spin coating  Aqueous, semiaqueous, and  Environmental and purity  Chemical mixing
 Immersion bath solvent-based formulations  Purification
 Acids, bases  Filtration
 Solvent  Packaging

Chemical mechanical  Spin coating  Particle-containing aqueous  Environmental and purity  Chemical mixing
planarization formulations  Purification
 Filtration
 Packaging
SOURCE: Internal knowledge from EMD Electronics, 2021.
New Directions for Chemical Engineering

Novel and Improved Materials for the 21st Century 195

The unmet current and future needs for advanced materials for the electronics
industry are being addressed through close collaboration with foundries and independent
device manufacturers (IDMs), precompetitive consortia, and industry roadmaps. Industry
roadmaps, now in their third generation (Gargini et al., 2020), take the longest view in
identifying key trends, industry challenges, potential solutions, and required innovations
over the next 15 years. These roadmaps have been invaluable in accelerating industry
advances by advancing the best technologies to keep pace with Moore’s law. Material
suppliers with appropriate resources can confidently make long-term strategic research
and business decisions based on these roadmaps.
Since the formation of SEMATECH as an industry–government partnership to
maintain U.S. leadership in the semiconductor industry (Irwin and Klenow, 1996) and its
evolution into an international precompetitive consortium (Carayannis and Alexander,
2004), precompetitive strategic partnerships have been instrumental in the electronics in-
dustry to advance the technologies needed to realize industry roadmaps (Logar et al.,
2014). Medium-term needs for materials suppliers can be understood through participa-
tion in these consortia. Near-term manufacturing needs can be understood only through
close collaboration with the foundries and IDMs. Collaborative innovation within the
semiconductor industry confers an advantage despite the challenges of maintaining rela-
tionships and controlling proprietary information (Kapoor, 2012). Materials suppliers can
receive timely feedback on new product performance and integration issues, which ena-
bles faster implementation of new processes for device manufacturers (O’Neill and
Zheng, 2019).
Once a material need has been specified, the research process of discovery begins.
That process has evolved over the past several decades but follows the basic cyclical sci-
entific method of relevant field study, hypothesis development, testing, and assessment.
In the past, this cyclical method was a slow, trial-and-error process. Leveraging the arrival
of more powerful computers and improved theoretical calculations in the second half of
the last century, the current so-called third age of quantum chemistry now enables calcu-
lations at least as accurate as the results of physical experimentation (Langhoff, 1995;
Richards, 1979). Thus, there is now a rational approach to the design of materials, one
that entails creating new molecules with desired properties using predictions from models
based on atomic or molecular structures. The accuracy and utility of these models are
directly tied to the rapid scaling of computational power, which itself was enabled by the
discovery of new processes and materials used by the semiconductor industry. These im-
provements led to better and faster computational results, such that screening experiments
can be conducted reliably in silico (Hafner et al., 2006; Hautier et al., 2012), limiting the
synthesis targets for physical experimentation to compounds or mixtures with the best
prospects of achieving the material needs of the industry. Additionally, laboratory exper-
imentation has been enhanced by the commonplace computer control of experimental ap-
paratus and analytical instrumentation (Ford, 1982). Thus, advanced computing has in-
creased efficiencies in the testing phase with high-throughput screening and combinatorial
material synthesis. These approaches revolutionized pharmaceutical research but have
now been demonstrated for discovery of new materials in semiconductors, especially

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

196 New Directions for Chemical Engineering

when the new material can be synthesized and evaluated as a thin film (Takeuchi et al.,
2005).

Strategies for Development of Electronic Materials

Chemical engineering merges chemistry fundamentals and material physics with


the hardware and equipment required to produce the desired commercial chemical prod-
ucts. The electronic materials manufacturing industry elevates this challenge to a new
level not experienced with the manufacture of many traditional chemical products. Com-
parison with the evolution of the pharmaceutical chemical industry offers a useful guide
for the advancement of the electronic materials industry.
Purity and product uniformity have improved throughout the progress of the elec-
tronic materials industry. Early in the history of integrated circuit (IC) manufacturing,
materials were used as they were available in the standard reagent and commodity chem-
icals market. Later, IC manufacturers, using statistical tools to characterize chip-manu-
facturing performance, drove the quality and specifications for materials far beyond what
was generally available. Thus, the development of purification technologies is a primary
driver for new products across the industry, and also drives the need for continuous quality
verification using process analytical technology and supporting statistical tools that must
be upgraded to handle the massive amounts of data generated with automation. The future
will look more and more toward AI to enhance problem solving and material and process
design (see Chapter 8).
Materials used in the electronics industry are constantly changing in response to
the new requirements of manufacturers. Many of these materials are unique in structure,
can be very hazardous, and require customized reactors and processes to produce. These
reactor and separation unit operations require a great deal of creative technology design.
The development of robust processes can be accelerated when capital costs are kept low
and process equipment can be kept small. Process-intensification methods, modular con-
tinuous reactors, microreactors, and purification platforms can speed up product commer-
cialization, with assets prepositioned globally (Tian et al., 2018).
Many of the raw materials used to manufacture electronic chemical products are
available only as commodities with limited purity, far below the necessary requirements.
Efforts are increasingly focused on the development of purification technologies using
traditional and novel purification equipment, methods, and schemes. Additionally, the
stringent requirements for product purity require that suppliers of electronic materials
evaluate and characterize the subtle and increasingly important impact of equipment ma-
terials of construction on process chemistry. The industry organization SEMI provides
standards for purity, analytical methods, and processing for many raw materials.1
Metallic impurities are notoriously problematic in all electronic materials, limit-
ing the materials of construction suitable for manufacturing and separation equipment.
Even extremely low levels of surface corrosion can lead to unusable products. Polymeric
materials, on the other hand, can leach low levels of unbound polymers or plasticizers.

1
See https://www.semi.org/en/products-services/standards.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Novel and Improved Materials for the 21st Century 197

Glass reactors and process equipment can be vulnerable to extremes in pH, which can
cause leaching of silica, boron, and metals. Surface science is a critical and ever more
important discipline, and chemical engineers will play a role in the development of new
products and processes for this industry.
Ion-exchange resin processing is commonly used to remove metal impurities in
aqueous and organic raw materials. The diverse chemical functional groups available on
resins provide a wide range of metal-removal selectivity. The choice of resins and com-
binations of resins can be made specific to the metals (and their oxidation state) to be
removed. Preparation and pretreatment of the resin bed are important to reduce polymer
contamination. Materials often need to be diluted before processing to achieve the neces-
sary removal purity and then reconcentrated (Alexandratos, 2009; Silva et al., 2018).
Other purification methods include many variations on distillation, filtration, and
extraction. They are usually applied for small-scale continuous processes, and include
packed-column distillation, molecular distillation, wiped-film and spinning-band distilla-
tion, sublimation, liquid extraction using Karr columns and rotating-disc columns, crys-
tallization, adsorption for both liquid- and gas-phase materials, ion exchange, crossflow
and flow-through filtration on novel modified membranes and plastics, and simulated
moving-bed chromatography. Hydrogen peroxide, for example, is a critical material used
in chemical mechanical polishing, and a new approach to reverse osmosis is proposed as
a novel method for purification (Abejón et al., 2010; Gao et al., 2017).
Cleaning procedures as they pertain to such equipment as piping, pumping, and
storage vessels are frequently underappreciated. Especially in the pharmaceutical, fine
chemical, and electronic chemicals industry, cleaning of equipment between changes in
product, raw material lots, or batches is often a complex process. Cleaning processes re-
quire detailed metrics, and in many cases may be as complex and time-consuming as the
chemical synthesis or purification process. Chemical surface preparation of unit operation
equipment and packaging containers will become a project in surface science and a subset
of the overall process development.

CHALLENGES AND OPPORTUNITIES

Chemical engineers can contribute to the materials development domain across a


range of material types and applications. In particular, they have a unique role to play in
the continued development of polymer science and engineering because of their under-
standing of chemical synthesis and catalysis, thermodynamics, transport and rheology,
and process and systems design. Chemical engineering is the logical home for research
and development of complex fluids and soft matter. The combination of molecular-level
understanding and thermodynamic and transport concepts yields important insights and
enables advances. The science and application of nanoparticles by chemical engineers in
both industry and biomedicine are rapidly accelerating, offering the opportunity for ap-
plication breakthroughs. Chemical engineers play an essential role in advancing the de-
velopment of biomaterials for both regenerative engineering and organ-on-a-chip tech-
nology, and chemical engineering principles are at the heart of understanding and
improving targeted drug delivery both spatially and temporally. As the United States has

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

198 New Directions for Chemical Engineering

lost dominance in the area of semiconductor processing, chemical engineering expertise


around reactor design, separations, and process intensification has become critical to the
success and growth of the electronic materials industry.

Recommendation 7-1: Federal and industry research investments in materials


should be directed to

 polymer science and engineering, with a focus on life-cycle considera-


tions, multiscale simulation, artificial intelligence, and structure/prop-
erty/processing approaches;
 basic research to build new knowledge in complex fluids and soft matter;
 nanoparticle synthesis and assembly, with the goal of creating new ma-
terials by self- or directed assembly, as well as improvements in the safety
and efficacy of nanoparticle therapies; and
 discovery and design of new reaction schemes and purification processes,
with a steady focus on process intensification, especially for applications
in electronic materials.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

8
Tools to Enable the Future of Chemical Engineering

 Current and future chemical engineers will need to navigate the interface between
the natural world and the data that describe it, as well as use the tools that turn
data into useful information, knowledge, and understanding.
 Some emerging and future tools will be developed in other fields but will have a
significant impact on the work of chemical engineers; others will be developed
directly by chemical engineers and have an impact in science and engineering
more broadly.
 Some tools and capabilities will be evolutionary, with gradual and predictable
development and applications, while others will be revolutionary and will change
chemical engineering research and practice in ways that may be difficult to pre-
dict or anticipate today.

Earlier chapters concentrate on specific aspects of the discipline of chemical en-


gineering (and by extension, of its students, academicians, researchers, and practitioners);
this chapter is concerned with the tools and capabilities that will be required in the field’s
endeavors in the future. Consequently, the content of this chapter cuts across the topics
covered in earlier chapters and is motivated by and organized around a defining question:
What existing and anticipated tools and capabilities will drive chemical engineering and
chemical engineers of the future?
Some of these tools and capabilities currently exist in some form but will evolve
in the future, some are emerging as this report is being written, and yet others will need
to be created and will coevolve with chemical engineering over the next few decades.
Some of these tools will emerge from current and future trends in technologies, with little
or no involvement of chemical engineers in their development even though they will have
significant impacts on chemical engineering. In other cases, chemical engineers will be
involved directly in the development of tools and technologies that will advance science
and engineering more broadly. Some of these tools and capabilities will be evolutionary
in the sense that their development and application will be gradual and more predictable;
others will be revolutionary in that either their development, their application, or both will
alter the landscape of chemical engineering research and practice in ways that may be
difficult to predict or anticipate today.
Whether engaging in cutting-edge medicine or manufacturing; revolutionizing
the design, optimization, and operation of efficient energy systems that respect environ-
mental constraints; discovering and designing new materials; or innovating in any of the
other endeavors discussed in earlier chapters, the chemical engineer of today and tomor-
row will, conceptually, need to navigate the interface between the natural world and the

199

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

200 New Directions for Chemical Engineering

data that describe it. What devices (physical or otherwise), tools, or capabilities will make
it possible to turn data into useful information, knowledge, and understanding in the fu-
ture? While the discussion here could address a virtually endless list of tools and capabil-
ities—many of which, when used in combination, will drive innovation—the focus is on
four categories: data science and computational tools, modeling and simulation, novel
instruments, and sensors.

DATA SCIENCE AND COMPUTATIONAL TOOLS

Chemical engineering has been a data-intensive field from the very beginning.
The chemical industry was built on systematic measurement and cataloging of experi-
mental measurements, including thermodynamic properties, phase diagrams, rate con-
stants, flow rates, and heat- and mass-transfer coefficients. Early steam tables from the
1930s exemplify the importance of data-intensive approaches in chemical engineering.
These data led to the development of correlations, equations of state, and process simula-
tors, which have allowed industry to design, optimize, and innovate. Until the late 1990s,
chemical engineering datasets were stored primarily in print—notebooks, publications, or
books. After a glossary of definitions, the first technical chapter of Perry’s classic chem-
ical engineering handbook is a compilation of physical property data (Green and
Southard, 2019). The National Institute of Standards and Technology (formerly the
National Bureau of Standards) was entrusted by the federal government with the collec-
tion of such data, which were subsequently published in the form of much-sought-after
compilations and books.
Chemical engineering is undergoing a transformation driven largely by the way
data can be acquired, curated, shared, and utilized. Many labor-intensive measurements
traditionally performed by humans—from calorimetry, to composition measurements us-
ing chromatographs equipped with autosamplers, to more sophisticated systems capable
of carrying out and quantifying chemical reactions—are now routinely automated, mak-
ing it easier than ever to acquire data in the laboratory (Sanderson, 2019). Open-source
databases are now commonplace in such areas as biology (Karsch-Mizrachi et al., 2012);
thermophysical properties; materials properties (CHiMaD, 2021; Jain et al., 2013; NIMS,
2021; PRISMS Center, 2021); and climate, water management, and agriculture (Univer-
sity of Hertfordshire, 2021). The venerable steam tables, once available only in print, can
now be accessed on the internet (NIST, 2021). These transformations are propelling ad-
vances on multiple fronts, but they are just the beginning. For example, the chemical in-
dustry only recently began the process of digitizing data previously stored over the course
of many decades in laboratory notebooks. However, the infrastructure needed to ingest
and organize such data most efficiently is not yet available, limiting the potential use of
these vast stores of existing data. The way that infrastructure is developed over the next
decade will determine how rapidly and how effectively companies and researchers can
benefit from gaining access to vast amounts of information from multiple sources.
Moving forward, industry will develop systems that include not only internal,
proprietary data but also data mined from the literature and from publicly accessible and
commercial databases. Natural language processing tools are improving at a rapid pace,

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Tools to Enable the Future of Chemical Engineering 201

now capable of extracting sought-after information, such as experimental data needed to


validate computer models, from the published literature (Olivetti et al., 2020). How one
uses data and integrates data from multiple sources to generate valuable information is
expected to become a major differentiator in highly competitive industries. Recent exam-
ples that illustrate this point include the optimization of water and energy use in paper
mills (Ahmetović et al., 2021); the optimization of raw materials production from natural
sources, such as switchgrass (Galán et al., 2021); and the production of renewable fuels
from multiple sources (König et al., 2020). If they are to contribute to the development of
such tools, chemical engineers will need to develop skills to exploit new possibilities as
they emerge (Rzhetsky et al., 2015).
The growing ubiquity of data, along with increasingly powerful computers, net-
works, and algorithms, is creating opportunities for integration of data from disparate
sources on scales never before imagined. It is now possible to monitor every aspect of a
chemical plant through a myriad of sensors capable of collecting data continuously, and
it will become increasingly common to couple such “lower-level” process data to other
types of “higher-level” data pertaining to global supply chains, raw materials availability
and characteristics, failure data, responsible water management (Caballero et al., 2020),
or climate conditions throughout the world. This capability will enable chemical engineers
to optimize enterprise-wide performance at levels not envisaged a decade ago (Hubbs et
al., 2020). The COVID-19 pandemic has revealed the fragility of worldwide supply
chains, arising from their inherent interconnectedness and the “just-in-time” operating
philosophy of many businesses. Chemical engineers will have enormous opportunities to
reimagine and redesign these systems to exploit the new ubiquity of and instantaneous
access to data to achieve unprecedented robustness. Modern approaches to the design of
all processes and products will revolve around the instant availability of relevant data.
Growth in the volume of worldwide data is creating both urgency and opportunity for
chemical engineers to develop the tools needed to assess data quality and in turn transform
all the data into information and knowledge.
Other opportunities for chemical engineers will arise from combining technical
and sociological data to achieve advances that could not have been anticipated just a few
years ago. Examples include quantifying the dynamics of failure in scientific endeavors
or technology translation (Yin et al., 2019) and reorganizing in real-time teams that are
optimal for a given collaborative task (Wang et al., 2020). Combining sociological data
with, for example, real-time geographic data on pollution, energy distribution, or water
availability will allow chemical engineers to integrate sociological implications into en-
gineering considerations to ensure that technology solutions are equitable and inclusive
and do not disadvantage any population groups.
Indeed, the possibilities for integrating data and information from disparate
sources and domains of inquiry are limitless. A massive amount of data will be available
on climate, pollution, water transport, and the environment; on global and local supplies
of raw materials; on how processes operate or perform when materials are altered; on such
outcomes as quality and consumer demand for the corresponding products; and on human
health (e.g., heart rate, blood pressure, oxygen level), both aggregated and individualized.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

202 New Directions for Chemical Engineering

What new solutions, technologies, and industries will be possible through clever integra-
tion of disparate data? The following are some concrete examples.
Process automation. Highly automated processes will need to be developed
based on vast sensing capabilities that previously did not exist, allowing real-time process
management and dynamically optimized performance. This domain will grow rapidly as
part of Industry 4.0.1 Given its long history with data-driven approaches, the chemical
industry will be quick to benefit from such innovations, while other industries, such as
food and biopharmaceuticals, may be slower to benefit because of regulatory issues and
relative conservatism in the adoption of new technologies.
Human health. Chemical engineers will aid in advancing the biomedical sci-
ences to facilitate the practice of personalized medicine by addressing the human body as
a system, informed by the massive amounts of data generated by medical devices and
fitness trackers. Wearable devices invented by teams of chemical engineers and medical
professionals will synthesize a multitude of real-time data streams to enable the instanta-
neous transmission of health information to physicians and phones. In the future, such
information may be used to prevent everything from headaches to heart attacks. With their
expertise in microfluidics, nanotechnology, and process control, combined with their deep
understanding of biological processes within a systems framework, chemical engineers
will drive invention and innovation in this space.
Additive manufacturing. Additive manufacturing enables on-demand produc-
tion of materials, metamaterials, and parts. New data science tools will enable ever-more-
rapid prototyping and scale-up, real-time assessment of performance in situ, and judicious
adjustments based on continuous feedback from finite-element models running in parallel
and in real time. New tools may one day make it possible to adapt manufacturing pro-
cesses autonomously and in real time and to use immediately available raw materials
based on local supply chain data.
Materials and molecular design. The rapid design of molecules or materials for
targeted applications will be enabled by the development of widely available “living” da-
tabases that are updated automatically with experimental data on newly developed mate-
rials, and/or with computational data providing predictions about such materials and in-
corporating updates as they become available. Similarly, understanding of structure–
property relationships ranging from the molecular to the manufacturing scale will enable
“inverse design and synthesis,” whereby one specifies the end-use characteristics and per-
formance objectives for materials, and the computational tool determines the molecular
components required to achieve them. Such tools will enable precision material design
from the nanometer to meter scale.
Quantitative structure–property relationships (QSPR) and quantitative
structure–activity relationships (QSAR). Statistics and probability—the classic tools of
data science—also have a long history in chemical engineering, as exemplified by the use

1
Industry 4.0, or the fourth industrial revolution, denotes the automation of traditional manu-
facturing and industrial processes.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Tools to Enable the Future of Chemical Engineering 203

of statistical methods for QSPR and QSAR—mathematical relationships linking, for ex-
ample, a particular molecule to its structure or pharmacological activity, respectively
(Cherkasov et al., 2014).
Automated/self-driving laboratories. As discussed further below, a future in
which artificial intelligence (AI) is used to power automated laboratories is already taking
form. To illustrate this concept, consider a polymer material designed to achieve certain
end-use performance objectives. Given the raw materials and laboratory process needed
to manufacture this material, a robot produces it by running the predesigned process, sub-
sequently characterizing the polymer, and making any necessary feedback corrections au-
tomatically until the desired material performance is achieved. Of course, realizing this
scenario in practice will require first developing all the sensors and robotics, along with
the chemistries and characterization, necessary to enable high-throughput operation. To
carry out syntheses, today’s automated laboratories require extensive human intervention
for setup and programming. Methods for automated extraction of chemical synthesis ac-
tions from experimental procedures documented in papers and laboratory notebooks have
been described that can enable automatic execution of arbitrary reactions using robotic
systems (Vaucher et al., 2020). IBM Research has described a system2 in which a dataset
of chemical procedures derived from cloud-based literature systems is integrated with
cloud-based control of a robotic synthesis machine. In a live demonstration, this system
identified a route to synthesis of 3-bromobenzylamine and carried out the process using
air-sensitive reagents (Extance, 2020).
The preceding examples represent only a fraction of the ways in which the gen-
eration of and access to data will change the way chemical engineers work. Tomorrow’s
chemical engineers will need to be capable of using the massive amount of data of all
sorts whose ready availability will continue to increase. It is easy to imagine a not-too-
distant future characterized by data-on-demand, where data on any subject, at any level
of granularity, will be readily and instantly accessible. Such a future suggests profound
and exciting opportunities for chemical engineers, who are trained in process integration
and systems-level thinking—skills that will be required to synthesize disparate data
streams into information and knowledge.

Artificial Intelligence

AI is one of the modern tools of data science that is rapidly transforming all fields
of science and engineering, including chemical engineering. Over the next few decades,
AI is expected to transform all industry sectors in profound ways; indeed, the ways in
which AI is used will differentiate companies with respect to gaining a competitive edge.
Whereas prior revolutions in chemical engineering were driven by reductionist models,
there is ample evidence that the next revolution surely will be data-driven and powered
by AI.
AI is a general term used to describe “machines” (fashioned primarily by models
and algorithms) designed to mimic human intelligence. With their unique combination of

2
See https://rxn.res.ibm.com.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

204 New Directions for Chemical Engineering

knowledge and skills—in particular in mathematics, statistics, computers, analytical


thinking, systems integration, and process control—chemical engineering researchers
have been among the first to integrate AI into their toolset, and now into the discipline’s
curriculum. Engineering colleges are challenged to keep pace with the increasing demand
for courses in AI, and in particular in machine learning (ML), the subfield of AI concerned
with the development and use of computer systems to learn adaptively from patterns con-
tained in data, using algorithms and statistical models, without human supervision. At the
same time, chemical engineering departments are under pressure to teach data science and
ML in the context of their discipline, in part because of changing expectations and needs
of employers for those who hold chemical engineering degrees. Each year, an increasing
number of chemical engineering graduates take nontraditional jobs as data scientists at
such companies as Facebook, Airbnb, Microsoft, Google, Uber, Amazon, and more. In a
growing number of instances, these companies are hiring chemical engineers for their data
science teams preferentially over computer scientists and even data scientists because they
recognize that a large part of “doing” data science lies in asking the right questions using
existing AI tools rather than, for example, inventing new AI machines.
While chemical engineering has been data-driven since its inception, what is new
to the discipline is the explosive growth in the use of AI—and in particular ML—in mak-
ing predictions from data. Perhaps the most rapidly growing use of AI by chemical engi-
neers is in the area of deep learning—a subset of ML referring to the use of deep neural
networks (DNNs), a class of ML methods whereby multilayered neural networks (NNs)
are trained to perform tasks or discover hidden correlations in datasets far too large to be
handled any other way. Simply put, NNs use structures motivated by networks of biolog-
ical neurons to learn, through iteration (training), the relationship between inputs and out-
puts. Because the outputs are labeled and defined explicitly, the learning is said to be
“supervised,” as opposed to “unsupervised learning,” exemplified by clustering, where
the output is not predefined explicitly, and the algorithm’s task is to discover groupings
within data. In NNs, the network architecture is essential to its success. Examples of NN
architectures used in chemical engineering research today include convolutional NNs
(CNNs, used, for example, to learn images); generative adversarial networks (GANs);
recurrent NNs (RNNs); and long short-term memory networks, where each descriptor is
indicative of the unique characteristic of the NN structure in question. DNNs can have
millions of parameters (node weights) to train and therefore require massive amounts of
data to make useful predictions—a challenge for many research problems. Nevertheless,
NNs and DNNs have seen success in a number of application areas within and adjacent
to chemical engineering. For example, Shell employed four-layer gated RNNs to predict
valve failures a month prior to failure (Gupta, 2018).
While ML concepts have existed for decades, only in recent years have computer
chip speeds and architectures become fast enough for ML to become sufficiently practical
to realize its promise. In addition to enabling faster computations, advances in computers
have made it easier to collect, store, and manage large datasets. Furthermore, significant
advances in the theory of statistical learning have led to new ML algorithms. The explo-
sive adoption of ML by nearly every technology sector and research area has been fueled
by the development and sharing of powerful, open-source, easily accessible libraries that

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Tools to Enable the Future of Chemical Engineering 205

simplify many of the most cumbersome tasks. This accessibility and ease of use of librar-
ies is in large part responsible for the rapid adoption of ML by chemical engineers.3 Such
libraries encode expertise outside the typical knowledge domain of chemical engineers,
providing naturally effective ways of expressing and working with complex computa-
tions, and thereby facilitating the implementation of data science and ML pipelines.
The relationship between science and engineering disciplines and AI software
development has been mutually beneficial: the aforementioned libraries have seen mas-
sive investments and improvements because of the value they have provided to data sci-
ence as a discipline, while AI’s successes with scientific and engineering problems,
among others, are driving continued increases in the use of ML software. Science and
engineering generally can expect to continue to reap the benefits of industry investments
in AI. A 2020 report from Grand View Research suggests that the AI industry will see a
compounded annual growth rate of 40.2 percent between 2021 and 2028 (Grand View
Research, 2021).
Today, chemical engineering researchers at universities are already using deep
learning alongside experimentation and computer simulation. Examples of the many
problems for which chemical engineers are making exciting breakthroughs using ML in-
clude mapping equilibrium phase diagrams, predicting system failure well in advance,
evaluating metabolic networks, designing new molecules and materials, developing im-
munotherapies, understanding cellular processes and disease, and controlling nonequilib-
rium processes (Dobbelaere et al., 2021; Sanchez-Lengelingand Aspuru-Guzik, 2018;
Venkatasubramanian, 2019). The swift adoption of AI by chemical engineering research-
ers is reflected in the steep rise in the number and frequency of American Institute of
Chemical Engineers (AIChE) fall meeting presentation titles and abstracts containing AI-
related terms. Between 2006 and 2020, the total number of presentations involving work
using data science in some form increased from 0 to nearly 400, while the fraction of such
presentations at a given meeting increased from 0 percent to nearly 7 percent (Figure 8-
1). (In each dataset, an abstract is counted only once if the term of interest appears.) The
appearance of no other words or two-word phrases increased as rapidly over this period.
For comparison, the two most commonly found technical words in fall meeting AIChE
abstracts are “bio” (in ~30 percent of all abstracts) and “nano” (in ~20 percent of all ab-
stracts), frequencies that have not grown substantially since 2006.
AI has many potential future roles in chemical engineering. The following sec-
tions describe two areas in which the impact of AI in chemical engineering is expected to
be profound: manufacturing and materials discovery and design.

3
Examples of such libraries written for the Python program language include NumPy (performs
numerical calculations), pandas (enables data cleaning and analysis), Scikit-learn (implements ML
models), SciPy (provides numerical routines and solvers common in science and engineering),
TensorFlow (provides a scalable platform for performing ML calculations), Matplotlib (creates
data visualizations), and Keras (simplifies the definition and training of neural networks for deep
learning).

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

206 New Directions for Chemical Engineering

FIGURE 8-1 Count (top panel) and fraction (bottom panel) of American Institute of Chemical
Engineers annual meeting abstracts that include terms related to data science, 2006–2020.

Artificial Intelligence and Data Science Applications in Manufacturing

Enabled by the internet of things (IoT), the availability of and role played by data
in large-scale manufacturing have changed dramatically in recent years. It is now reason-
able to expect that ubiquitous real-time sensors will be a core part of any future manufac-
turing operation, and that in general, relevant detailed information will be available at all
scales, from the individual plant to the overall supply chain. The result will be a rich set
of opportunities not only in the productive use of data but also in the methods for gener-
ating high-quality data and for maintaining the security of these processes. In each of
these three areas, manufacturing industries are innovating rapidly without waiting for ac-
ademic research to lead.

Productive Use of Data with Artificial Intelligence

An immediate implication of the above trends is the existence of a “data flood”


in many manufacturing settings. Numerous AI tools have been developed to use such data
effectively. A major challenge for chemical engineers, however, is that most data science
methods originated in fields in which the implications of “off-spec” predictions are less

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Tools to Enable the Future of Chemical Engineering 207

dramatic than would be the case at a large-scale chemical plant. A related challenge is that
many of these methods are not transparent, with components that have no recognizable
connection to the physical characteristics of the process in question. Their “black box”
predictions are therefore not readily interpretable, a challenge extending beyond manu-
facturing. These challenges highlight important research and development (R&D) direc-
tions for chemical engineers in combining data science methods judiciously with the best
aspects of traditional, physics-based models. An underappreciated aspect of data science
with particular importance in manufacturing is data quality, curation, and provenance.
Important R&D challenges lie in automated assessment of data quality—for example, in
the detection of sensor faults in large networks of sensors.

Innovative Modes of Obtaining High-Quality Data

Some aspects of introducing the IoT into manufacturing are as conceptually sim-
ple as measuring a process variable, such as temperature, at hundreds of points in a plant
instead of at a handful of points. However, there are important areas in which R&D could
enable collection of critical data that are currently challenging or impossible to obtain. In
a traditional facility, sensors are placed in fixed, predetermined locations. Efforts are al-
ready underway in industry to collect data from a variety of locations using mobile
sources, such as robots or autonomous vehicles. Chemical engineers have opportunities
to collaborate with adjacent disciplines to explore the full implications of these concepts
for manufacturing operations. They have opportunities as well in the development of im-
aging methods that provide low-cost, real-time insight into conditions inside inaccessible
spaces (e.g., in reactor vessels) or at nanometer scales, perhaps by leveraging insights
from associated areas, such as medical imaging.

Cybersecurity Implications in Manufacturing

Even as the IoT opens up unprecedented avenues for control and management of
manufacturing facilities, these same avenues create enormous cybersecurity challenges.
Well-publicized incidents of cyber-enabled attacks on physical resources in chemical
manufacturing have already taken place. Significant needs for R&D include adapting gen-
eral principles from the field of cybersecurity to the specific challenges of large-scale
manufacturing. Just as data science methods can be used for fault detection, similar meth-
ods could be used to detect possible breaches of cybersecurity. It is notable that, as is the
case with more familiar aspects of process safety, cybersecurity is a need shared by all
parties in the chemical industry, suggesting that public–private consortia focused on
precompetitive R&D would be well suited to addressing this challenge.

Artificial Intelligence Applications in Materials Discovery and Design

The nascent fourth paradigm of science has evolved rapidly over the past decade,
enabling the use of big data and ML for the design and realization of new, better materials

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

208 New Directions for Chemical Engineering

(Tolle et al., 2011). In the semiconductor industry, for example, ML is enabling the iden-
tification of property patterns and anomalies for new materials. Chemical engineers are
using ML to predict “sweet spots” for synthesis of new semiconductors, alloys, polymer
materials, and nanomaterials through training on previously obtained experimental and
simulation data. Force fields—necessary for molecular-dynamics simulations of materi-
als—are well known to fail to predict some material properties (e.g., dynamic properties)
accurately while reproducing others (e.g., thermodynamic properties) successfully.
Chemical engineers are using ML to generate ab initio and phenomenological force fields
that can accurately describe multiple material properties simultaneously, thereby enabling
more complex simulation studies of phase transformations in materials, for example. They
are designing new catalysts using a combination of electronic structure calculations and
ML. Nonequilibrium materials processes (such as self-assembly and crystallization) are
challenging to study for many reasons. Combining ML with molecular-dynamics simula-
tions of nanoparticle self-assembly into complex structures, chemical engineering re-
searchers are discovering microscopic details of assembly pathways that can be used to
engineer products with fewer defects. In the future, both reaction engineering and assem-
bly engineering will benefit from deep learning and other AI tools.
Scientific inquiry typically begins with a hypothesis generated by a researcher.
Using AI successfully for hypothesis generation requires a high-quality training dataset.
Depending on the maturity of the field, the exercise may produce a large set of well-
characterized existing materials, or the result may be no more than a potential solution
space with gaps. Fortunately, these gaps can be partially filled through modeling and sim-
ulation, using, for example, ab initio computational methods. Examples of areas in which
high-throughput computational screening of materials have been demonstrated include
optoelectronic semiconductors (Luo et al., 2020), thermoelectric materials (Sarikurt et al.,
2020), and metal organic frameworks (Daglar and Keskin, 2020). In many cases, such AI-
based hypothesis generation can be computationally expensive. In the future, the emerg-
ing field of quantum computing could greatly reduce the cost of these calculations or
enable solutions to problems that would otherwise be intractable using classic computing
architectures (Cao et al., 2019). However, technical and programming challenges need to
be overcome before quantum computing can become a tool in the chemical engineering
toolbox (Almudever et al., 2017).
Once a dataset of relevant materials and their properties (measured or calculated)
is assembled as a training set for ML, generative AI models can be used to fill out the
design space and suggest optimized synthesis targets, even if they were not part of the
original training set. While AI-based discriminative learning (that is, the ability to dis-
criminate among different kinds of data instances) is now commonplace (Ng, 2016), re-
cent improvements in generative modeling (the ability to generate new data instances) are
now allowing the generation of hypothetical new materials (Elton et al., 2019; Maziarka
et al., 2020). Improved generative algorithms can restrict the output of hypothetical ma-
terials to those that are physically possible as manufacturable product (Dan et al., 2020).
In the near term, human experts will be required to select potential, viable synthesis can-
didates from the set of predicted materials.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Tools to Enable the Future of Chemical Engineering 209

In the future, the testing phase of the research process also may be enhanced by
AI. Standardized machine-readable representations of chemical species (simplified mo-
lecular-input line-entry system [SMILES]) and reactions (SMILES arbitrary target speci-
fication [SMARTS]) have been developed. These representations have been recognized
as analogues to a context-free language (Sidorova and Anisimova, 2014). Atoms and mol-
ecules (SMILES) are the letters and words, while reactions (SMARTS) are sentences.
Thus, state-of-the-art computer processing for language may be used to improve models
for predicting chemical reactions and properties (Whiteside, 2019). Earlier approaches to
predicting chemical reactions relied on expert-crafted reaction rules and heuristics to de-
scribe potential retrosynthetic disconnections (Socorro et al., 2005), with mixed results.
More recent approaches are taking advantage of modern AI algorithms and the large re-
action corpora, such as the U.S. Patents and Reaxys databases, to construct purely data-
driven prediction engines (Coley et al., 2018). Beside retrosynthetic approaches, AI algo-
rithms for solving the forward problem of predicting products given a set of reactants and
reagents have been described (Schwaller et al., 2019).

Computing

From mainframes, to desktops, to cloud computing, to graphics processing units


(GPUs), chemical engineers have been among the first to exploit the most powerful com-
puter architectures available. The first uses of computers in chemical engineering can be
traced back to the early 1950s (Wilkes, 2002). Since the early 1990s, chip speed has dou-
bled roughly every 18 months. As a result, computing ability—in terms of both time to
solution and complexity of problems that can be handled—has reliably increased by
roughly three orders of magnitude every decade. What took months to compute in 2000
takes only minutes today and will likely take only a small fraction of a second in 2030.
At the same time, the cost of computing has plummeted, democratizing and increasing
access to supercomputing resources. The implications of this increase in computing capa-
bility and concomitant decrease in cost are profound. It now costs next to nothing to store
the output of large simulations for subsequent interrogation offline, or it may now be
faster to rerun a simulation when new calculations are desired or to perform real-time
calculations as the simulation is running. This flexibility afforded by the broad availability
and speed of both computing and data storage means that it is becoming faster and easier
to explore vast, multidimensional design spaces; predict phase behavior in complex mul-
ticomponent or reactive systems more accurately; or apply ML to large, heterogeneous
datasets. It also means that for many chemical engineering problems, simulation and ex-
perimental length and time scales coincide. This makes it possible for experiments to in-
form model development directly and for computation to be a true equal partner with
experiment in R&D.
Several relatively recent game changers have emerged in the computing tools
available to chemical engineers. An example is the GPU. Driven largely by the multi-
billion-dollar computer gaming industry, GPUs are now ubiquitous throughout scientific
computing and are today a key component of modern supercomputers. Central processing
units (CPUs)—the original main processors in personal computers—have a handful of

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

210 New Directions for Chemical Engineering

compute “cores,” exhibit low latency, and excel at serial tasks, but can handle only a few
operations at a time. By grouping tens of thousands of moderately fast CPUs or hundreds
of powerful CPUs together in a single supercomputer, computations can be carried out in
parallel. Since the first parallel computers appeared in the early 1990s, chemical engineers
have exploited their capabilities to great effect.
CPU-based parallel computing was the cutting-edge tool in scientific computing
until about 2010, when GPUs emerged in personal and business computers. Since the
invention of the GPU by NVIDIA in 1999, every desktop or laptop computer has con-
tained a GPU to process graphics displayed on the screen more quickly than was possible
with the CPU, which now handles all remaining tasks. In contrast to CPUs, GPUs contain
many cores, are high-throughput, excel at parallel processing, and can handle many thou-
sands of operations simultaneously. Game developers and professional graphics designers
exploited and drove R&D to increase on-chip parallelism to render ever-more-realistic
images in real time. NVIDIA’s public release of the parallel computing language CUDA
in 2007 made the power of GPUs accessible to the science and engineering communities,
which then could take advantage of GPUs by inserting a few simple commands into their
existing codes.
Today, nearly all major scientific software has been or is being ported to or re-
written for GPUs and GPU/CPU hybrid supercomputer architectures, which are expected
to be the primary architectures for at least the next decade. In molecular simulation, for
example, open-source software such as LAMMPS,4 HOOMD-Blue,5 GROMACS,6 and
NAMD (Kass et al., 2011) runs on GPUs, enabling operations up to several orders of
magnitude faster than those using CPU-based codes. Together with standards provided,
for example, by Dockers,7 computational science and engineering applications can be de-
veloped and tested on low-cost laptop and desktop computers; scaled up to run on local
clusters, on federally funded computer facilities (e.g., XCEDE),8 or on cloud-based ser-
vice providers; and then scaled up again to run on the fastest, most powerful supercom-
puters (including the current fastest open-science supercomputer, Summit,9 at Oak Ridge
National Laboratory). The growing use of software “containers”10 greatly facilitates port-
ing simulation code from one architecture to another, making it easier for chemical engi-
neering researchers to use state-of-the-art software without having to know the nuts and
bolts of the computer they are using.
Another commercial driver of GPU technology is AI, with GPUs enabling the use
of deep learning. The ubiquity of GPUs for deep learning has been accelerated by Tensor
Cores, which not only speed up the linear algebra and large matrix operations at the heart
of NNs but also carry out, as a single operation, multiple operations common in gaming.
By enabling faster graphics and faster, more powerful AI algorithms, the video games

4
See https://www.lammps.org.
5
See https://hoomd-blue.readthedocs.io/en/latest/#.
6
See http://www.gromacs.org.
7
See www.docker.com.
8
See www.xsede.org.
9
See https://www.olcf.ornl.gov/summit.
10
See https://www.docker.com/resources/what-container.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Tools to Enable the Future of Chemical Engineering 211

industry, the financial sector, and nearly every major industrial sector are today indirectly
driving breakthroughs in computing for science and engineering. GPUs are also game
changers for processing of imaging data, fueling everything from improved computed
tomography (CT) scans to self-driving cars.
The continued applicability of Moore’s law—achieved by still-decreasing na-
nometer-scale chip features, continued densification and thus multiplication of transistors
on single chips, continued densification of chips on computer boards, faster networking
and switch speeds for faster communications and input/output, new and faster computing
algorithms, declining costs, and ever-increasing accessibility—will continue to enable
chemical engineers to tackle bigger, more difficult, and more complex problems. Yet
while today’s computing power continues to increase, advances in AI are now pushing
the limits of what is currently available, motivating the search for alternative options with
even greater computing capability (Olson et al., 2016). The holy grail of computing is
quantum computing, which until recently has been somewhat of a laboratory experiment
but is now becoming more mainstream as an option available through some cloud provid-
ers. In traditional computing, a transistor holds the information as a 1 or 0 depending on
whether there is a current applied (1) or not (0). In the case of quantum computing, be-
cause the qubits are in a superposition of states, they can hold a continuum of values
between 0 and 1. Thus, in contrast to traditional computing’s binary system, quantum
computers can theoretically hold much more information within each qubit. Given the
reliance on managing quantum states to control the computation, the quantum computer’s
surrounding environment needs to be tightly controlled against vibration and temperature
variation. This is one of the fundamental challenges in quantum computing today, respon-
sible for its high error rates and limiting its adoption. Accordingly, scientists across the
globe are working to reduce error rates and foster confidence in quantum computer results.
Today’s approaches to this error-reduction problem have been based on the use of several
materials, including superconducting materials and ionic fluids, but the expectation is that
mainstream implementation may require a material not yet developed. It is widely ex-
pected that the requirements for strict environmental controls and high-purity materials
for fabricating today’s computer chips will become even more stringent as the computer
industry moves toward exascale and quantum computers. These challenges offer addi-
tional opportunities for chemical engineers.

MODELING AND SIMULATION

Mathematical modeling—the development and use of mathematical equations as


a convenient surrogate for studying, understanding, and even designing a physical en-
tity—has been an indispensable tool for the modern chemical engineer. In general, math-
ematical models range from simple equations based on material and energy balances of
large-scale industrial equipment or empirical correlations connecting one observed phe-
nomenon to another to complex multiscale models that span the entire length scale, cap-
turing atomistic molecular interactions and integrating up to macroscale phenomena.
Traditionally, modeling has almost always been connected to simulation in which
a computer is used to solve the set of modeling equations representing a physical system

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

212 New Directions for Chemical Engineering

so the behavior of that system can be explored under various conditions. However, simu-
lation has evolved to be understood in the more general sense of a computer representation
of a physical entity with or without the use of an explicit mathematical model. As a con-
sequence, any computational representation of a physical system, whether via explicit
mathematical equations or a computer simulation, is understood to be a “model” of the
system in question.
Simulation also refers to systems in which interaction potentials between particles
are developed to describe many interacting species, and emergent collective behavior is
revealed. These approaches are leveraged to understand particles ranging from the quan-
tum to colloidal scale, with detailed or approximate potentials. The results of these simu-
lations often inspire reductionist modeling in terms of mesoscale mathematical models
and can be used to understand or predict the parameters in these mesoscale models. Chem-
ical engineers have developed and applied these approaches to understand biomolecular
interactions essential to drug design and particle interactions and assembly at the heart of
nanotechnology. This cross-talk between simulation to reveal emergent behavior and to
inform model development can be further developed to embrace data-driven and equa-
tion-free approaches to simulating interactions across scales in complex, far-from-equi-
librium systems.
Chemical engineers continue to innovate and drive the field of molecular dynam-
ics (MD) simulation, which was invented in 1957 at Argonne National Laboratory and
further developed over the next several decades, primarily in the United States and United
Kingdom, by chemists and chemical engineers. Today, the most widely used MD codes—
including LAMMPS, HOOMD-Blue, NAMD, and GROMACS, all led by U.S. research-
ers—are open source. Chemical engineers have developed some of the most popular mo-
lecular simulation algorithms, such as free-energy methods (Frenkel and Ladd, 1984),
thermodynamic integration (Kofke, 1993), enhanced sampling methods, ensemble simu-
lation techniques (Panagiotopoulos, 1987; Yan and de Pablo, 1999), and acceleration al-
gorithms (e.g., bond-boost; Pal and Fichthorn, 1999). New generalized ensemble simula-
tion techniques, such as digital alchemy (van Anders et al., 2015), use MD and Monte
Carlo simulation for inverse design of materials building blocks for targeted structures
and properties.
Rapid advances in computer technology and an unprecedented increase in com-
puting power and data storage, combined with the ready availability of user-friendly mod-
eling and simulation software packages powered by extremely efficient computational
methodology, have facilitated modern modeling and simulation in chemical engineering.
It is now possible, for example, to begin with simple rules for interactions among particles
(electrons, protons, atoms) and simulate emergent behavior. On the other hand, such
emergent phenomena may be so complex that the computer model itself is an “equation-
free” model developed to capture the behavior on the basis of large volumes of experi-
mental observations (using such data analysis tools as ML algorithms; Kevrekidis and
Samaey, 2009). It is possible, therefore, that in the future, the distinction between first
principles–based and data-based modeling/simulation will become sufficiently blurred as
to become almost irrelevant. The following sections describe the role that the dual tools

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Tools to Enable the Future of Chemical Engineering 213

of modeling and simulation will play in chemical engineering in the future in the specific
areas of education, research, and industrial applications.

Education Applications

Since computers were first introduced into chemical engineering, (Katz, 1966),
simulation has been an integral part of chemical engineering education; this trend will
continue and will likely accelerate in line with the advent of faster and more sophisticated
computer hardware and software and transformative approaches (such as equation-free
models). The three traditional central components of simulation are (1) the mathemati-
cal/computational model or interaction potentials upon which the simulation is based, (2)
the algorithms used to carry out the simulation and/or solve the equations, and (3) the
computer hardware and software used to carry out the necessary computations and display
the results.
The chemical engineering classroom of the future will leverage modeling and
simulation to expand student understanding. In the past, student intuition was developed
in advanced undergraduate courses using simple canonical examples of chemical pro-
cesses that permit analytical closed-form solutions. Advances in simulation are already
making it possible to complement and expand these pedagogical approaches. For exam-
ple, molecular-scale phenomena, phase transitions, and fluid flow can all be taught effec-
tively using computer simulations. With virtually limitless computing power, the chal-
lenge is to convey this information in consumable form so that students understand the
limitations of simple models; develop intuition; and become adept at leveraging this in-
formation to design, control, and exploit the systems of interest. Simulation allows stu-
dents to learn through concrete examples how electronic interactions give rise to inter-
atomic and intermolecular interactions; how these interactions produce structure, proper-
ties, and behavior; and how changes at the nanometer scale manifest at the mesoscopic
and macroscopic scales.
The future will see improvements not only in the knowledge base upon which to
build the models but also in the ability to compile, validate, and deploy such information
rapidly and efficiently. Achieving the goal of conveying the information from simulations
in “consumable form” will be facilitated by synergistic integration of model development
and computation within unified simulation software, coupled with efficient graphical out-
put capabilities and user-friendly application programming interfaces (APIs). Different
parts of the chemical engineering curriculum will require different suites of modeling and
simulation tools (e.g., MATLAB/SIMULINK for process control, ASPEN for process de-
sign, computational fluid dynamics [CFD] tools for fluid dynamics, MD and Monte Carlo
simulation for thermodynamics). There is no question as to the increasing role of model-
ing and simulation tools in the education of chemical engineers of the future. The com-
mittee imagines a future in which modeling and simulation tools, in combination with AI
and virtual reality technology, will afford students the experience of operating an entire
chemical manufacturing plant virtually, in a manner that transcends even that which flight
simulators provide to pilots in training today.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

214 New Directions for Chemical Engineering

Research Applications

Chemical engineers have used simulation to contribute significantly to under-


standing of systems in which interaction potentials are developed to describe interacting
atoms, molecules, polymers, nanoparticles, and larger particles, and emergent collective
behavior is revealed. These approaches make it possible to predict the behavior of matter,
whether a catalyst for a chemical reaction or a material for a specific application. Elec-
tronic structure calculations are used at the smallest level, which can inform classic force
fields to access longer length and time scales. Chemical engineers have made leading
contributions using such approaches in a wide variety of systems. Examples in biomolec-
ular systems include protein folding and aggregation, hydration, and ligand–receptor in-
teractions; self-assembly of lipids; and the role of hydration in ion complexation. In sur-
factant systems, examples include prediction of equilibrium mesostructures used to
formulate personal care products or to template advanced materials. In nanotechnology,
the same simulation methods are used to predict the self-assembly of “colloidal mole-
cules” and patchy particles to form aggregates and assemblies for next-generation soft
materials or their directed assembly in the presence of biasing fields.
Mesoscale models, which exploit coarse-grained potentials or even continuum
free-energy functionals, are also used to access emergent behaviors (e.g., for protein ag-
gregation in lipid bilayers, or colloidal assembly in complex fluids at length and time
scales inaccessible to the most detailed simulations). The challenges for the future include
better connecting simulation from one scale to another and developing interoperable sim-
ulation tools to inform model development for the systems scale. The coarse-grained or
continuum-scale descriptions are valuable in that they include descriptors and parameters
that may be related to experimental observations or may be exploited in systems-level
mathematical models. Close coupling of these simulations with experiments can form a
virtuous circle of simulation, prediction, and identification/determination of mesoscale
parameters and experiments to improve understanding of complex systems.
Significant challenges remain in simulation of systems that are far from equilib-
rium. For example, it is currently challenging to move from the molecular scale, to the
mesoscale, to the cellular or tissue level in biomolecular engineering; to relate surfactant
mesostructure to rheology in systems under flow; to understand reacting systems with
changes in mesostructure and physicochemistry over prolonged time scales; or to under-
stand driven assembly of colloidal suspensions in complex soft matter. As simulation
costs decline and data-driven approaches mature, cross-talk between simulation to reveal
emergent behavior, mesoscale description, and mathematical models will advance. Such
advances will facilitate, for example, environmental modeling, making it possible to start
from detailed chemical reactions at the quantum level and integrate up to represent large-
scale, macrolevel behavior, which in turn will enable high-fidelity climate prediction over
periods of years.
Other rapidly developing tools that will be important for chemical engineering
include CFD, computational quantum chemistry, and predictive reaction kinetics. Im-
proved CFD models will contribute to understanding of turbulence. Direct numerical sim-
ulation (DNS) and large-eddy simulation (LES) for fluids and methods for two-phase

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Tools to Enable the Future of Chemical Engineering 215

flows, combined with faster computers, are yielding new insights. Computational quan-
tum chemistry has become invaluable to chemical engineers for modeling force fields,
electronic materials, catalysts, thermochemistry, and safety analyses. Chemical engineers
are deeply involved in developing methods and codes that leverage these calculations with
ML. Finally, tools of predictive reaction kinetics are advancing rapidly. Reactive force
fields such as ReaxFF and QMDFF and improvements in ab initio molecular dynamics
promise to enable model-based chemical experiments.
Chemical engineers can also meet the challenge of capturing and manipulating
the information from simulation and experiment in forms that provide mechanistic insight
to inform the design, control, and exploitation of the systems of interest. As research ad-
vances on this front, so will pedagogy, as these new approaches and capabilities will be
integrated into the classroom.
One other tool that could revolutionize chemical engineering research is the de-
velopment of interoperable models and software. Currently, there exist software for pre-
dicting properties, other software for materials design that use the properties, and yet dif-
ferent software for simulating the material and predicting its performance in end use.
Integrating these disparate systems within a framework that facilitates interoperability
would greatly speed up materials discovery and design.
With the growing power of AI and its adoption in chemical engineering, chemis-
try, physics, and materials science, molecular simulation is poised for a revolution. Stand-
ard approaches to developing interatomic force fields may become a thing of the past,
replaced by the use of ML to develop force fields (Chmiela et al., 2018; Noé et al., 2020).
Already there are a growing number of success stories. A recent study (Batzner et al.,
2021) shows that with the use of equivariant neural networks, massive datasets are not
required for deep learning, and that high accuracy can be achieved with far less computa-
tional effort than previously thought.

Industrial Applications

Chemical Manufacturing

Modeling and simulation have jointly played a significant role in the modern
chemical and biomanufacturing industries, although applications have focused primarily
on large-scale processes and equipment, mainly for operator training and for the design
and implementation of advanced control systems. The increasing operational complexi-
ties and shrinking economic margins in the petrochemical industry, coupled with stringent
environmental and quality demands on the manufacture of specialty chemicals and poly-
mers, have spurred increased use of modeling and simulation. While the pharmaceutical
industry currently lags behind the chemical industry, fundamental changes in regulatory
requirements are motivating greater use of mathematical models and simulation in that
industry, especially in the rapidly growing biomanufacturing sector. With the advent of
such Food and Drug Administration initiatives as process analytical technologies and
quality by design, an increasing number of leading pharmaceutical companies are incor-
porating process modeling and simulation systems into their manufacturing enterprises.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

216 New Directions for Chemical Engineering

One concept penetrating pharmaceutical manufacturing is that of the “digital twin,” de-
scribed most simply as a virtual representation that serves as the real-time digital coun-
terpart of a physical object or process. Park and colleagues (2021) provide a brief review
of potential bioprocess applications.
With increasing capacity for collecting, curating, warehousing, and visualizing
massive quantities of process data and with easy access to relevant ancillary data, the next
challenge is to develop systems that will integrate process operating data seamlessly with
ancillary data from the supply chain, policy, economic trends, global markets, and climate
predictions to optimize production planning, scheduling, and operation. This capability
will become especially indispensable for enterprises with multiple manufacturing facili-
ties distributed across the globe. Such modeling is also important for scaling up, especially
for large-scale plants that represent a large capital investment. The computer hardware
technology necessary to facilitate the implementation of such systems currently exists and
is advancing rapidly; the necessary software is sure to follow in the coming years.
Systems are currently being developed to allow the industrial practitioner to con-
nect molecular-level models to large-scale process models and insert process models
within the context of a facility. Consider as an illustrative example a dynamic model of a
CO2 capture unit within a power plant that is capable of predicting the plant’s performance
as the electricity output changes dynamically. Imagine that the systems-level model of the
power plant is integrated with the market to enable the user to determine optimal plant
configurations and product slates in a given market. Such a simulation system will, among
other things, allow users to explore options for how best to design and optimize energy
production/conversion systems. A beta version of such a system currently exists and of-
fers a glimpse into what is possible in the future (Arent et al., 2021).
To be most useful, however, such systems need to provide different kinds of in-
terfaces for different categories of users, making customized information available in real
time as appropriate for each user: technical information for the research engineer and the
practicing engineer, and high-level economic and business information and policy impli-
cations for decision makers. The committee envisages such systems also possessing the
capability to incorporate past experiences (e.g., failures, safety incidents). Chemical en-
gineers already feature prominently in, and in some cases have founded, companies that
are commercial vendors of process modeling, simulation, control, and operations soft-
ware, hardware, and services (e.g., the pioneering process design software ASPEN was
developed and the company ASPENTECH founded by Larry Evans, a professor of chem-
ical engineering at the Massachusetts Institute of Technology). Chemical engineers are
likely to contribute significantly to the development, implementation, and commerciali-
zation of these systems of the future.

Life-Cycle and Technoeconomic Analyses

Life-cycle assessment (LCA) and technoeconomic analysis (TEA) are systems-


level analysis tools used by chemical engineering primarily for new chemical technolo-
gies; they represent a completely different category of industrial applications of modeling
and simulation. As mentioned throughout earlier chapters, the ability to use these tools

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Tools to Enable the Future of Chemical Engineering 217

will enable chemical engineers to ensure that their innovations have a net lower environ-
mental impact relative to alternatives. LCA is fundamentally about the mass and energy
balances of processes and is a standardized methodology for assessing the environmental
impact of a product over its entire life cycle (i.e., production, use, and disposal) (Billiet
and Trenor, 2020; Muralikrishna and Manickam, 2017; Nuss, 2015; Weisbrod et al., 2016;
Zhang et al., 2018b); it can also be used to compare two or more products objectively.
LCA has a fixed structure and follows the standards in International Organization for
Standardization (ISO) 14040:2006 (ISO, 2006). It can be carried out as a

 cradle-to-grave analysis, which provides the full LCA from the raw material
extraction step, through the use step, to the disposal step (grave);
 cradle-to-gate analysis, which provides a partial product LCA from the raw
material extraction step to the factory gate (i.e., before the product is trans-
ported to the consumer); or
 cradle-to-cradle analysis, in which the end-of-life disposal step is a recycling
process.

TEA, a methodology for analyzing the technical and economic performance of a


process or product, combines engineering design, process modeling, and economic eval-
uation. It depends heavily on the technology readiness level (TRL) and includes goal,
scope, and scenario definition; cost estimation; market analysis; profitability analysis; and
results interpretation. Specifics on TRL titles, descriptions, tangible work results, and
workplaces for the chemical industry can be found in Buchner et al. (2018). As the TRL
increases, data availability and accuracy increase, and the TEA becomes more reliable.
Integrating LCA with TEA yields what is called environmental technoeconomic
assessment (Thomassen et al., 2019), most useful when further combined with energy and
mass balances, thermodynamics (Banholzer and Jones, 2013), government policies (e.g.,
potential carbon tax, packaging recycling content targets, greenhouse gas [GHG] ceil-
ings), social acceptance assessment, and project control structure. This overall integrated
assessment structure affects technology directions and needs to be incorporated into the
thinking of chemical engineers as they develop new products and processes in the circular
economy, discussed in Chapter 6 (see the example discussed in Zimmermann and
Schomäcker, 2017). Appropriate modeling, simulation, and data assimilation tools tai-
lored specifically to facilitate such large-scale systems analysis do not currently exist,
creating an area of opportunity for chemical engineers as they facilitate the transition from
the linear to the circular economy. AI will play an important enabling role in this regard
(McKinsey Sustainability, 2019), allowing chemical engineers to harness information
contained in large datasets to learn faster and more efficiently from highly complex sys-
tems. Specifically, AI can enable chemical engineers to

 design circular products and materials (via ML-assisted design processes for
rapid prototyping and testing),
 operate circular business models (via AI’s predictive capabilities from histor-
ical datasets), and

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

218 New Directions for Chemical Engineering

 optimize circular infrastructure (via improving the processes for sorting and
disassembling products and recycling materials).

Medical Applications

Chemical engineers have contributed significantly to understanding of the molec-


ular systems and mechanisms underlying many of the body’s physiological functions.
From signal transduction in cells (how signals are passed from the surrounding environ-
ment into the nucleus of a cell) to changes in gene expression in response to these signals,
ultimately resulting in the regulation of various biological processes, chemical engineers
have developed models that have facilitated the development of systems biology (Kinney
et al., 2019). In an earlier generation, Bischoff and his coworkers pioneered the use of
compartmental models to study the macroscale distribution of drugs in the human body
(Bischoff, 2015). Mathematical models of metabolic pathways combined with novel
measurement techniques have facilitated unprecedented high-fidelity analysis of human
metabolism (Lachance et al., 2021). At the physiological level, the recent development of
the artificial pancreas, an insulin delivery system for those with type 1 diabetes, is based
on modeling, simulation, and control mechanisms developed primarily by chemical engi-
neers in conjunction with clinicians (Dassau et al., 2017).
Understanding of human physiology is rapidly approaching a sufficiently quanti-
tative level to permit the development of multiscale simulators of the entire body, from
single cells, to tissues, to organs and organ systems, to the entire organism. Such systems
can then be used for a limitless number of applications, including clinical trials. These so-
called in silico trials will save a significant amount of money and time by allowing pre-
diction of the clinical outcomes of drugs at various stages of development on the one hand,
and on the other, facilitating the design and synthesis of complex biological drugs and
predicting the behavior in vivo. Such simulation systems can also be used to implement
personalized medicine, with the potential to revolutionize the practice of medicine and
drug development (Ogunnaike, 2019) (see Chapter 5).
The building blocks necessary to actualize this class of simulators already exist
in the form of single-cell models, multiscale biological system information, metabolic
analyses, carbon labeling, flux analysis, single-cell genomics, high-throughput screening,
and so on. With rapid advances in biological/medical knowledge, data-collection capabil-
ities, and computer hardware and software technologies, the development of such medical
simulation systems may occur sooner rather than later.

NOVEL INSTRUMENTS

Chemical engineers have had a transformative impact on instrument develop-


ment, especially in the establishment of fundamentals underpinning measurement and
characterization, in the development of hardware, and in early adoption. Analytical meth-
ods and visualization tools are two particular areas that have benefited from the participa-
tion of chemical engineers. These methods and tools have had an impact on chemical
engineering research and practice in two distinct ways. First, in some cases (e.g., super-

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Tools to Enable the Future of Chemical Engineering 219

high-resolution microscopy and single-molecule detection), the tools have improved sig-
nal-to-noise ratios and improved accuracy and/or precision, thereby making higher-fidel-
ity characterization possible. Similarly, advances in atomic-force microscopy and trans-
mission electron microscopy have provided new information about synthetic as well as
biological objects with unprecedented precision and accuracy. Second, improved speed
in genomics, proteomics tools, and flow cytometry has allowed more rapid and more com-
prehensive explorations of the parameter spaces of interest to the researcher. The impact
on biology and medicine has been particularly significant, since the high-throughput ca-
pabilities of these tools now make the routine collection of massive amounts of data a
practical means of addressing biological complexity.
The field of tool development offers opportunities for chemical engineers to con-
tribute to the development of next-generation instruments that will provide both funda-
mental and practical insights not possible today. For example, methods that track protein
abundance dynamically inside a cell will provide unprecedented information about cellu-
lar function. In contrast, current protein analysis methods are rather simplistic, with capa-
bilities limited to quantifying specific, predetermined proteins one at a time by antibody-
based methods or mass spectrometry. Even if current proteomic methods were to be en-
hanced to become high-throughput, the technological basis would still be such that anal-
ysis could be performed only ex situ, requiring the removal of proteins from their native
environment, with a two-fold undesirable consequence: some of the key information
about in situ characteristics would be altered, and some of the essential features of pro-
tein–protein interactions would be lost. For proteins for which information on dynamic
behavior is accessible, advances in visualization of such dynamic characteristics have
yielded unique insights into cellular transport and cytoskeletal networks. Another exam-
ple of tools that will have a significant impact on chemical engineering is dynamic in situ
super-high-resolution microscopy, especially electron microscopy. Current transmission
electron microscopy methods can provide only temporally averaged information.
The ability to track individual cells in real time in the body represents another
opportunity to obtain potentially game-changing insight into the body’s metabolism in
both healthy and diseased states. Current methods for assessing cellular composition can
provide only infrequent snapshots of macroscopic information at discrete points in time.
Histology remains the primary clinical mode of assessing the cellular composition of tis-
sues. Such assessment is rather limiting because it provides only a two-dimensional static
view of cellular composition. The ability to visualize cellular motion in the body in real
time will provide useful information about the immune response to infections, sepsis, and
metastasis, all of which involve the complex, coordinated motion of multiple cell types.
Tools for building materials with molecular-scale precision will open up new op-
portunities in nearly all fields associated with chemical engineering. With such tools,
unique structural features that are currently infeasible will be possible because the tools
will offer the precision and control of 3-D printing and allow building of materials at the
individual-molecule scale, making it possible to control functional attributes and identify
and address manufacturing defects in real time.
“On-chip” systems offer a cost-effective means of generating information, syn-
thesizing materials, and screening them for their important interactions. Examples of such

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

220 New Directions for Chemical Engineering

systems include a chemical-plant-on-a-chip, human-on-a-chip, organ-on-a-chip, or col-


loidal-particle-generator-on-a-chip. (Opportunities for cell-, organ-, and organism-on-a-
chip are discussed in Chapter 5.) Such chip-based methods offer many key advantages.
From a synthesis point of view, they offer the advantage of small volumes and subse-
quently reduced costs. This attribute is helpful especially for research operations that aim
to generate large libraries of materials in small quantities (for example, chemical librar-
ies). Reduced scales are especially significant if the chemicals involved are toxic, since
the on-chip systems use very low volumes of material. Such on-chip systems are also
beneficial for colloidal synthesis. Traditional macroscopic methods of colloidal synthesis
are often subject to a high degree of heterogeneity associated with interfacial phenomena.
On-chip methods are very precise and have been used to synthesize colloids with more
uniform dimensions. Such on-chip systems have also played an important role in the de-
ployment of sensors that provide finer control over the flow and interactions of samples
and reagents. Human-body-on-chips, which include systems comprising cellular chips
representing one or more organs, are also poised to make a strong impact on drug devel-
opment.
All on-chip systems share some common features and opportunities for impact
from chemical engineering. First, microfluidics is a central component of all on-chip sys-
tems, and many of the fluids of practical interest (e.g., blood in the case of body-on-chips,
or colloidal suspensions in the case of nanoparticle synthesis) are complex. Consequently,
an understanding the flow of complex fluids in confined spaces is of prime importance.
Second, all on-chip systems consist of a complex interconnection of a number of individ-
ual components, each complex in its own right. The design, analysis, and effective de-
ployment of these systems offer opportunities for chemical engineers to contribute their
expertise in fluid dynamics, reaction engineering, tissue and cellular engineering, and pro-
cess systems engineering.

SENSORS

Sensors for Process Monitoring

Sensors play an important role in the design, development, and monitoring of


processes and systems in all areas of chemical engineering. Development of “smart sen-
sors,” “ubiquitous sensors” (tool or enabling technology depending on specific use),
“smart actuators,” and “smart transmitters” hold the potential to transform a wide variety
of manufacturing applications in the future. Such sensors can be classified broadly into
those that measure physical parameters (e.g., temperature or volume), chemicals (e.g.,
reactants or cell culture media), or states (e.g., cell viability or particle sizes).

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Tools to Enable the Future of Chemical Engineering 221

Physical Parameters

Sensors that measure such physical parameters as temperature, pressure, or flow


rate have long been central to process monitoring in the chemical industry. With the ad-
vent of inexpensive, lightweight, and wearable types enabled by the development of flex-
ible electronics, sensors that monitor physical parameters have found widespread appli-
cations. For the most part, sensors that monitor physical parameters are sufficiently
mature and have already been incorporated into many practical applications. A key re-
maining opportunity moving forward lies in real-time analysis of data from such sensors
for advanced process and product optimization. An emerging opportunity is in the adop-
tion of these sensors and sensor networks for remote operation and remote laboratories.
While remote operations have been used in the past largely for processes with safety con-
cerns or with inherently limited access, the practical limitations imposed by COVID-19,
which made remote operations inevitable, have brought this technology to the forefront
not only for manufacturing processes but also for instruction in laboratory courses, repre-
senting an area of potential future impact for chemical engineers.

Chemicals

Sensors with which to measure the composition of chemicals play an important


role in collecting molecular-scale information about a monitored system. For example,
chemical analysis of reactants/products has long been used to assess reaction progress.
Decades of advances in analytical tools have provided an array of methods (e.g., nuclear
magnetic resonance [NMR], Raman spectroscopy, high-performance liquid chromatog-
raphy, gel permeation chromatography, mass spectrometry [MS]) for assessing the mo-
lecular composition of complex chemical mixtures. Two key opportunities exist for the
future. First, while the ability to measure (sense) chemicals in large laboratory equipment
is practically unlimited, capabilities to do so using small sensors that can be readily in-
serted into the system for continuous monitoring are limited. Such sensors often require
the development of new sensing mechanisms, robust hardware to implement those mech-
anisms, and novel technologies to enable continuous measurements in a seamless way. In
addition, sensors for specific molecules (e.g., protein glycan groups) may require the de-
velopment of molecularly and/or biologically based sensors interfaced with an appropri-
ate device to generate the required measurement in the form of an electronic signal.
A second opportunity is the use of AI and ML to extract more meaningful infor-
mation from complex data. Measurements offered by analytical tools such as NMR and
MS are rich in information, and the information extracted from these measurements is
heavily dependent on the question posed. The availability of new AI/ML algorithms can
facilitate the extraction of new information from existing data, offering the opportunity to
develop means by which new tools can extract information from preexisting data.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

222 New Directions for Chemical Engineering

State

Another emerging opportunity lies in the capability to sense a state, which is in a


relatively early stage of development. In this context, the “state” of a process or system
refers in general to characteristics of the process or system that otherwise cannot be cap-
tured in a simple single parameter—for example, the viability of a phenotype of cells for
application in protein or cell manufacturing, or particle-size distribution in a granulation
or agglomeration process. Sensors capable of going beyond the current acquisition of
physical parameter measurements have played and will continue to play a major role in
the monitoring of chemical and biochemical processes. Future opportunities lie in the de-
velopment of sensors for measuring the abundance of specific molecules and the estab-
lishment of a network of such sensors supported by advanced data analytics tools to enable
optimum extraction of information contained in the massive quantities of acquired data.
The development of self-learning, automated, self-monitored reactors is yet another op-
portunity that will depend critically on the availability of smart sensors.
Although significant progress has been made in real-time monitoring for quality
control during production in dissolution, crystallization, drying, and other important unit
operations, the implementation of similar monitoring during cell growth and protein pro-
duction is lagging. For example, most antibodies and other therapeutic proteins require
posttranslational modifications, such as glycosylation and disulfide bond formation. The
product profile approved by the regulatory agency includes a range of allowed modifica-
tions typically not determined until postpurification. In addition, the extent of other, un-
desired modifications—such as oxidation, aggregation, charge variants, and truncation—
needs to be determined and quantified because these modifications affect drug perfor-
mance in end use. Two approaches for addressing these product characterization needs
are (1) modifying existing analytical methods and using data science methods to deter-
mine the composition of complex mixtures (e.g., cell culture supernatant), and (2) devel-
oping altogether novel analytical optical and bioelectronic sensors. For example, infrared
(IR)–based sensors are often used for measuring chemical composition, primarily because
they can be used in situ for the most part. However, using Fourier transform infrared
spectroscopy or near IR requires offline sensing and spectral deconvolution. Determining
the level of contaminants, such as heavy metals, in a bioreactor feed is an important aspect
of biomanufacturing. The development of sensors capable of providing these critical
measurements requires not only development of the actual hardware of the sensor but also
establishment of the fundamental parameters that define the state. Correct placement of
sensors is another practical challenge that needs to be addressed. As new sensors are de-
veloped, proper placement within a network and the connectivity within the configuration
will be central to successful implementation. The availability of transmitters to transmit
the necessary data within the network with the requisite resolution will be critical. The
development of such tools will require advances in hardware related to wireless transmis-
sion, as well as data analytics tools for efficient and effective information extraction.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Tools to Enable the Future of Chemical Engineering 223

Sensors for Health Monitoring

Design for Biological Applications

In the field of health monitoring, care for patients with diabetes has been impacted
by sensor innovation through the development of continuous glucose monitoring technol-
ogy. Monitoring of blood glucose levels, which, decades ago, depended on infrequent
laboratory-based sampled measurements, is now carried out virtually continuously with
wearable sensors (Lipani et al., 2018). This development has been made possible by sig-
nificant advances in glucose sensing chemistries, a deeper understanding of the biocom-
patibility of sensors, and the development of electronics that enable a user-friendly patient
interface. Such developments have transformed the sensing paradigm from “sample to the
sensor,” to “sensor to the sample,” to “sensor on the patient.” Sensing technology was
once based on large analytical equipment, requiring that the patient sample (e.g., blood)
be withdrawn and shipped to a central analytical laboratory (e.g., laboratory-based glu-
cose measurements). As sensors became smaller and portable, they went where the sample
was (e.g., home-based glucose tests), which eliminated the sample shipment process. As
sensors get even smaller, they can go directly on the patient (e.g., wearable glucose sen-
sors), thus eliminating the step of collecting the sample and putting it into the sensor.
These continuous sensors utilize a needle-based sensor inserted subcutaneously to meas-
ure the glucose concentration in the extracellular space just under the skin. Efforts have
also been made to develop completely noninvasive means of measuring glucose concen-
trations using spectroscopic methods or noninvasive collection of tissue fluid through the
skin, including sweat-based sensors. Sweat-based sensors measure the composition of
various analytes present in the sweat (including such small molecules as glucose and hor-
mones, and such large molecules as proteins) and use these measurements to assess the
patient’s physiological state (Chung et al., 2019). Recent advances in sweat-based sensors
are fueling optimism that the development of means of measuring additional analytes,
including hormones, toxins, and allergic responses in the body, will likely have a similar
transformative impact on health.
In addition to molecule-based diagnosis, wearable sensors now routinely perform
a variety of other diagnostic measurements, such as temperature, blood oxygenation, and
pulse (Gao et al., 2019). The ability to measure these key vitals through a wearable sensor
has already transformed the collection of human physiological data, facilitating unprece-
dented acquisition of information about human responses in real-life situations in real
time and continuously. Traditionally, these measurements were possible only using large,
bulky sensors often connected to stationary electronic processors and displays.
Therapeutic drug monitoring is another area in which the impact of novel sensors
can be transformative. Current drug-dosing regimens are rudimentary, based largely on
average male body mass and drug trials that ignore the effects of race, ethnicity, gender,
and many other patient-specific characteristics. The ability to measure drug concentra-
tions in the body in real time can allow clinicians to achieve tighter control over

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

224 New Directions for Chemical Engineering

pharmacokinetics, thus potentially reducing harmful variability in drug distribution and


preventing adverse events due to large excursions from desired concentration levels. Suc-
cessful implementation of such technology will require the development of drug-specific,
nonfouling, long-term implantable sensors that provide real-time feedback to clinicians.
Significant progress is already being made in this regard. For example, an aptamer-based
sensor that can measure certain chemotherapeutic agents has been developed and tested
at the preclinical level. The challenges associated with expanding the scope of such tech-
nologies for a broad range of drugs offer future opportunities for chemical engineers.

Data Handling and Processing

The ability to extract new, better information from existing data promises to have
significant impacts on health and medicine. As one example, consider medical diagnosis,
which has long relied on qualitative analysis of images derived from magnetic resonance
imaging (MRI), CT scans, radiology, and histology. Tools that enable better image pro-
cessing can extract better information and yield more accurate diagnoses compared with
qualitative human analysis. Studies have already demonstrated that AI-based analysis of
histology is likely to be more accurate and less likely to miss a rare event. At the same
time, AI-based analyses and diagnoses can help eliminate human bias in such procedures
(although if not implemented correctly with appropriate training data, AI can itself am-
plify biases).
Extrapolation of current trends into the future suggests an ever-increasing capa-
bility of wearable sensors to collect even more massive amounts of data about human
behavior in healthy and diseased conditions. These data will certainly include such basic
vitals as temperature, pulse, and oxygenation, which in themselves are sufficient for cer-
tain applications. However, the available data could contain real-time measurements of
such quantities as blood glucose, hormones as an indication of stress, and alcohol as a
measure of intoxication, among others. Some of the technical challenges to overcome in
realizing such capabilities include development of sensors that can work efficiently with
small amounts of analytes that are available in sweat and determine compositions with
appropriate accuracy.
A challenge also exists in correlating these measurements accurately with blood
concentrations, in particular with respect to the transport of analytes from the blood into
the interstitial fluid. An opportunity exists to collect massive amounts of data and establish
correlations across the population in a way that is simply not possible through the collec-
tion of small amounts of data. Can model-based analysis be used to predict catastrophic
physiological events from these data? Can drug reactions be better understood or pre-
dicted through appropriate analysis of such massive amounts of data? Given the well-
documented challenge posed for drug design and development by the heterogeneity of
drug response across a population, the prospect of personalization of drug therapies pre-
sents an exciting opportunity for reducing adverse events without compromising thera-
peutic efficacy. However, actualizing this concept will require having adequately in-
formative data available to support such personalization at the design stage and the

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Tools to Enable the Future of Chemical Engineering 225

reduction of adverse events at the follow-up stage. Continuous physiological monitoring


offers the potential for producing precisely this sort of data.

CHALLENGES AND OPPORTUNITIES

The development of tools that synthesize available data in real time and frame-
works or models that transform data into information and actionable knowledge could
become one of the key contributions of chemical engineering to society over the next
decades. It is easy to imagine a not-too-distant future characterized by data-on-demand—
where data on anything, at any level of granularity, will be readily and instantly accessi-
ble. Such a future suggests profound and exciting opportunities for chemical engineers,
who are trained in process integration and systems-level thinking—skills that will be re-
quired to synthesize disparate data streams into information and knowledge.
The systems thinking, analytical approaches, and creative problem-solving skills
of today’s chemical engineering graduates give them a distinct advantage in using AI in
real-world contexts. The evolution of AI in the next decade will have enormous implica-
tions not only for the types of problems chemical engineers will be able to solve but also
for how they will develop those solutions. Chemical engineers are poised to contribute
significantly to the development of modeling and simulation tools that will influence ed-
ucation, research, and industry. They will continue developing and disseminating meth-
ods, algorithms, techniques, and open-source codes, making it easier for nonexperts to
use computing tools for scientific research.
The increasing operational complexities and shrinking economic margins in the
petrochemical industry, coupled with stringent environmental and quality demands on the
manufacture of specialty chemicals and polymers, will continue to drive the increased use
of modeling and simulation. While the pharmaceutical industry currently lags behind the
chemical industry in its use of simulation tools, fundamental changes in regulatory re-
quirements are motivating its greater use of mathematical models and simulation, espe-
cially in the rapidly growing biomanufacturing sector.

Recommendation 8-1: Federal and industry research investments should be directed


to advancing the use of artificial intelligence, machine learning, and other data sci-
ence tools; improving modeling and simulation and life-cycle assessment capabili-
ties; and developing novel instruments and sensors. Such investments should focus
on applications in basic chemical engineering research and materials development,
as well as on accelerating the transition to a low-carbon energy system; improving
the sustainability of food production, water management, and manufacturing; and
increasing the accessibility of health care.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

9
Training and Fostering the
Next Generation of Chemical Engineers

 Chemical engineers are in high demand across most professions and job levels,
and chemical engineering provides an excellent foundation for many career paths.
 The undergraduate chemical engineering curriculum has served the discipline
well and has continued to evolve in response to scientific discoveries, technolog-
ical advances, and societal needs.
 Educational attainment in chemical engineering for women and Black, Indige-
nous, and People of Color (BIPOC) individuals has remained essentially un-
changed for more than a decade.

The chemical engineering curriculum today provides a robust foundation of tools


and practices founded in an understanding of systems and molecular-level phenomena,
including fundamental concepts of mass and energy balances, transport phenomena, ther-
modynamics, reaction engineering, control, and separations. Although the core subjects
of the curriculum were first built around manufacturing processes, primarily petrochemi-
cal, they can be applied in most fields and professions. Indeed, previous chapters have
highlighted how these concepts can be and have been applied across a wide variety of
applications in energy, environmental sustainability, health and medicine; manufacturing;
and materials development. As a result, chemical engineers are in high demand across
most professions and job levels, and chemical engineering provides an excellent back-
ground for many career paths (NAE, 2018; see Figure 9-1).
In the past, chemical engineers tended to find industrial employment in manufac-
turing and process engineering. The connections between basic research and the work-
place were usually made through industrial research and development (R&D) organiza-
tions that were aligned with internal business units with needs for both operational
efficiency and new products or process developments. Chemical engineers would often
begin a career in R&D or manufacturing and move, over time, into senior technical and
business leadership positions. Faculty members would engage with industry as consult-
ants and through university–industry collaborations. The net result of that model was a
feedback loop from the market back to basic research at universities.
More recently, a strong shift in academic research topics in the field has occurred,
driven primarily by changes in federal funding priorities, leading to a movement away
from process research and toward basic and applied scientific research. This discovery-
focused research includes areas, such as materials and life sciences, relatively new to
chemical engineering. At the same time, many companies have globalized, shortened their
time horizons, and reduced or eliminated longer-term research programs and laboratories.

226

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Training and Fostering the Next Generation of Chemical Engineers 227

The resultant growing gap between university research and market needs is often referred
to as the “valley of death.”
During the past decade, the world has undergone a major technology-led transi-
tion enabled by global networks, computing power, sensors, artificial intelligence, and
machine learning. This transition will continue and likely accelerate, creating an ongoing
need for new skills and capabilities. As technology continues to transform how work is
performed, a growing need for collaboration skills—in communication, interdisciplinary
teamwork, and project management—can be anticipated, as can educational needs yet to
be identified. And all of these trends will require lifelong learning.
This chapter examines the current state of chemical engineering education, in-
cluding the broader context of the existing academic education model (Box 9-1) and the
value of the current undergraduate core curriculum. The committee proposes strategies
for growing and diversifying the profession—both of which are essential to the field’s
survival and potential for impact—by making it more broadly accessible.1 Following a
discussion of the aspects of undergraduate and graduate education that will need to change
to prepare the next generation of chemical engineers, the chapter turns to emerging trends
that are shaping new models of learning and innovation for the future.

FIGURE 9-1 Career paths available to chemical engineers in a range of industries, shown here
with median salary in industry categories with at least 30 respondents to an American Institute of
Chemical Engineers salary survey. SOURCE: AIChE (2021).

1
E.g., https://www.aiche.org/chenected/2021/02/ideal-path-equity-diversity-and-inclusion.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

228 New Directions for Chemical Engineering

BOX 9-1
The Existing Academic Education Model

Academic institutions are tasked with three important societal responsibilities—education,


research, and service—whose fulfillment has long brought, and will continue to bring, tremen-
dous societal benefit. A simple “mass and energy balance” of U.S. universities is shown in the
figure below.
Undergraduate education and graduate research and education are two distinct but con-
nected functions of research universities. Undergraduate programs collectively receive more
than $100 billion in tuition and state support for public institutions, and produce about 2 million
college graduates each year (Atkinson, 2018; Kastner, 2018; NCES, 2021; NCSES, 2020; The
Pew Charitable Trusts, 2019). More than 30 percent of the U.S. population over age 25 has a
college degree (U.S. Census Bureau, 2019b), putting the United States among the top coun-
tries—but not at the top—in the number of college graduates per capita. The increasing costs of
and inequities in access to education are major challenges that require attention to ensure that
universities meet the societal needs of the future. In this regard, the widespread use of remote
teaching during the COVID-19 pandemic may offer models for educational practice, while po-
tentially improving access and decreasing cost.

A “mass and energy balance” of U.S. research universities showing the flow of funding streams
and various outputs of graduates and research results. Data from Atkinson (2018), Kastner
(2018), NCES (2021), NCSES (2020), The Pew Charitable Trusts (2019).

Graduate students are less numerous than undergraduates. Annually, about 700,000 stu-
dents graduate with a master’s degree, and about 35,000 with a PhD; about 1 percent of the U.S.
population holds a PhD (NCSES, 2018; U.S. Census Bureau, 2019a). Since the mid-1970s,
graduate research in STEM (science, technology, engineering, and mathematics) fields has been
funded primarily by federal grants, although the relative fiscal contributions of philanthropy and
industry to academic research are growing at some institutions.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Training and Fostering the Next Generation of Chemical Engineers 229

Chemical engineers with PhDs have driven research and development (R&D) in a variety
of fields, including energy, water, food, health, and biotechnology. In the field of biotechnology,
they have played a significant role in advances that include antibody design, biologics manufac-
turing, cell therapies, nanotechnology, nucleic acid therapeutics, and tissue engineering, among
others. Collectively, the biotechnology/pharma industry accounts for a productivity of just over
$1 trillion/year—an achievement not possible without the graduate education provided by uni-
versities. A fraction of graduating PhDs in chemical engineering pursue careers in academia and
support the growth of the undergraduate education enterprise. These two career paths together
support a clear and important connection between graduate education in the field and society at
large.
Generally speaking, PhD programs in chemical engineering focus on teaching graduate stu-
dents how to conceptualize and carry out research by using an individual research project as a
training ground. The output at the end of the degree program encompasses both trained research-
ers and the products of their research (frequently in the form of published papers). A third key
outcome of graduate research and education is invention, captured tangibly in patents. U.S. uni-
versities account for about 6,000 granted U.S. patents each year, representing about 4 percent
of the total annual patent output in the United States (NSB, 2018; USPTO, 2021).
Recent years have seen a growing interest among both students and faculty in championing
the translation of academic ideas into practice. There are many successful examples of commer-
cialization of academic inventions driven by the entrepreneurial enthusiasm and expertise of
their inventors. However, sole reliance on the business skills of individual inventors signifi-
cantly limits the efficiency of translation and will certainly leave many important inventions
behind. Thus it is important for universities to complement and collaborate, rather than compete,
with industry. Academic–industry consortia provide a natural opportunity not only to support
academic research but also to bring inventions into the marketplace and bring professional man-
agers to spin-off companies. Universities can improve translational efforts by accepting and
recognizing such endeavors and measuring their impact; by offering stronger and longer protec-
tion of intellectual property, thus allowing technologies to mature; and by investing in inventions
emerging from their programs to develop and derisk them prior to licensing, encouraging inven-
tors to address translational hurdles, and incorporating translation and entrepreneurship into the
education they provide. Such efforts can provide additional avenues for researchers to pursue
societal impact, enable universities to build stronger bridges with industry, and allow both uni-
versities and society to extract more value from a currently underutilized resource.

THE UNDERGRADUATE CORE CURRICULUM

Throughout its history, chemical engineering has been defined as a profession by


its core undergraduate curriculum, a curriculum that has for more than a century prevented
the “spalling of the profession” (p. 573, Scriven, 1991). At the same time, this core un-
dergraduate curriculum has evolved with the incorporation of new topics reflecting
emerging areas of impact and relevant practice, as well as ongoing dialogue about the
very nature of the profession. The resilience of the discipline in the face of change reflects
the nature of its core curriculum and how it brings together the underpinning sciences
(chemistry, physics, mathematics, biochemistry, and biology) into an interdisciplinary
problem-solving context. Together, the enduring nature of this canon and its history of
adaptation and impact speak to the resiliency of the chemical engineering discipline.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

230 New Directions for Chemical Engineering

This curriculum has been aimed at transmitting a body of knowledge that is foun-
dational and translating it into solutions for technological and societal challenges. It has
enabled chemical engineers to respond nimbly to the unpredictable nature of these chal-
lenges by redirecting fundamental concepts as codified in a historical sequence of core
courses: mass and energy balances, transport of mass and energy, chemical kinetics, pro-
cess design and control, and a capstone undergraduate laboratory. The examples used in
transmitting this knowledge and applying it in practice have changed and will continue to
do so as chemical engineering finds new applications, even as the foundations have ac-
quired additional complexity through mathematical dexterity fueled by transformational
changes in computing power, as well as greater breadth through the growth of the bio-
chemical aspects of the discipline.
The core undergraduate curriculum provides a problem-solving approach to the
mastery of concepts in the dynamics and thermodynamics of physical and chemical pro-
cesses, with a historical evolution from the physical to the chemical; most recently to the
biochemical and electrochemical, and at present, toward the photochemical realm. The
problems addressed have changed because “engineers solve problems. If they are suc-
cessful, those problems disappear. Then we find new problems to solve…, but the princi-
ples used to solve successive generations of problems change very slowly…” (p. 243,
Schowalter, 2003).
The core curriculum as taught represents a method of inquiry and a toolbox for
solving problems. It is entirely general in its most abstract mathematical form; perhaps
for this reason, it has remained useful for a remarkable breadth of relevant practice. At
the same time, however, it can appear to lack merit and utility when first learned, espe-
cially if the content is delivered absent context within current modern problems and prac-
tice. The consensus of a selected group of graduate students and postdocs is that “process
science,” their nomenclature for the core curriculum, has proved complete enough and
adaptable enough to persist for the next 25 years and longer (Westmoreland and McCabe,
2018). Yet the profession and its undergraduate curriculum do not always succeed in cre-
ating the messaging landscape required to attract and retain individuals with the diverse
backgrounds and interests that future challenges will demand. As it evolves, then, the
curriculum will need to convey with greater purpose and success how “no profession un-
leashes the spirit of innovation like engineering” and how few other disciplines “have
such a direct and positive effect on people’s everyday lives” (p. 46, NAE, 2008).
To those embedded within the field at a given time, the evolution of the curricu-
lum has often seemed too slow and the survival of the discipline fraught with perils. This
is not a new perception. Decades ago, Denn (p. 565, 1991) observed, “We have been
hearing a great deal in recent years about the changing nature of chemical engineering.
The emphasis on new fields of research has created the appearance of a fragmented pro-
fession….” Nonetheless, as suggested above, the curriculum “has endured, not because it
is frozen, but because it has adapted dynamically to new ideas, emphases challenges, and
opportunities” (p. 7, Luo et al., 2015). As described later in this chapter, some topics
within the curriculum will need to evolve more rapidly, and in some cases, components
removed from the canon during earlier cycles of evolution will need to be restored.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Training and Fostering the Next Generation of Chemical Engineers 231

Today, as throughout its history, the field of chemical engineering needs to con-
sider what minimal requirements and what set of principles should define the content of
its core undergraduate curriculum. This is a question perhaps best posed as: What should
an evolving undergraduate curriculum deliver as its product? Nearly 20 years ago, the
answer to that question was, an undergraduate chemical engineer trained through “a pro-
gram of study sufficient for entry-level positions in engineering practice and engineering-
related fields,” but also exposure to shoulder areas and to professional and personal ethical
guidelines and the foundational knowledge required for graduate studies (p. 243,
Schowalter, 2003). These requirements have not changed, but chemical engineers func-
tion and practice in a much different environment today. They are asked to address more
diverse challenges with a body of knowledge and a toolbox that extend beyond what a 4-
year core undergraduate curriculum can competently deliver in full. As discussed later in
this chapter, master’s degrees and continuing education are likely to become increasingly
important for working professionals who seek to specialize in one of these shoulder areas.
The toolbox delivered by the undergraduate curriculum provides a mathematical
framework for designing and describing (electro-/photo-/bio-) chemical and physical pro-
cesses across length and time scales spanning many orders of magnitude. It teaches that
(1) some quantities are conserved (energy, momentum, atoms, mass); (2) their balances
need to be carried out over “control volumes” small enough to be homogeneous but large
enough to be described by continuum equations; (3) thermodynamic relations define the
point of equilibrium, but also the dynamics by which systems approach such equilibria,
whether through chemical or physical changes; and (4) all of this extends, remarkably
unperturbed, to molecules and atoms in every state of matter (gas, liquid, solid, supercriti-
cal).
A survey of young professionals a few years after they had entered the profession
of chemical engineering identified features of the undergraduate curriculum that they con-
sidered important to their careers (Figure 9-2); the components of the enduring core cur-
riculum are well represented throughout these features. The four highlighted items repre-
sent those in need of revision and strengthening in the face of changes in both the nature
of chemical engineering practice and the employment landscape. Two items in particu-
lar—process and product safety; and data science and application: design of experiments,
statistics, analytics—reside within the core knowledge base and need to become more
prominent. Process and product safety will need to become a stronger component within
each core undergraduate course. Data science and statistics may be delivered most effec-
tively in a separate course embedded within the core curriculum and taught with specific
emphasis on matters of chemistry and engineering. This latter course would also bring a
greater emphasis on statistics in the modern context of larger datasets, more powerful
computing, and models and methods that are more robust and of greater fidelity.
The other features highlighted in Figure 9-2—business skills, leadership training,
management, and economics; and innovation and entrepreneurial skills—represent
“softer” skillsets that provide entry-level engineers with significant competitive ad-
vantages in today’s workplace. Along with other, related skills, such as written and oral

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

232 New Directions for Chemical Engineering

FIGURE 9-2 Career importance of various areas of study, as indicated in a survey of early-career
chemical engineers. SOURCE: Modified from Luo et al. (2015).

communication and a baseline understanding of policy and regulatory issues, however,


they reside outside the technical core of chemical engineering. In the committee’s view,
the core undergraduate curriculum is not the most effective vehicle for delivering the nec-
essary foundational knowledge and skills in these areas. Elective courses, postgraduate
training in specialized industrial settings, and lifelong learning through professional soci-
eties and relevant literature provide more effective routes for acquiring and sharpening
these skills, and the application and sharpening of these ancillary skills can be made part
of each core course, with emphasis on how they enable and enhance the technical contri-
butions of chemical engineers.
The remainder of this section describes some of the challenges that represent im-
portant considerations in the near-term evolution of the undergraduate curriculum, as
identified by members of this committee and shared by invited external speakers in dis-
cussions and presentations. Three challenges are discussed: the need for experiential
learning and greater connectivity among the concepts/tools of the discipline and their ap-
plication in practice through (1) more effective connections among the individual core
courses (“the silos”); (2) experiential learning through virtual or physical laboratory ex-
periences earlier in the undergraduate course of studies; and (3) a more effective and
seamless embedding of statistics and of mathematical and computational thinking into the
core. The committee emphasizes that these challenges are inextricably connected, and
notes that actions suggested in the course of the discussion are meant to be illustrative,
and not to prescribe modes of implementation, which will best be identified by experts in
discipline-based education research (e.g., NRC, 2012; Paul and Brennan, 2019).

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Training and Fostering the Next Generation of Chemical Engineers 233

Connecting the Silos

In the core curriculum, concepts of balances (mass, energy, momentum), fluid


flow, thermodynamics, kinetics, and process/design control are taught in separate courses
that may make them seem disconnected to students. Students are then asked to connect
these seemingly disparate concepts with each other, as well as with the bond-making and
bond-breaking rules from the chemistry curriculum, the biochemical tenets of the biology
curriculum, and the mathematical mechanics taught within the mathematics curriculum.
Not surprisingly, students face significant hurdles in bringing these historical repositories
of knowledge together to form the problem-solving and reasoning strategies that consti-
tute the practice of chemical engineering. Some of these connections emerge, with varying
levels of effectiveness and rigor, in a laboratory or unit operations course near the end of
the curriculum. But these connections are better made by anticipating in earlier courses
how these concepts will ultimately coalesce at later stages and what kinds of engineering
challenges require their combined application—how the balances, the thermodynamics
and dynamics, the fluid flow, and the chemistry and biology ultimately merge into the
design and control of reaction, separation, biological, and materials synthesis processes.
These earlier connections can be made, in the committee’s view, by making the bounda-
ries between the silos more porous, highlighting how the individual core concepts first
presented within their respective silos ultimately merge into the practice of chemical en-
gineering.

Experiential Learning and a “Laboratory within Each Core Course”

The dense nature of the core undergraduate curriculum leaves few openings for
incorporating an additional hands-on laboratory course earlier in the curriculum. In some
instances, this has been successfully accomplished, albeit as a broad engineering design
course (e.g., the Coffee Lab at the University of California, Davis; see Box 9-2). In other
cases, a freshman-level introductory course has attempted to place the curriculum in con-
text at an early stage, but without the rigor of analysis and treatment that will follow later
on and with some duplication of the content of subsequent core courses. Such efforts need
to continue and expand, but they are likely to miss those students that enter at a later stage
of the curriculum through transfer from community college or other majors. An alterna-
tive strategy would be to use advances in real-time simulation and demonstration in vir-
tual/digital form to illustrate an “experiment” representing the behavior of a (bio-/photo-
/electro-) chemical or physical system as described by a mathematical representation im-
mediately following this “experience.” In its interactive mode, this kind of visualization
would allow students to design and control the behavior and performance of such systems,
to explore how they respond to perturbations in parameters or conditions, and to address
and resolve safety and ethical matters in the practice of engineering without risking any
direct physical or professional consequences of their actions. Such approaches, currently
implemented as more ad hoc strategies, would expose students to issues of design, control,

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

234 New Directions for Chemical Engineering

BOX 9-2
Coffee Lab: An Example of Experiential Learning

At the University of California, Davis an experiential freshman laboratory experience has


been developed for both majors and nonmajors.a Each version of this course uses the design of
a coffee-brewing process, optimizing flavor with respect to energy use, as a vehicle for experi-
ential learning. For nonmajors, the goal is providing a nonmathematical introduction to how
chemical engineers think by using the unit operations of coffee brewing and experiments as an
illustration. This course for nonmajors has a very large enrollment (>1,000 students/year) and
serves the goals of both marketing the major to non–chemical engineering freshmen and provid-
ing a basic chemical engineering skillset to other disciplines. The analogous course for majors
is mathematical in nature and uses the roasting, grinding, and brewing of coffee to introduce
process flow diagrams, mass and energy balances, transport phenomena, separations, and the
basics of design. With this experience in hand, students are demonstrably more sophisticated in
their understanding of how individual courses and concepts fit into the discipline and various
applications as they complete their degrees.
The Coffee Lab also uses relatively inexpensive and readily available household appliances
that require minimal training and safety measures. This makes it an accessible introduction to
chemical engineering for institutions that do not currently support a chemical engineering pro-
gram or laboratory and could facilitate interest and preparedness for students who transfer to
schools with, or pursue graduate degrees in, this field.
a
https://coffeecenter.ucdavis.edu/facilities/undergraduate-coffee-lab.

safety, and even economic impact that typically appear in their more formal and founda-
tional contexts later in the curriculum. They could also be used to incorporate sensitivity
and statistical analyses, design of experiments, “what if” assessments of realistic scenar-
ios, and a view of mathematical treatments underpinned by their role in analysis and de-
cision making in the practice of the chemical and physical processes that such treatments
intend to describe. The application of these approaches will continue to benefit from ad-
vances in real-time description and visualization of process (and product) performance,
and will sharpen the process synthesis and analysis skills needed for chemical engineering
practice. And students will be more effective and feel less intimidated when examining
the validity of assumptions made in describing real-world systems that would otherwise
require descriptions too complex for analytical solutions or even numerical analysis of
more complete equations.

Bringing Mathematics and Statistics into the Core

In most cases, students acquire the mathematical machinery of calculus, differen-


tial equations, and linear algebra in courses taught by college math departments or through
advanced high school coursework. In such courses, they acquire limited (if any) skills in
numerical methods or in the construction of the equations that describe the physical and

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Training and Fostering the Next Generation of Chemical Engineers 235

chemical content they are later asked to implement in chemical engineering courses. Sta-
tistical analysis, specifically in the context of acquiring knowledge and analyzing dense
datasets, is essentially absent from the curriculum until the capstone laboratory course,
where students first encounter the imperfections of the data gathered through their own
actions. Previously, such statistical and mathematical methods were more closely inte-
grated into and introduced earlier in the curriculum. The reversal of that pattern seen today
in the case of statistics is due to the emergence of data science as a modern catchall for
the learning and practice of such methods. The committee believes that training in math-
ematics and statistics needs to be brought into the core curriculum in a more structured
manner, either complementing or replacing some of the education that currently occurs
outside the core curriculum. This might take the form of a course in mathematical methods
taught within chemical engineering departments focused on illustrating how analytical,
numerical, and statistical methods are used in the context of the equations that emerge
later within specific core courses. The return of this content to the core needs to occur
reasonably early in the course of studies for greatest impact, creating several challenges
given the dense nature of the curriculum and the needs of students entering at different
stages and with different backgrounds and skillsets in mathematics and statistics.
In summary, the undergraduate chemical engineering curriculum has served the
discipline well and will continue to evolve in response to scientific discoveries, techno-
logical advances, and societal needs. In this evolution, it will benefit from rapid changes
in the ways knowledge is disseminated and transferred within and among disciplines. As
part of this evolution, it is also necessary to consider the imperative to attract and retain
practitioners of chemical engineering with increasingly diverse backgrounds, as discussed
in the next section. Later sections of this chapter address the need to enhance and improve
training in business, economics, innovation, and entrepreneurship, as well as lifelong
learning. The committee considers these skills to be essential, well suited to being illus-
trated within the core undergraduate curriculum but entailing foundational knowledge that
lies beyond the core.

BECOMING A CHEMICAL ENGINEER: THE IMPORTANCE OF DIVERSITY

As a discipline, chemical engineering is unique in its pervasive contributions to


society—in areas ranging from energy; to food, water, and air; to health and medicine; to
manufacturing; to materials, as described in earlier chapters of this report. Consequen-
tially, the field is in a strong position to attract a broad range of individuals interested in
the many areas associated with chemical engineering who also are seeking a career with
the potential for societal impact. Research has shown that members of historically ex-
cluded groups are often motivated by the altruistic career goals of making the world better
and giving back to their communities (Thoman et al., 2015). Emphasizing the role of
chemical engineering in addressing societal issues might therefore help attract more high
school students from diverse backgrounds to the undergraduate major.
Women and Black, Indigenous, and People of Color (BIPOC) are underrepre-
sented in chemical engineering relative to the general population, even by comparison

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

236 New Directions for Chemical Engineering

with chemical and biological sciences and related fields. While a career in a STEM (sci-
ence, technology, engineering, and mathematics) field is an attractive path toward altru-
istic work, persisting barriers impede the entry of women and BIPOC into the field. Some
of these barriers affect all historically excluded groups, while others affect students based
on their gender or racial identity, and these barriers can be compounded for students who
are members of more than one historically excluded group. The National Academies and
others have reported extensively on such structural and cultural barriers as unwelcoming
and unsupportive cultures and environments; “gatekeeping”; biases; lack of mentors and
role models; and inequitable policies and practices that impact recruitment, retention, and
career success (see, e.g., McGee, 2020, on barriers for underrepresented and racially
minoritized students; NASEM, 2020b, on barriers for Black students; NASEM, 2016, on
barriers for women and BIPOC broadly; and NASEM, 2018 and 2020a, on barriers for
women).
As a result of such systemic barriers, chemical engineering benefits from the tal-
ents of only a fraction of the population. Educational attainment in chemical engineering
for women (Figure 9-3) and BIPOC (Figure 9-4) has remained essentially unchanged for
more than a decade. The demographics in chemical engineering today reflect the past.
Historically, science and engineering have not been welcoming to BIPOC and women and
have been particularly harsh to Black Americans. In his essay “The Negro Scientist,”
published in The American Scholar in 1939, W. E. B. Du Bois challenged assumptions
held by Whites regarding the propensity of African Americans for science. These types
of biases persist, and after starts and stumbles with interventions designed to counter sys-
temic barriers (NASEM, 2016), work still remains to provide clear and inclusive pathways
in STEM fields, including chemical engineering, for historically excluded groups.
To fully support members of historically excluded groups in chemical engineer-
ing training and education, specialized programs and cultural shifts will be necessary.
Interventions and support mechanisms will vary based on which groups are targeted
(NASEM, 2021c); the focus may be on supporting women,2 or on Black3 or Latinx/His-
panic and Indigenous4 students. The design of specific support mechanisms will vary as
well according to the unique needs and goals of each institution. In this section, and in the
later section on graduate education, the committee presents strategies applicable for most
chemical engineering departments that are likely to improve the recruitment and retention
of and outcomes for multiple underrepresented groups.
Research has illuminated how children’s early pathways can be determined, along
with some of the critical factors that dictate their future educational options and career
trajectories (Akee et al., 2017; Chetty et al., 2016, 2019). These studies have revealed that
children with high scores on third-grade math tests who come from high-income families,
children who grow up in geographic areas with high rates of invention, and girls who are
exposed to women inventors are more likely to become inventors. These findings speak

2
E.g., Society of Women Engineers (https://swe.org/).
3
E.g., National Organization for the Professional Advancement of Black Chemists and Chem-
ical Engineers (https://www.nobcche.org/).
4
E.g, Society for the Advancement of Chicanos/Hispanics and Native Americans in Science
(https://www.sacnas.org/).

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Training and Fostering the Next Generation of Chemical Engineers 237

to the social factors that require attention in any technical field seeking to spark creativity
and build pathways to include the full pool of talent. The importance of role models and
access to opportunities is clear, as is the need for adequate academic preparation for any
STEM field, including chemical engineering. In primary and secondary education, studies
specific to chemical engineering are lacking; studies that include qualitative data and ex-
periences specific to chemical engineering are lacking even more.
Chemical engineering as a field is not immune from systemic and other barriers
to inclusivity, and the field can draw insights and apply the lessons from STEM-wide
studies. In short, if chemical engineering is to reach its full potential as a discipline and a
major enabler of solutions for societal needs, it will need to address opportunity gaps and
ensure that its educational, research, and professional environments support the success
of everyone, regardless of their identity.

FIGURE 9-3 Percentage of bachelor’s, master’s, and PhD degrees awarded to women in chemical
engineering (ChemE), biomedical engineering (BME), and engineering overall, 2008–2018. Data
from the National Center for Science and Engineering Statistics.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

238 New Directions for Chemical Engineering

FIGURE 9-4 Demographic breakdown of degrees awarded to chemical engineers by race: (a)
bachelor’s degrees, (b) master’s degrees, and (c) PhDs. NOTE: In the data for PhDs, the category
of Asian and Pacific Islander is disaggregated, with separate categories for Asian and for Native
Hawaiian and Pacific Islanders; the categories also included an option for “more than one race”
rather than “other race or unknown.” Therefore, these data do not sum to 100 percent because data
were redacted for privacy reasons. Data from the National Center for Science and Engineering
Statistics.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Training and Fostering the Next Generation of Chemical Engineers 239

Cech (2013) speaks to the (mis)framing of engineering as meritocratic and depo-


liticized (asocial). This mischaracterization results from the false assumptions that an in-
herently technical field will inevitably be a meritocracy as though technical fields can
operate outside of social influences (Cech, 2013; Cech and Blair-Loy, 2010). This framing
has been debunked with research revealing the myriad structural and cultural factors that
turn students away from the field regardless of their skills and competencies (Seymour
and Hewitt, 1997), such as structural racism (McGee, 2020), gatekeeping, and weeding
out through historical exclusion in the education system (Malcom, 1996; Malcom and
Malcom-Piquex, 2020), and stereotyping regarding who has innate talent (Leslie et al.,
2015). Science and math courses have long been gatekeepers for entry into engineering.
Retention data disaggregated by discipline are not readily available, but based on the ob-
servations of members of this committee, such courses as the sophomore-level course in
mass and energy balances in chemical engineering serve as additional gatekeepers.
Those chemical engineers who are retained in the field play many roles in practice
and face significant challenges in retaining relevance and excellence as the field evolves;
they do so through diverse educational trajectories and with endpoints in industry, aca-
demia, and elsewhere. When considering what draws people to chemical engineering, it
is important to acknowledge these different educational and career trajectories and the
pressures involved in achieving and maintaining them. Stability, work–life balance, pro-
fessional support structures and mentorship, and opportunities for growth are important
to people in any sector but are not distributed equally across sectors or demographics
within them. In academia, professors act as educators, administrators, mentors, research-
ers, communicators, and fundraisers, but all individuals are not equally suited to all of
these tasks, and some can become overburdened by the need to fulfill them all.
At the same time, members of historically excluded groups bear a disproportion-
ate responsibility for promoting diversity, providing representation on committees, and
supporting the academic and career progression of other members of underrepresented
groups. Institutionalizing the work of diversity requires shared responsibility, not just the
efforts of those with a personal stake in improving access to equitable opportunities. Both
industry and academia still fail to pay and support women and BIPOC and promote them
to executive positions at a proportional rate (Funk and Parker, 2018; Gumpertz et al.,
2017; Renzulli, 2019). Promoting and retaining meaningfully diverse talent will have a
multiplier effect, engendering greater diversity moving forward as more diverse groups
decide who joins the workforce. These issues are pervasive throughout STEM fields and
certainly not unique to chemical engineering. In looking to the future, however, chemical
engineers have an opportunity to be leaders among STEM fields in increasing diversity
and inclusion within their profession.
Greater visibility and connectivity within the broader community can also support
diversity, equity, and recruitment, and chemical engineers can make contributions in other
areas of public interest beyond diversifying the field. Like all scientists and engineers,
they have the opportunity to use effective scientific and popular communication to engage
with the general public, as well as improve resources for K-12 educators. Social media
and science entertainment have been vital for accessibility and visibility, but given the
integral roles of science and engineering in society, scientists and engineers need to be

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

240 New Directions for Chemical Engineering

more integrated into policy making. Chemical engineers have a stake in society equal to
that of people with law or business backgrounds who commonly take on societal concerns,
and are trained in a thought process that would lend itself well to addressing those con-
cerns with respect to both scientific and systemic issues. It is not clear to the average high
school senior that chemical engineering will have a critical role in many of the domains
that will be central to progress in the coming decades. As demonstrated in earlier chapters,
many policy issues in such domains as energy, food security, clean air and water, and
health care, and public health involve chemical engineering. Maintaining dialogue among
chemical engineers in policy-making positions, communities to be served, and those in
academia and industry would also benefit the field, identifying needs and potential prior-
ities.
Science policy is another area in which students are driving growth at universities,
founding student groups and advocating for formalized programs. Being responsive to
those student interests and facilitating this career path can help chemical engineering pro-
grams attract those who want to pursue public-sector work and to develop both the tools
and the influence needed to have direct impact in bettering their community.
This section has addressed recruitment of chemical engineers and barriers to en-
tering the field, in particular for women and BIPOC. Recruiting is critical, but so is reten-
tion. Mentorship has been shown to have a positive effect on underrepresented students,
yet underrepresented individuals enrolled in STEM programs typically receive less men-
torship than their well-represented peers (NASEM, 2019e). Institutionalized developmen-
tal support needs to evolve from “Are you cut out for this?” to “How can we help you
succeed?” Formal support systems for academic success are enhanced by the deliberate
formation of peer and mentoring networks. Beyond mentoring, systemic approaches to
ensuring success for all individuals along the entire career path will ensure that chemical
engineering remains equipped to attract, develop, and retain a diverse cadre of future
chemical engineers.

MAKING CHEMICAL ENGINEERING BROADLY ACCESSIBLE

A long-running national dialogue about college affordability and the impact of


student loan debt on the overall economy has recently become more visible. The relative
affordability of community colleges is a major attraction for a diverse body of students,
ranging from talented budget-minded high school seniors to nontraditional students. In
2021, average annual tuition and fees at 2-year public schools was $3,372, versus $9,580
for in-state students at 4-year public schools (Hanson, 2021). In addition, at least 17 states
have programs that make community college attendance tuition-free for at least a portion
of the student population (Farrington, 2020). Students enrolled at 2-year schools are more
racially diverse than those at 4-year schools (NASEM, 2016), and the majority of tribal
colleges and universities in the United States are 2-year institutions. Further, many states
have existing contractual agreements promising not only admission to public 4-year insti-
tutions for students who demonstrate success at a community college, but also the ability
to graduate within 2 years after transferring. Indeed, it is possible that 2-year community
colleges could become the default choice of the middle class in the relatively near future,

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Training and Fostering the Next Generation of Chemical Engineers 241

requiring a major adaptation of chemical engineering undergraduate programs to become


viable options for those students.
This body of transfer students represents an untapped opportunity for chemical
engineering to broaden participation in and access to the profession. Over the last decades,
science communication and outreach at the K-12 level have resulted in significantly in-
creased interest in STEM fields, but that increase (particularly among diverse groups) has
been focused in areas in which high school courses are available—physics; biology; com-
puter science; and, to a lesser degree, mechanical engineering/robotics. Many high school
graduates have little exposure to the relevance or potential impact of chemical engineering
as a career, and few community colleges offer chemical engineering courses, though many
provide transfer pathways to 4-year schools. Building bridges to actively recruit both full-
time community college students and nontraditional students and identifying and imple-
menting pathways to support them after they transfer could greatly democratize the pro-
fession.
Chemical engineering transfer students face a remarkable challenge beyond the
abrupt change from the community college to the larger university environment (some-
times referred to as “transfer shock”; Flaga, 2006). Chemical engineering curricula gen-
erally have required courses beginning early in the sophomore year, with many programs
offering an introductory course in the first year. Further, education research has under-
scored the importance of providing early hands-on experiences in engineering to improve
students’ motivation to complete their degrees (Cui et al., 2011). In fact, such experiences
have been shown to be particularly successful in the retention of women and BIPOC (re-
spectively, a 27 percent, 54 percent, and 36 percent retention gain for women, Latinx, and
African American students; Hoit and Ohland, 1998; Knight et al., 2003; Napoli et al.,
2017; Willson et al., 1995). These gains are attributed not only to increased design, team-
work, and communication skills, but also to the development of a peer support network
(Richardson and Dantzler, 2002). Challenges for transfer students are compounded be-
cause, in contrast with prerequisites in chemistry, physics, and mathematics, most 2-year
community colleges lack chemical engineering departments to offer these courses, much
less hands-on experiences. Further, because students at community colleges do not fulfill
any major requirements, they do not form a peer support network with other chemical
engineering majors. As a result, transfer students are asked to compress most of 3–4 years
of chemical engineering curriculum into a 2-year period, and to do so without the same
peer support or foundational experience in engineering enjoyed by their nontransfer peers.
This is not a recipe for success.
The challenge of accommodating a 2-year path to graduation for community col-
lege transfer students is already facing many undergraduate programs at public universi-
ties. This is an ideal opportunity for the widespread deployment of online course offerings
within the sophomore chemical engineering curriculum (mass and energy balances, a first
course in transport phenomena and/or thermodynamics, and likely a course in mathemat-
ics for chemical engineering applications). Further, from a student perspective, these
courses need to be widely accepted so as to open up options for transfer to a variety of 4-
year programs. Despite the obvious administrative hurdles, these courses would be most
beneficial if crafted by and offered as a collaboration among leading large universities

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

242 New Directions for Chemical Engineering

(public and private), along with the accreditation agencies and the American Institute of
Chemical Engineers (AIChE), thereby promoting universal acceptance of these courses
as satisfying prerequisites for the junior-level curricula at individual 4-year institutions.
In addition to academic hurdles, community college transfer students may face
financial or other challenges that, while unrelated to their academic abilities, affect their
performance. For example, the lack of study groups or support systems noted above
(Lenaburg et al., 2012) can translate into a sizable drop in their grade point average during
the first term, which can have long-term impacts on the future potential for graduate study
and career options. In addition, given the relatively short time spent at a 4-year university,
such transfer students typically do not participate in campus undergraduate research ex-
periences, thus missing out on important opportunities for professional skills development
and resumé building. Universities need to develop systems to support and engage these
students early on and establish peer networks that will support their acclimation and aca-
demic success (Eris et al., 2010; Litzler and Young, 2012). Importantly, doing so will
build students’ confidence and teach important workplace skills.
The committee recognizes that adding more experiential learning earlier in a tra-
ditional 4-year curriculum (as discussed previously in this chapter) and making that same
4-year curriculum more welcoming to transfer students from 2-year institutions are seem-
ingly at odds with one another. The experience of a transfer student in a 4-year program
will never be the same as that of a student who entered as a freshman, but it is the com-
mittee’s hope that more practicum-like experiences will become standard across all intro-
ductory STEM courses, whether offered at a 2-year or 4-year institution. For this reason,
the committee also chose to highlight an example of experiential learning (Box 9-2) that
does not require expensive or specialized equipment, as well as the development of virtual
experiential learning experiences and courses that satisfy the chemical engineering degree
requirements typically offered during the sophomore year.

TEACHING UNDERGRADUATE STUDENTS TODAY AND TOMORROW

Delivery Methods

In spring 2020, the COVID-19 pandemic caused almost all U.S. higher education
institutions to move abruptly to an online format, greatly accelerating ongoing trends to-
ward efficiency and scale within higher education. The result was the creation of signifi-
cant online content of widely varying quality and a deeper understanding of what does
and does not work in synchronous and asynchronous online modes both for chemical en-
gineering and more generally.
To some degree, chemical engineering courses, regardless of the subject, tradi-
tionally start with fundamentals and end with practice (if time permits). This pattern re-
flects the desire of educators to teach tools that can be adapted throughout a student’s
career rather than a vocational skill. In practice, however, this approach has resulted in
the derivation of fundamental equations in lecture, with application and problem solving
occurring in discussion sections and problem sets. The online experience of 2020–2021
amplified existing trends toward classroom delivery that encourages more problem- and

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Training and Fostering the Next Generation of Chemical Engineers 243

project-based and group learning, which appears to be welcomed by students. One mode
of implementing this “flipped” approach that became prominent in the online format was
including the derivation in an asynchronous recording and then solving the problems live.
The obvious weaknesses of this approach are a potential lack of understanding of asyn-
chronous content or noncompliance with requirements to watch it. There are also some
indications that such flipped approaches or greater use of team-based learning may dis-
proportionately affect women, BIPOC, and low-income students; however, the results of
early research are conflicting in this regard (e.g., Cruz et al., 2021; Deri et al., 2018; Dixon
and Wendt, 2021; Raišienė et al., 2021; Sarsons, 2017; Winter et al., 2021). More research
is needed in this area, and any move toward more asynchronous learning and/or team- and
project-based learning will need to ensure equitable outcomes for all students.
In this regard, it may be more realistic to consider online delivery as the new form
of “self-teaching.” Historically, the U.S. higher education system (regardless of disci-
pline) has strived to impart to students both concrete knowledge and skills and the ability
to learn new skills throughout their career. In this sense, the purpose of courses is to teach
content that is critical, but that is either too difficult or not obviously motivating to learn
on its own. If viewed through this lens, the curation of online content that is either of a
complexity level appropriate for self-teaching or related to subjects for which the student
is motivated offers an exciting prospect for lifelong learning, allowing for the uniform
distribution of better content at lower cost and democratizing the offering of specialized
courses around the world. Augmentation of existing courses (with modules, examples, or
alternative explanations), communication and other soft skills, business/entrepreneur-
ship/management, policy and regulatory issues, and extensions of course content are areas
in which online and classroom learning may dovetail within a single course or curriculum.
All online content carries the curse of the internet—namely, the varying quality
and accuracy of information and content. Within a single course or curriculum, curation
of content will become a major faculty responsibility. With respect to extension learning,
curation of content may become a major endeavor of professional societies such as
AIChE. In the past, these societies have served their membership in terms of information
dissemination and continuing professional education via conferences, workshops and
seminars, and the publication of journals. Going forward, this role will likely shift not just
to one that is online in nature but to one that is more geared toward serving as a trusted
curator of outside resources, perhaps via a subscription model rather than content gener-
ation and ownership.

Curricular Content Evolution

As discussed earlier in this chapter, the chemical engineering curriculum has in


recent years sought to balance the goals of retaining core rigor (mathematical modeling,
thermodynamics, kinetics, and design) and incorporating new important topics (most re-
cently, biochemical engineering and data science). With a massive expansion of the num-
ber of STEM majors in many institutions (Figure 9-5), enrollment in the introductory se-
quences of chemistry, physics, mathematics, and biology has grown to the point that

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

244 New Directions for Chemical Engineering

chemical engineering students make up a small portion of STEM students. This shift has
led to the question, posed differently at every university based on its politics and financial
construct, of the degree to which these other departments should teach introductory ma-
terial relevant to chemical engineering and to what degree a chemical engineering pro-
gram is responsible for its own introductory material. For example, each chemical engi-
neering major in the United States is required to take a calculus sequence from a
mathematics department. Some chemical engineering departments have developed sup-
plemental undergraduate mathematics content to incorporate coverage of relevant partial
differential equations, linear algebra, numerical methods, and the data science and statis-
tics tools needed by chemical engineering students. This trend is likely to be limited, how-
ever, by the relatively small size of chemical engineering departments and an inability to
teach an entire undergraduate curriculum without the aid of sister departments on campus.
Further, as discussed above, one major drive toward college affordability and diversity
and inclusion is the broadening of a path for transfer students from lower-cost community
colleges and students who change their majors. Neither is well served by a curriculum
that is monolithically specialized starting in the freshman year.

FIGURE 9-5 Total STEM degrees (bachelor’s, master’s, PhD) awarded between 2008 and 2018.
Data from National Center for Science and Engineering Statistics.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Training and Fostering the Next Generation of Chemical Engineers 245

TEACHING GRADUATE STUDENTS TODAY AND TOMORROW

Chemical engineering research has expanded considerably in breadth and scope


over the past decades, and it is likely to continue doing so. This expansion requires that
graduate students acquire deep knowledge in adjacent fields and subfields, including, for
example, biology, materials science, and applied physics. At the same time, increasing
demands on undergraduate education programs have necessitated reduced coverage of
core chemical engineering topics, such as thermodynamics and transport phenomena. The
question facing graduate education in the field is whether to compensate for the depth and
content that are no longer provided by a chemical engineering undergraduate degree or to
focus on giving students the flexibility and opportunities to largely tailor their own grad-
uate program.
A core graduate program consisting of thermodynamics, statistical and quantum
mechanics, transport phenomena, chemical kinetics, and applied mathematics will con-
tinue to be necessary for chemical engineering research, but that content will have to be
delivered in a manner that allows students to apply it in a wider range of contexts. Courses
in thermodynamics, for example, will have to rely on approaches and examples that illus-
trate the general applicability of the underlying concepts, from problems related to issues
of protein stability; to general free-energy minimization techniques; to phase transitions
in mixtures of solids, liquids, or gases. The core curriculum will need to be limited to
foundational concepts, thereby giving students the flexibility to pursue coursework in ar-
eas of direct relevance to their research.
While graduate preparation in chemical engineering builds on undergraduate ma-
terial, this exclusivity comes at the cost of diversity in terms of both the number of women
and BIPOC and the breadth of scientific backgrounds in the chemical engineering gradu-
ate population. The imperative to recruit talent from a more diverse range of backgrounds
will require, in addition to the measures discussed above, the opening up of chemical
engineering by providing background content in a manner that creates opportunities for
students from other disciplines (e.g., chemists, physicists, biologists) to join a graduate
chemical engineering program. That material would include elements from the core sub-
jects covered in the undergraduate curriculum but organized and delivered in a way that
is easily accessible to postgraduate scientists or engineers. Interestingly, anecdotal evi-
dence from the members of this committee indicates that while many chemical engineer-
ing graduate programs use undergraduate preparation in the field as a major gateway to
admission, faculty of their own departments include many members whose training was
in related disciplines.
Relative to undergraduate programs in chemical engineering, those in chemistry
and biology graduate significantly more women (50 percent and 63 percent, respectively,
compared to 35 percent in chemical engineering; NCSES, 2018). Similarly, undergradu-
ate programs in chemistry and biology are more racially diverse (58 percent and 55 per-
cent White, respectively, compared with 64 percent white in chemical engineering;
NCSES, 2018). By considering admitting more graduate students with undergraduate de-
grees in these related disciplines, chemical engineering departments could provide oppor-
tunities for more diverse applicant pools. Significantly, a decision not to accept graduate

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

246 New Directions for Chemical Engineering

students from other, related fields would generally rule out the enrollment of graduates of
many historically Black colleges and universities and other minority-serving institutions
that lack undergraduate chemical engineering programs. At the same time, a search for
diverse graduate students cannot be limited solely to those institutions. Even when re-
cruiting within undergraduate chemical engineering programs, there is an elitism in some
graduate programs that excludes those who may have chosen a bachelor’s institution be-
cause of its affordability or location. For a graduate program, it is justifiable to want stu-
dents to understand what it means to do research before committing them to, and support-
ing them for, a program lasting many years. But how is genuine interest cultivated for
those who did not know earlier about or did not have access to undergraduate research
opportunities, or for those members of historically excluded groups who did have the
chance to participate in such programs as the Research Experience for Undergraduates
but are then not actively recruited to the host institutions?
Another vehicle for graduate education that has until recently been largely miss-
ing from the graduate chemical engineering curriculum is internships in industry, govern-
ment, or the nonprofit sector. Experiential learning in the form of graduate internships is
currently rare, and providing sufficient opportunities for systematic placement of graduate
students will require a conversation among industry, federal funding agencies, universi-
ties, and professional organizations such as AIChE to enable the development of suitable
frameworks capable of administering effective training programs, perhaps even on a na-
tional scale. As with coursework, new opportunities are emerging through remote and
virtual access. New models will likely be needed that address issues of equity and inclu-
sion, suitable compensation, intellectual property considerations, and a commitment to
the mentoring of graduate interns. Encouraging companies to create educational/intern-
ship opportunities by creating model programs would be beneficial.
Master’s degrees are likely to play an increasingly important role in graduate ed-
ucation. For the reasons outlined above—whether a need to acquire additional depth in
core chemical engineering concepts or to gain breadth in ancillary disciplines such as
bioengineering or computing or data science, among many others—master’s degrees
could offer an attractive solution for chemical engineers needing to adapt and respond to
a rapidly changing marketplace. One obstacle that remains to be addressed is the cost of
such degrees. However, with the emergence of improved options for remote learning, and
with compelling examples of high-quality degrees (e.g., the Online Master of Science in
Computer Science from the Georgia Institute of Technology, with more than 5,000 grad-
uates in its 8 years of existence; McMurtrie, 2018; Nietzel, 2021) being offered for less
than the cost of the typical undergraduate tuition at a state institution, students and em-
ployers alike will benefit from the flexibility offered by master’s and graduate certificate
programs. The chemical engineering community will have to develop carefully conceived
degrees that can not only provide in-depth, topical chemical engineering content for stu-
dents and practitioners but also attract students from other disciplines. Such programs will
also need to provide the flexibility required by working professionals who wish to con-
tinue their education and earn an advanced degree through evening and/or weekend pro-
grams.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Training and Fostering the Next Generation of Chemical Engineers 247

NEW LEARNING AND INNOVATION PRACTICES


TO ADDRESS CURRENT CHALLENGES

The preceding sections outline significant steps that could be taken to grow and
diversify the field of chemical engineering and deepen its impact on society. These ob-
servations about curriculum, experiential learning, approaches to teaching, and ways of
building a more diverse and inclusive profession, as well as the broader issue of control-
ling the costs of higher education, raise the question of whether education in the field can
be better designed to deliver on future opportunities. Is it possible to double the quality of
an education (including an outcome measure of student success) delivered in half the time
and at half the cost? How can education be made globally accessible in real time? What
are possible new options for addressing the identified challenges facing chemical engi-
neering education? Answers to these questions could reflect and incorporate the ways in
which technology is transforming how work is performed in many professions and how
global networks have transformed knowledge management.
In the past, problems were solved based on what an individual or group of indi-
viduals knew, acquired from discrete sources (e.g., books, articles, other publications, and
their own lived experiences). Today, in contrast, essentially all of the world’s public in-
formation has been indexed and is quickly available, often at no cost, via internet search
engines. When confronted with a problem in almost any setting, those with internet access
first search for possible solutions or known information. Their initial findings connect to
nearly limitless information about related topics and solutions and opportunities for an
individual to find and build upon what is known. An education is necessary to provide
sufficient background in the subject matter so the problem can be formulated, to curate
and validate information, and then to know how to apply the information to create the
needed solutions.
The nature of the education needed to solve problems in this way is evolving.
Several companies in the technology arena have eliminated previously held requirements
for a 4-year degree for many jobs and are leading the development of more targeted cer-
tificate programs to create a pool of talent for the range of jobs that are open. For example,
Google has launched “Grow with Google,” a program wherein completion of an online
certificate program available on Coursera can lead to entry-level jobs with competitive
starting salaries.5 There is no requirement for a 4-year degree or equivalent experience.
The Google program engages more than 100 partner companies that also have positions
available upon completion of the common certificate programs. The traditional higher
education model is linear, with a student moving from K-12 to some amount of university
education and then to work, and there are usually limited feedback loops and a lack of
integration across the steps. That linear model could be transformed into a general learn-
ing model based on a shared platform integrating the educational silos found today.

5
See http://grow.google.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

248 New Directions for Chemical Engineering

The degree to which truly new models of skills-based education will penetrate
chemical engineering is unclear. The advantages of the integrated approach discussed
above are numerous and are likely to remain attractive for many students. On the other
hand, there is ample evidence that U.S. research universities are not designed to deliver a
low-cost undergraduate education. Their many missions result in high costs (to support
faculty and research) and high overhead rates (to pay for people and facilities), and they
are not responsive over the time scales of the connected world because of the nature of
their research and scholarship. New connected models of learning and innovation that
span traditional boundaries could provide solutions more responsive to some of today’s
needs, although major barriers, including the incorporation of laboratory classes, would
have to be overcome.
Learning and innovation could be designed to span the boundary between univer-
sities and workplaces. Instead of universities taking on more responsibilities, contributors
could build content that would be shared and could be used at all stages of education. That
content could be both scalable and used locally in the classroom or as a supplement to
classroom learning. Given the accelerating pace of change, such changes would assist
learners of all ages in thinking about future career options that can be aligned with their
learning and skills development (see Box 9-3).
The move to skills-based hiring opens up other new possibilities. An existing de-
gree is at some basic level a collection of skills. Different disciplines have both unique
and common skills, the latter of which enable new mapping of those skills to other disci-
plines, as well as to different jobs. Skills-based modules offered as complements to exist-
ing degree programs could provide a lower-cost path to a first job, along with continued
support for additional lifelong learning in response to the evolution of the job market.
New learning networks could also help build a more diverse STEM workforce by enhanc-
ing access to much-needed career opportunities, background preparation, and support.
As discussed earlier in this chapter, much remains to be learned about the relative
advantages of online and classroom teaching and learning. While online programs have
gained popularity and were widely used during the COVID-19 pandemic, their advantages
clearly come at the cost of the interpersonal interactions and discussions that occur in a
traditional classroom. Chemical engineers have an opportunity to lead innovation in
STEM fields by building a model that emphasizes scalability as well as human connec-
tion. Scalability can come from online content that is curated jointly by companies and
universities and made available for local use in the classroom, or from the use of stand-
alone modules to address particular topics that are accessed entirely online. Human inter-
action can then come from internships or work on extended projects at a company. The
possibilities are numerous, even as the existing business model and the set of priorities
now in place in universities, companies, and government create barriers to change (e.g.,
Conn et al., 2021). As with all major disruptions, change will likely be generated first by
companies as they struggle to find, hire, and develop future talent.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Training and Fostering the Next Generation of Chemical Engineers 249

BOX 9-3
The American Institute of Chemical Engineers (AIChE)
Institute for Learning and Innovation

The AIChE Institute for Learning and Innovation (ILI) provides a “horizontal” connection
from university to workplace. In this university/local business model, shared content can be
accessed by all stakeholders for use in the classroom. The model also enables sharing across
universities and building new collaborations among multiple stakeholders. The ILI recognizes
that technology is rapidly transforming most market sectors, that learners of all ages need to
keep pace with evolving skill requirements, and that companies will seek contemporary skills
and capabilities as their business continues to change. A particular focus of the effort is on
providing high-quality and low-cost content for institutions of all sizes.
The initial ILI has four basic modules: Career Discovery, Academy, Practice, and Creden-
tial, illustrated below.

The Career Discovery module, which has been piloted and launched through the ILI, begins
with asking students “What might you like to do?” rather than “What do you want to do?”
Through a series of exercises, lectures, and group discussions, students develop a personal career
plan for working through the university curriculum. Additional skills and experiences are high-
lighted, and through the broad network of the ILI students have options for engaging in addi-
tional skills training, internships, and other outreach activities that will help them get their first
job after graduation. The same approach is being used within companies by midcareer profes-
sionals seeking new skills.
The Academy, Practice, and Credential modules are all designed cooperatively by industry
and universities. The modules provide training in high-priority skills in the marketplace, as well
as in specific skills and applications prioritized by companies. AIChE will make shared content
available for use in universities, allowing university education to shift in emphasis toward ex-
ploring new engineering or business problems. New internship models to engage students with
market problems are also under development, and companies can engage with classroom exer-
cises in order to share case studies. Finally, “Lessons from Leaders” is an emerging module in
which senior leaders share their career experiences and learning with students considering var-
ious career options.
A related scholarship program for Black, Indigenous, and People of Color (BIPOC), the
Future of STEM Scholars Initiative (FOSSI), is providing expanded opportunities for students
with limited exposure to or understanding of possible market options. Building on studies of
disruptive innovation provides a major opportunity to create new education models for histori-
cally excluded groups through partnerships among companies, community colleges, and histor-
ically Black colleges and universities. Such models would provide a low-cost pathway to high-
quality jobs with a direct connection to market need while at the same time broadening outreach
to a more diverse talent pool.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

250 New Directions for Chemical Engineering

CONCLUSION

The core chemical engineering curriculum has contributed to the long-term suc-
cess and impact of the discipline. The undergraduate curriculum provides a mathematical
framework for designing and describing (electro-/photo-/bio-) chemical and physical pro-
cesses across spatial and temporal scales of many orders of magnitude. Data science and
statistics may be delivered most effectively within a separate course embedded within the
core curriculum and taught with specific emphasis on matters of chemistry and engineer-
ing. At the same time, experiential learning is important, and the majority of industrial
and academic chemical engineers interviewed by the committee stressed the importance
of internships and other practical experiences. However, far fewer internships are availa-
ble than the number of students who would benefit from them, and the density of the core
undergraduate curriculum leaves few openings for incorporating an additional hands-on
laboratory course earlier in the curriculum.
The current chemical engineering curriculum is well suited to preparing students
for a wide variety of industrial roles. Graduate research increasingly encompasses a di-
verse range of topics that do not all require the level of traditionally curated knowledge
currently delivered in graduate chemical engineering curricula, and so graduate curricula
may need to be adjusted. Internships for graduate students are currently rare, and new
models will need to address issues of equity and inclusion, suitable compensation, intel-
lectual property considerations, and adequate mentoring of interns.
Women and members of historically excluded groups are underrepresented in
chemical engineering relative to the general population, even by comparison with the
chemical and biological sciences and related fields. Diversifying the profession is essen-
tial to the field’s survival and potential for impact. At all points along their academic path,
chemical engineering students need role models and effective, inclusive mentors, includ-
ing mentors that reflect the diversity of backgrounds needed by the field. Leveraging of
professional societies and associated affinity groups could provide valuable support for
people of diverse backgrounds entering the field. Strong university support for student
chapters of professional organizations would improve access and success.
The general affordability of community colleges is a major attraction for a diverse
body of students, ranging from budget-minded high school seniors to nontraditional stu-
dents. Increased engagement of transfer students is an untapped opportunity for chemical
engineering to broaden participation in and access to the profession. Students from 2-year
colleges and those who change their major to chemical engineering would benefit from a
redesign of the curriculum allowing them to complete the degree in less time. Better aca-
demic and social support structures are needed to enable successful pathways for these
students. New methods for offering portions of the curriculum in a distributed manner and
more general restructuring may require flexibility in curriculum design and changes in
university policies, graduation, and accreditation requirements.

Recommendation 9-1: Chemical engineering departments should consider revisions


to their undergraduate curricula that would

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Training and Fostering the Next Generation of Chemical Engineers 251

 help students understand how individual core concepts merge into the
practice of chemical engineering,
 include earlier and more frequent experiential learning through physical
laboratories and virtual simulations, and
 bring mathematics and statistics into the core curriculum in a more
structured manner by either complementing or replacing some of the ed-
ucation that currently occurs outside the core curriculum.

Recommendation 9-2: To provide graduate students with experiential learning op-


portunities, universities, industry, funding agencies, and the American Institute of
Chemical Engineers should coordinate to revise graduate training programs and
funding structures to provide opportunities for and remove barriers to systematic
placement of graduate students in internships.

Recommendation 9-3: To increase recruitment and retention of women and Black,


Indigenous, and People of Color (BIPOC) individuals in undergraduate programs,
chemical engineering departments should emphasize opportunities for chemical en-
gineers to make positive societal impacts, and should build effective mentoring and
support structures for students who are members of such historically excluded
groups. To provide more opportunities for BIPOC students, departments should
consider redesigning their undergraduate curricula to allow students from 2-year
colleges and those who change their major to chemical engineering to complete their
degree without extending their time to degree, and provide the support structures
necessary to ensure the retention and success of transfer students.

Recommendation 9-4: To increase the recruitment of students from historically ex-


cluded communities into graduate programs, chemical engineering departments
should consider revising their admissions criteria to remove barriers faced by, for
example, students who attended less prestigious universities or did not participate in
undergraduate research. To provide more opportunities for women and Black, In-
digenous, and People of Color (BIPOC) individuals, departments should welcome
students with degrees in related disciplines and consider additions to their graduate
curricula that present the core components of the undergraduate curriculum tai-
lored for postgraduate scientists and engineers.

Recommendation 9-5: A consortium of universities, together with the American In-


stitute of Chemical Engineers, should create incentives and practices for building
and sharing curated chemical engineering content for use across universities and
industry. Such sharing could reduce costs and advance broad access to high-quality
content intended both for students and for professional engineers intending to fur-
ther their education or change industries later in their careers.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

252 New Directions for Chemical Engineering

Recommendation 9-6: Universities, industry, federal funding agencies, and profes-


sional societies should jointly develop and convene a summit to bring together per-
spectives represented by existing practices across the ecosystem of stakeholders in
chemical engineering professional development. Such a summit would explore the
needs, barriers, and opportunities around creating a technology-enabled learning
and innovation infrastructure for chemical engineering, extending from university
education through to the workplace.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

10
International Leadership

 U.S. leadership in chemical engineering has decreased significantly in the past 15


years as measured by number of publications, now about half the number pro-
duced by China.
 U.S. leadership in the impact of scholarly research in chemical engineering as
measured by citation impact has likewise lost ground to international competitors.
 The United States is in a leadership position in some areas of chemical engineer-
ing technology, but lags in many niches compared with various other countries.

This report presents a U.S.-based perspective on the field of chemical engineer-


ing, but like most areas of science and engineering, chemical engineering is a globally
connected enterprise both commercially and academically. Assigning advances within the
field to an individual country is difficult, particularly for commercial innovations devel-
oped by multinational companies, but some measures of national scientific (mainly aca-
demic) output can be measured quantitatively. To that end, the committee used a method-
ology developed during a previous National Academies study for benchmarking the
competitiveness of U.S. chemical engineering (NRC, 2007). Applying this methodology,
the committee examined scientific outputs, defined as papers published in peer-reviewed
journals, over 5-year intervals from 1981 to 2020—a period encompassing that covered
by the previous study and extending it through the most recent complete calendar year.
The committee also estimated the quality and impact of those papers across geographic
regions by comparing the field-weighted citation impact of the top 100 papers and the
number of highly cited papers.
This chapter provides a perspective on U.S. leadership in chemical engineering
that is focused on publications and research output. It should be noted that that the em-
phases and directions of chemical engineering as a discipline and a profession vary across
countries and regions. These variations can result from differences in resource availabil-
ity, political and economic systems, and chemical engineering education. A more com-
prehensive view considering national investments in chemical engineering research and
development (R&D), training curricula, and impact on gross domestic product, for exam-
ple, could provide a more nuanced and valuable perspective. Such data, however, are dif-
ficult to obtain and are often not disaggregated by field or discipline. Furthermore, such
an extensive data analysis was beyond the committee’s charge. In many areas of chemical
engineering, moreover, there are clear examples in which national or regional challenges,
such as water scarcity, have driven research and technology development at a high level,
as well as cases in which existing infrastructure, such as refineries, is well matched to
such challenges and therefore enables regional advances.

253

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

254 New Directions for Chemical Engineering

PUBLICATION RATES AND CITATION ANALYSIS

The methodology of the 2007 study was based on analysis of a large base of sci-
entific journals (listed in Appendix 3B of NRC, 2007). For the present analysis, journals
on the original list that are no longer being published were removed, and those whose
publication began in 2007 or thereafter were incorporated (Appendix B). The total number
of papers published was found by searching Scopus for all publications for which any
author or coauthor had a chemical engineering affiliation in the address field in each of
these journals over each time interval. As indicated in the 2007 report, a chemical engi-
neering affiliation is a reliable measure of a researcher’s being involved in chemical en-
gineering in the United States, but likely undercounts authors in other regions where de-
partment names do not necessarily include the term “chemical engineering.” These data
include all countries affiliated with each paper. Over the entire 40-year period, 1,250 pa-
pers were published with coauthors from the United States, Asia, and the European Union
(EU); 6,619 with coauthors from the EU and Asia; 9,722 with coauthors from the United
States and the EU; and 16,153 papers were published with coauthors from the United
States and Asia. Note that Asia comprises China, Korea, Taiwan, India, and Japan, and
the EU includes the current membership and England.
Publishing by U.S. chemical engineers has continued to grow since 2007, but as
was already seen in 2007, the rate of growth of the other regions dramatically exceeds
that of the United States. Some of this growth was previously attributed to the increased
number of researchers in those other regions rather than increased publication rates by
individual chemical engineers; indeed, while China currently has a low number of re-
searchers per capita (Heney, 2020), a larger share of their undergraduates is awarded
STEM (science, technology, engineering, and mathematics) degrees (49 percent) relative
to the share in the United States (33 percent; NSF, 2016). Publications from Asia have
grown nearly 100-fold since 1981, eclipsing other regions and making up 57 percent of
the total papers published between 2016 and 2020, compared with 23 percent for the
United States (Figure 10-1).
The same trend is visible in the regional shifts in publication in those journals
identified as core to chemical engineering in the 2007 study—the AIChE Journal, Indus-
trial & Engineering Chemistry Research, and Chemical Engineering Science (Table 10-
1). For these journals, a similar search was performed without specifying chemical engi-
neering affiliation. In the last 5 years for all three journals, Asia has become the dominant
source of publications, with 50 percent of the total, while the U.S. share has declined from
42 percent in 1991–1995 to 20 percent in the last 5 years.
While the quantity of articles is easy to determine, their quality and impact are
more difficult to ascertain. In the short term, a measure of the impact of a publication is
the number of times it is cited (Table 10-2). Asia has recently accounted for the largest
share of the 100 most cited articles in the three core chemical engineering journals. For
the list of top 100 cited articles, countries within a geographic region that were not in-
cluded in the original 2007 region search list were binned with that region; for example,
Singapore was included in the count for Asia.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

International Leadership 255

FIGURE 10-1 Number of publications by region from selected journals with at least one author
with a chemical engineering department affiliation over 5-year intervals from 1981 to 2020. Asia
comprises China, Korea, Taiwan, India, and Japan, and the European Union (EU) includes current
membership and England. “Original journals” includes all those used in the 2007 study (NRC,
2007) and “new journals” includes those added for the current analysis whose publication and/or
indexing by Scopus began in 2007 or thereafter.

Arguably, the work published by chemical engineers that has the most impact is
published in journals whose scope is much broader than chemical engineering. Such jour-
nals include Science, Proceedings of the National Academy of Sciences, and Nature. In-
deed, the fraction of publications in these three journals with a coauthor having a chemical
engineering affiliation has grown over time (Table 10-3), which speaks to a general broad-
ening of the scope and impact of chemical engineering research (or at least research done
by chemical engineers). As is the case for the three narrower chemical engineering jour-
nals discussed above, contributions from Asia have increased substantially, but in the
broader three journals, papers authored by chemical engineers are still primarily of U.S.
origin. Interestingly, the number of papers resulting from collaborations across regions
has increased substantially in recent decades.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering
TABLE 10-1 Allocation of All Articles in AIChE Journal, Industrial & Engineering Chemistry (I&EC) Research, and Chemical

256
Engineering Science Published within Previously Defined Regions and Percentages over 5-Year Periods from 1991 to 2020
Total Articles 1991–1995 1996–2000 2001–2005 2006–2010 2011–2015 2016–2020
United States 716 (66%) 749 (52%) 666 (47%) 504 (31%) 637 (35%) 672 (35%)
Europe 180 (16%) 308 (21%) 409 (29%) 448 (28%) 426 (23%) 382 (20%)
AIChE Journal
Copyright National Academy of Sciences. All rights reserved.

Asia 117 (11%) 190 (13%) 224 (16%) 429 (26%) 587 (32%) 741 (39%)
Canada 75 (7%) 80 (6%) 83 (6%) 108 (7%) 146 (8%) 138 (7%)
S. Am 14 (1%) 23 (2%) 28 (2%) 45 (3%) 24 (1%) 49 (3%)

Total 1,091 1,440 1,431 1,619 1,825 1,922


United States 1,049 (45%) 1,183 (38%) 1,463 (33%) 1,489 (24%) 1,604 (18%) 1,615 (19%)
Europe 495 (21%) 782 (25%) 1,245 (28%) 1,526 (25%) 1,756 (20%) 1,515 (17%)
I&EC Research

Asia 486 (21%) 658 (21%) 1,010 (23%) 2,083 (34%) 4,304 (49%) 4,832 (56%)
Canada 125 (5%) 206 (7%) 266 (6%) 372 (6%) 434 (5%) 447 (5%)
S. Am 49 (2%) 86 (3%) 174 (4%) 230 (4%) 316 (4%) 330 (4%)

Total 2,317 3,094 4,410 6,114 8,765 8,698


United States 587 (27%) 553 (23%) 512 (19%) 599 (19%) 541 (17%) 441 (14%)
Chemical Engineering

Europe 757 (35%) 954 (40%) 990 (37%) 1092 (35%) 1164 (36%) 969 (30%)
Asia 262 (12%) 362 (15%) 536 (20%) 724 (23%) 938 (29%) 1403 (43%)
Science

Canada 140 (6%) 153 (6%) 168 (6%) 256 (8%) 208 (6%) 201 (6%)
S. Am 23 (1%) 65 (3%) 78 (3%) 78 (2%) 60 (2%) 112 (3%)

Total 2,159 2,377 2,686 3,135 3,232 3,265


New Directions for Chemical Engineering
TABLE 10-2 Region of Authors of the Top 100 Most Cited Articles Published in a Given 5-Year Period between 1991 and 2020 in Selected
Core Chemical Engineering Journals: AIChE Journal, Industrial & Engineering Chemistry Research, and Chemical Engineering Science
100 most cited 1991–1995 1996–2000 2001–2005 2006–2010 2011–2015 2016–2020
United States 69 59 52 35 54 32
Europe 14 20 28 24 12 15
Copyright National Academy of Sciences. All rights reserved.

AIChE Journal

Asia 5 9 15 30 27 45
Canada 10 6 1 8 3 2
S. Am 1 1 0 0 0 2
Australia 1 4 1 2 4 2
Other 0 1 3 1 0 2
United States 51 41 43 32 14 16
Europe 22 27 30 25 12 13
I&EC Research

Asia 14 19 17 20 56 57
Canada 3 5 4 10 7 2
S. Am 1 0 1 3 0 2
Australia 4 6 3 9 2 2
Other 5 2 2 1 9 8
United States 30 26 19 20 24 13
Chemical Engineering

Europe 53 51 48 37 34 19
Asia 4 14 24 22 24 49
Science

Canada 10 7 4 11 6 7
S. Am 0 0 0 2 1 1
Australia 1 2 3 7 10 5
Other 2 0 2 1 1 6
NOTE: Individual articles may be attributed to more than one country.

257
New Directions for Chemical Engineering
TABLE 10-3 Number and Percentage by Region of Articles in Science, Proceedings of the National Academy of Sciences (PNAS), and

258
Nature with At Least One Author with a Chemical Engineering Department Affiliation
1991–1995 1996–2000 2001–2005 2006–2010 2011–2015 2016–2020
Total Papers 11,834 12,816 11,922 11,005 10,347 12,488
Total ChemE 76 (1%) 56 (0%) 84 (1%) 143 (1%) 174 (2%) 277 (2%)
Copyright National Academy of Sciences. All rights reserved.

U.S. ChemE 60 (79%) 53 (95%) 81 (96%) 134 (94%) 145 (83%) 225 (81%)
EU ChemE 12 (16%) 5 (9%) 15 (18%) 26 (18%) 41 (24%) 64 (23%)
Science

Asia ChemE 4 (5%) 3 (5%) 6 (7%) 23 (16%) 37 (21%) 101 (36%)


Canada ChemE 0 (0%) 0 (0%) 1 (1%) 3 (2%) 5 (3%) 20 (7%)
S. Am ChemE 0 (0%) 0 (0%) 0 (0%) 2 (1%) 1 (1%) 9 (3%)

Internationalization 0% 9% 23% 31% 32% 51%

Total Papers 12,889 13,994 15,642 19,697 21,065 19,683

Total ChemE 40 (0%) 66 (0%) 138 (1%) 377 (2%) 528 (3%) 719 (4%)

U.S. ChemE 38 (95%) 63 (95%) 131 (95%) 336 (89%) 486 (92%) 604 (84%)
EU ChemE 3 (8%) 2 (3%) 19 (14%) 57 (15%) 100 (19%) 142 (20%)
PNAS

Asia ChemE 1 (3%) 3 (5%) 13 (9%) 40 (11%) 109 (21%) 219 (30%)
Canada ChemE 0 (0%) 2 (3%) 4 (3%) 19 (5%) 22 (4%) 43 (6%)
S. Am ChemE 0 (0%) 0 (0%) 1 (1%) 2 (1%) 5 (1%) 9 (1%)

Internationalization 5% 6% 22% 20% 37% 41%


New Directions for Chemical Engineering
Total Papers 15,048 14,067 12,871 12,156 12,294 11,938

Total ChemE 48 (0%) 61 (0%) 51 (0%) 73 (1%) 97 (1%) 186 (2%)

U.S. ChemE 38 (79%) 52 (85%) 45 (88%) 58 (79%) 76 (78%) 153 (82%)


Copyright National Academy of Sciences. All rights reserved.

EU ChemE 7 (15%) 4 (7%) 10 (20%) 14 (19%) 28 (29%) 58 (31%)


Nature

Asia ChemE 5 (10%) 3 (5%) 4 (8%) 6 (8%) 19 (20%) 75 (40%)


Canada ChemE 0 (0%) 0 (0%) 2 (4%) 2 (3%) 6 (6%) 17 (9%)
S. Am ChemE 0 (0%) 0 (0%) 0 (0%) 0 (0%) 4 (4%) 3 (2%)

Internationalization 4% 3% 20% 10% 37% 65%


NOTES: “Internationalization” indicates the minimum percentage overlap across regions. Not all countries are included in the defined regions,
so this is only an approximate measure of international collaboration.

259
New Directions for Chemical Engineering

260 New Directions for Chemical Engineering

The committee also made use of the Elsevier SciVal tool,1 particularly field-
weighted citation impact,2 which indicates how the number of citations of publications
from a region compares with the average number of citations of all other similar publica-
tions in the data universe, or how the citations of a particular publication compare with
the world average. Field-weighted citation impact was averaged for the 100 most cited
papers with an author having a chemical engineering affiliation published in each year
from each country or region within the comprehensive publication list described in Ap-
pendix C (Figure 10-2). In parallel with the other publication metrics reported above,
Asia, driven primarily by China, has surpassed other regions and countries in recent years.

FIGURE 10-2 Comparison of the field-weighted citation impact score for the top 100 cited pub-
lications from each region from the full list of journals with an author having a chemical engineer-
ing department affiliation. NOTE: 2005 is the first year in which South America published at least
100 papers in these journals, but it did not produce any notable outliers in earlier years.

1
See https://www.elsevier.com/solutions/scival.
2
See https://service.elsevier.com/app/answers/detail/a_id/28192/supporthub/scival/p/10961/28
192/.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

International Leadership 261

OBSERVATIONS

Beyond the above publication trends, the committee noted differences in chemi-
cal engineering advances across specific areas of the world. In energy, for example, the
United States is behind both China and Brazil in demonstrating large-scale technologies.
The United States remains the leader in catalysis research, which is usually found in chem-
istry departments in other countries. In the water–energy–food nexus, countries with wa-
ter scarcity (e.g., Australia, Israel, Saudi Arabia) have advanced R&D much further than
has the United States; in fact, in many places in the United States, aspects of the nexus
have been used as if they were not independent limited resources.
In the pharmaceutical space, threats from offshore manufacturing include both
challenges to the supply chain and concerns about quality control. While the United States
still leads the world in the invention and innovation of new therapeutic molecules, India,
China, and Singapore are leading the way in the economical production of biosimilar
molecules.
In the area of advanced recycling of plastics, the pyrolysis pathway has seen more
R&D in Europe than in the United States, a reflection of the difference in refinery capa-
bilities in the two regions. This regional difference has resulted in two trends: many more
pyrolysis companies are at pilot and demonstration scales in Europe than in the United
States, and more companies in the United States and Canada than in Europe are develop-
ing gasification recycling technologies.
In the area of plastics end-of-life management generally, the EU is ahead of the
United States in terms of policy drivers toward end-of-life management, as well as such
advances in waste conversion as anaerobic digestion and the use of municipal solid waste
for energy and other applications. Although there has been a recent increase in funding
for and interest in process intensification in the United States, progress in this area has for
an extended period been led largely by efforts in Europe.
With respect to the development of tools for future use, the United States leads
the world in high-performance computing and the development of new computer archi-
tectures. Most codes for molecular simulation were developed in the United States and
are open source. The United States also leads in the development of sensors. U.S. leader-
ship is clear as well in the development of tools for medicine (e.g., organs-on-chips), while
early work on in vitro models was done in Europe, driven mainly by the region’s ban on
animal testing of cosmetics.

CONCLUSION

The growth of research output from Asia has been driven mainly by the explosive
growth of publications from China (Figure 10-3). This growth reflects a dramatic increase
in that country’s research intensity and productivity, which was easy to anticipate simply
by extrapolating the 2007 study. More concerning for U.S. leadership in chemical engi-
neering and U.S. security is the quality of the work from China as measured both by cita-
tions (Figure 10-4) and by its appearance in elite journals (Figure 10-5). Likewise con-
cerning is patent analysis demonstrating that while the United States currently holds an

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

262 New Directions for Chemical Engineering

above-average revealed technological advantage in chemical engineering and adjacent in-


dustries, that advantage shrank over the period analyzed (2012–2014 compared with
2010–2012; Daiko et al., 2017).

FIGURE 10-3 Number of publications from the United States and China having at least one author
with a chemical engineering department affiliation over 5-year intervals from 1981 to 2020.
NOTE: “Original journals” includes all those used in the 2007 study (NRC, 2007) and “new jour-
nals” includes those added for the current analysis whose publication and/or indexing by Scopus
began in 2007 or thereafter.

FIGURE 10-4 Number of publications from the United States and China in the top 100 cited
articles in a given year over 5-year intervals from 1991 to 2020 within selected core chemical
engineering journals: AIChE Journal, Industrial & Engineering Chemistry Research (I and EC),
and Chemical Engineering Science (CES).

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

International Leadership 263

FIGURE 10-5 Number of publications from the United States and China having at least one author
with a chemical engineering department affiliation over 5-year intervals from 1981 to 2020 in
selected elite journals: Science, Proceedings of the National Academy of Sciences (PNAS), and
Nature.

There is little doubt regarding China’s commitment to research in areas central to


chemical engineering. Even a casual review of China’s Made in China 2025 plan
(McBride and Chatzky, 2019) shows a sharp focus on expanding technology and ad-
vanced manufacturing—in particular, electric and other alternative-energy vehicles, next-
generation information technology, artificial intelligence, new materials, and biomedi-
cine. China also emphasizes semiconductor development and manufacturing. All of these
research areas are discussed at various levels of detail in previous chapters of this report.
It is clear that anything less than vigorous and sustained investment in these and other
areas by the United States will by default cede leadership in technology to China in the
next 20 years (Mozur and Myers, 2021).
The increased research output from China is a result of large investments in tech-
nology areas, many of which are either central or highly relevant to chemical engineering,
investments that are approaching the gross expenditures on R&D of the United States
(Heney, 2020).
Similar levels of investment in the U.S. research enterprise are imperative. The
committee acknowledges that issues of international leadership are complex, and it is ex-
tremely challenging to find a balance between a healthy competitive national aim to lead
in a field and national security–driven aspirations of technological superiority, especially
given that national needs and priorities differ and can evolve with changing political lead-
ership.
Previous chapters outline the numerous opportunities for chemical engineers to
contribute in the areas of energy; food, water, and air; health and medicine; manufactur-
ing; materials; and tools. Without a sustained investment by federal research agencies

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

264 New Directions for Chemical Engineering

across these areas, it will be impossible for the United States to maintain leadership in
chemical engineering. At the same time, U.S. chemical engineering will be strengthened
through increased coordination and collaboration across disciplines, sectors, and political
boundaries. Almost all of the areas of research discussed in this report are multidiscipli-
nary in nature, and close collaboration will be required among researchers in academia
and government laboratories and industry practitioners to develop applications that are
economically viable and scalable. Such collaborations, as recommended throughout this
report, will have additional benefits for graduate education and faculty member develop-
ment while still satisfying the need of industry to achieve fast results.
International collaborations will continue to yield benefits in all the areas dis-
cussed herein. A good example of a format for collaboration is the Global Grand Chal-
lenges Summits organized by the National Academy of Engineering, the Chinese Acad-
emy of Engineering, and the Royal Academy of Engineering.3 This series of conferences
brings together professionals and students to explore solutions to the Grand Challenges
for Engineering, and could serve as a venue for creating and sustaining such collabora-
tions.

Recommendation 10-1: Across all areas of chemical engineering, in addition to ad-


vancing fundamental understanding, research investments should be set aside for
support of interdisciplinary, cross-sector, and international collaborations in the ar-
eas of energy; water, food, and air; health and medicine; manufacturing; materials
research; tools development; and beyond, with the goal of connecting U.S. research
to points of strength in other countries.

3
See http://www.engineeringchallenges.org/14500.aspx.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

References

Abdelkareem, M. A., M. El Haj Assad, E. T. Sayed, and B. Soudan. 2018. Recent progress in the
use of renewable energy sources to power water desalination plants. Desalination 435:97-
113. DOI: 10.1016/j.desal.2017.11.018.
Abejón, R., A. Garea, and A. Irabien. 2010. Ultrapurification of hydrogen peroxide solution from
ionic metals impurities to semiconductor grade by reverse osmosis. Separation and
Purification Technology 76(1):44-51. DOI: 10.1016/j.seppur.2010.09.018.
ACS and RSC (American Chemical Society and Royal Society of Chemistry). 1999. The discovery
and development of penicillin 1928-1945. Retrieved August 19, 2021, from https://www.
acs.org/content/dam/acsorg/education/whatischemistry/landmarks/flemingpenicillin/the-
discovery-and-development-of-penicillin-commemorative-booklet.pdf.
Ahmed, F. 2007. Profile of Mark E. Davis. Proceedings of the National Academy of Sciences
104(52). DOI: 10.1073/pnas.0704959105.
Ahmetović, E., Z. Kravanja, N. Ibrić, I. E. Grossmann, and L. E. Savulescu. 2021. State of the art
methods for combined water and energy systems optimisation in Kraft pulp mills.
Optimization and Engineering 22:1831-1852. DOI: 10.1007/s11081-021-09612-4.
AIChE (American Institute of Chemical Engineers). 2017. Achievements in the environment.
Retrieved August 17, 2021, from https://www.aiche.org/community/students/career-
resources-k-12-students-parents/what-do-chemical-engineers-do/saving-
environment/achievements.
AIChE. 2020. Thinking about climate. CEP Magazine 116(13). https://www.aiche.org/sites/
default/files/cep/20210214.pdf.
AIChE. 2021. 2021 AIChE Salary Survey. CEP Magazine. https://www.aiche.org/resources/
publications/cep/2021/june/2021-aiche-salary-survey.
Akee, R., M. R. Jones, and S. R. Porter. 2017. Race matters: Income Shares, income inequality,
and income mobility for all U.S. Races. National Bureau of Economic Research Working
Paper Series No. 23733. DOI: 10.3386/w23733.
Albrecht, T. R., A. Crootof, and C. A. Scott. 2018. The water-energy-food nexus: A systematic
review of methods for nexus assessment. Environmental Research Letters 13(4). DOI:
10.1088/1748-9326/aaa9c6.
Alexandratos, S. D. 2009. Ion-exchange resins: A retrospective from industrial and engineering
chemistry research. Industrial & Engineering Chemistry Research 48(1):388-398. DOI:
10.1021/ie801242v.
Alger, M., D. Velegol, and R. Shi. 2021. Sustainable energy corps: Building a global collaboration
to accelerate transition to a low carbon world. Chemical Engineering Science: X 10. DOI:
10.1016/j.cesx.2021.100099.
Alias, A., R. Abhijith, and V. Thankachan. 2019. Review on applications of smart glass in green
buildings. Presented at Green Buildings and Sustainable Engineering: Singapore.
Allison, G., Y. T. Cain, C. Cooney, T. Garcia, T. G. Bizjak, O. Holte, N. Jagota, B. Komas, E.
Korakianiti, D. Kourti, R. Madurawe, E. Morefield, F. Montgomery, M. Nasr, W.
Randolph, J. L. Robert, D. Rudd, and D. Zezza. 2015. Regulatory and quality
considerations for continuous manufacturing, May 20-21, 2014: Continuous
manufacturing symposium. Journal of Pharmaceutical Sciences 104(3):803-812. DOI:
10.1002/jps.24324.

265

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

266 New Directions for Chemical Engineering

Almudever, C. G., L. Lao, X. Fu, N. Khammassi, I. Ashraf, D. Iorga, S. Varsamopoulos, C. Eichler,


A. Wallraff, L. Geck, A. Kruth, J. Knoch, H. Bluhm, and K. Bertels. 2017. The
engineering challenges in quantum computing. Presented at Design, Automation & Test
in Europe Conference & Exhibition. Lausanne, Switzerland.
Alonso, C., T. Alig, J. Yoon, F. Bringezu, H. Warriner, and J. A. Zasadzinski. 2004. More than a
monolayer: Relating lung surfactant structure and mechanics to composition. Biophysical
Journal 87(6):4188-4202. DOI: 10.1529/biophysj.104.051201.
Altabet, Y. E., and P. G. Debenedetti. 2017. Communication: Relationship between local structure
and the stability of water in hydrophobic confinement. The Journal of Chemical Physics
147(24). DOI: 10.1063/1.5013253.
America Makes, AMT, and Deloitte (Association for Manufacturing Technology and Deloitte
Consulting). 2021. Assessing the role of additive manufacturing in support of the U.S.
COVID-19 response. Advanced Manufacturing Crisis Production Response.
https://www.fda.gov/media/150615/download.
Anastas, P. T., and J. C. Warner. 1998. Green chemistry: Theory and practice. New York: Oxford
University Press.
Anastas, P. T., and J. B. Zimmerman. 2003. Design through the 12 principles of green engineering.
Environmental Science & Technology 37(5):94A-101A. DOI: 10.1021/es032373g.
Andrew, R. M. 2018. Global CO2 emissions from cement production. Earth System Science Data
10:195-217. DOI: 10.5194/essd-10-195-2018.
Anna, S. L., N. Bontoux, and H. A. Stone. 2003. Formation of dispersions using “flow focusing”
in microchannels. Applied Physics Letters 82(3):364-366. DOI: 10.1063/1.1537519.
Anselmo, A. C., and S. Mitragotri. 2014. An overview of clinical and commercial impact of drug
delivery systems. Journal of Controlled Release 190:15-28. DOI: 10.1016/j.jconrel.
2014.03.053.
API (American Petroleum Institute). 2021. What are alternatives to make fracking less impactful?
Retrieved August 17, 2021, from https://www.api.org/oil-and-natural-gas/energy-
primers/hydraulic-fracturing/what-are-alternatives-to-make-fracking-less-impactful.
Ardo, S., D. Fernandez Rivas, M. A. Modestino, V. Schulze Greiving, F. F. Abdi, E. Alarcon
Llado, V. Artero, K. Ayers, C. Battaglia, J.-P. Becker, D. Bederak, A. Berger, F. Buda,
E. Chinello, B. Dam, V. Di Palma, T. Edvinsson, K. Fujii, H. Gardeniers, H. Geerlings,
S. M. H. Hashemi, S. Haussener, F. Houle, J. Huskens, B. D. James, K. Konrad, A. Kudo,
P. P. Kunturu, D. Lohse, B. Mei, E. L. Miller, G. F. Moore, J. Muller, K. L. Orchard, T.
E. Rosser, F. H. Saadi, J.-W. Schüttauf, B. Seger, S. W. Sheehan, W. A. Smith, J.
Spurgeon, M. H. Tang, R. van de Krol, P. C. K. Vesborg, and P. Westerik. 2018. Pathways
to electrochemical solar-hydrogen technologies. Energy & Environmental Science
11(10):2768-2783. DOI: 10.1039/c7ee03639f.
Arent, D. J., S. M. Bragg-Sitton, D. C. Miller, T. J. Tarka, J. A. Engel-Cox, R. D. Boardman, P. C.
Balash, M. F. Ruth, J. Cox, and D. J. Garfield. 2021. Multi-input, multi-output hybrid
energy systems. Joule 5(1):47-58. DOI: 10.1016/j.joule.2020.11.004.
Aresco. 2021. Microwave fracking: The new hydraulic fracturing? Retrieved August 17, 2021,
from https://www.arescotx.com/microwave-technology-the-new-fracking/.
Argus. 2020. EU ETS price €32-65/t under 2030 scenarios. Retrieved August 17, 2021, from
https://www.argusmedia.com/en/news/2142240-eu-ets-price-3265t-under-2030-
scenarios.
Arnold, F. H. 2018. Directed evolution: Bringing new chemistry to life. Angewandte Chemie
International Edition English 57(16):4143-4148. DOI: 10.1002/anie.201708408.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

References 267

ASEE (American Society for Engineering Education). 2020. Engineering & engineering
technology by the numbers 2019. Retrieved August 17, 2021, from https://ira.asee.org/
wp-content/uploads/2021/02/Engineering-by-the-Numbers-FINAL-2021.pdf.
Atkinson, R. D. 2018. Industry funding of university research: Which states lead? Retreived 2021,
from https://itif.org/publications/2018/01/08/industry-funding-university-research-which
-states-lead.
Auta, H. S., C. U. Emenike, and S. H. Fauziah. 2017. Distribution and importance of microplastics
in the marine environment: A review of the sources, fate, effects, and potential solutions.
Environment International 102:165-176. DOI: 10.1016/j.envint.2017.02.013.
AWEA (American Wind Energy Association). 2020. Wind powers America annual report 2019.
Retrieved August 18, 2021, from https://www.powermag.com/wp-content/uploads/2020/
04/awea_wpa_executivesummary2019.pdf.
Bachler, J., P. H. Handle, N. Giovambattista, and T. Loerting. 2019. Glass polymorphism and
liquid-liquid phase transition in aqueous solutions: Experiments and computer
simulations. Physical Chemistry Chemical Physics 21(42):23238-23268. DOI: 10.1039/
c9cp02953b.
Balicka, I. 2020. The food-energy-water nexus: A complex balance. Retrieved July 2020, from
https://www.aiche.org/chenected/2020/07/food-energy-water-nexus-complex-balance.
Banholzer, W. F., and M. E. Jones. 2013. Chemical engineers must focus on practical solutions.
AIChE Journal 59(8):2708-2720. DOI: 10.1002/aic.14172.
Barabino, G. A. 2021. Engineering solutions to COVID-19 and racial and ethnic health disparities.
Journal of Racial and Ethnic Health Disparities 8(2):277-279. DOI: 10.1007/s40615-
020-00953-x.
Bartel, M. A., J. R. Weinstein, and D. V. Schaffer. 2012. Directed evolution of novel adeno-
associated viruses for therapeutic gene delivery. Gene Therapy 19(6):694-700. DOI:
10.1038/gt.2012.20.
Bates, F. S., P. Brant, G. W. Coates, J. Lipson, C. Osuji, J. de Pablo, S. Rowan, R. Segalman, and
K. I. Winey. 2017. Frontiers in polymer science and engineering. NSF Workshop:
Frontiers in Polymer Science and Engineering. Arlington, VA. http://nsfpolymer
workshop2016.cems.umn.edu.
Batzner, S., T. Smidt, L. Sun, J. Mailoa, M. Kornbluth, N. Molinari, and B. Kozinsky. 2021. SE(3)-
equivariant graph neural networks for data-efficient and accurate interatomic potentials.
arXiv. DOI: 10.21203/rs.3.rs-244137/v1.
Bausch, M., C. Schultheiss, and J. B. Sieck. 2019. Recommendations for comparison of
productivity between fed-batch and perfusion processes. Biotechnology Journal 14(2).
DOI: 10.1002/biot.201700721.
Bazant, M. Z., and J. W. M. Bush. 2021. A guideline to limit indoor airborne transmission of
COVID-19. Proceedings of the National Academy of Sciences 118(17). DOI: 10.1073/
pnas.2018995118.
Beveridge, G. S. G., and R. S. Schechter. 1975. Optimization: Theory and practice (chemical
engineering). New York: McGraw-Hill Education.
Bezek, L. B., J. Pan, C. Harb, C. E. Zawaski, B. Molla, J. R. Kubalak, L. C. Marr, and C. B.
Williams. 2021. Additively manufactured respirators: Quantifying particle transmission
and identifying system-level challenges for improving filtration efficiency. Journal of
Manufacturing Systems 60:762-773. DOI: 10.1016/j.jmsy.2021.01.002.
Bhamla, M. S., B. Benson, C. Chai, G. Katsikis, A. Johri, and M. Prakash. 2017. Hand-powered
ultralow-cost paper centrifuge. Nature Biomedical Engineering 1(1). DOI: 10.1038/
s41551-016-0009.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

268 New Directions for Chemical Engineering

Bhatia, S. N., and D. E. Ingber. 2014. Microfluidic organs-on-chips. Nature Biotechnology


32(8):760-772. DOI: 10.1038/nbt.2989.
Bhojwani, S., K. Topolski, R. Mukherjee, D. Sengupta, and M. M. El-Halwagi. 2019. Technology
review and data analysis for cost assessment of water treatment systems. Science of the
Total Environment 651:2749-2761. DOI: 10.1016/j.scitotenv.2018.09.363.
Biddy, M. J. 2016. Chemicals from biomass: A market assessment of bioproducts with near-term
potential. Retrieved August 18, 2021, from https://www.energy.gov/sites/prod/files/2016/
11/f34/biddy_bioenergy_2016.pdf.
Bielenberg, J., and M. Bryner. 2018. Realize the potential of process intensification. Retrieved
March 2018, from https://www.aiche.org/sites/default/files/cep/20180341.pdf.
Biggs, E. M., E. Bruce, B. Boruff, J. M. A. Duncan, J. Horsley, N. Pauli, K. McNeill, A. Neef, F.
Van Ogtrop, J. Curnow, B. Haworth, S. Duce, and Y. Imanari. 2015. Sustainable
development and the water–energy–food nexus: A perspective on livelihoods.
Environmental Science & Policy 54:389-397. DOI: 10.1016/j.envsci.2015.08.002.
Billiet, S., and S. R. Trenor. 2020. 100th Anniversary of macromolecular science viewpoint: Needs
for plastics packaging circularity. ACS Macro Letters 9(9):1376-1390. DOI: 10.1021/
acsmacrolett.0c00437.
(BIO) Biotechnology Innovation Organization. 2011. New study shows the rate of drug approvals
lower than previously reported. Retrieved August 17, 2021, from https://archive.bio.org/
media/press-release/new-study-shows-rate-drug-approvals-lower-previously-reported.
Bioeconomy Capital. 2018. Bioeconomy dashboard: Economic metrics. Retrieved August 17,
2021, from http://www.bioeconomycapital.com/bioeconomy-dashboard.
Bird, R. B., Warren E. Stewart, and Edwin N. Lightfoot. 1960. Transport phenomena. New York:
John Wiley and Sons, Inc.
Bischoff, K. B. 2015. Pharmacokinetics and cancer chemotherapy. Journal of Pharmacokinetics
and Biopharmaceutics 1(6):465-480. DOI: 10.1007/bf01059786.
Bjorneholm, O., M. H. Hansen, A. Hodgson, L. M. Liu, D. T. Limmer, A. Michaelides, P.
Pedevilla, J. Rossmeisl, H. Shen, G. Tocci, E. Tyrode, M. M. Walz, J. Werner, and H.
Bluhm. 2016. Water at interfaces. Chemical Reviews 116(13):7698-7726. DOI: 10.1021/
acs.chemrev.6b00045.
BloombergNEF. 2020. Battery pack prices cited below $100/kWh for the first time in 2020, while
market average sits at $137/kWh. Retrieved August 17, 2021, from https://about.bnef.
com/blog/battery-pack-prices-cited-below-100-kwh-for-the-first-time-in-2020-while-
market-average-sits-at-137-kwh.
BLS (U.S. Bureau of Labor Statistics). 2021. Occupational outlook handbook—Chemical
engineers. Retrieved August 16, 2021, from https://www.bls.gov/ooh/architecture-and-
engineering/chemical-engineers.htm#tab-1.
Boerner, L. K. 2019. Industrial ammonia production emits more CO2 than any other chemical-
making reaction. Chemists want to change that. Chemical & Engineering News 97(24).
https://cen.acs.org/environment/green-chemistry/Industrial-ammonia-production-emits-
CO2/97/i24.
Boger, T., S. Roy, A. K. Heibel, and O. Borchers. 2003. A monolith loop reactor as an attractive
alternative to slurry reactors. Catalysis Today 79(2):441-451. DOI: 10.1016/s0920-
5861(03)00058-0.
Bozell, J. J., and G. R. Petersen. 2010. Technology development for the production of biobased
products from biorefinery carbohydrates—the US Department of Energy’s “Top 10”
revisited. Green Chemistry 12(4). DOI: 10.1039/b922014c.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

References 269

BP. 2019. BP statistical review of world energy, 68th edition. Retrieved August 17, 2021, from
https://www.bp.com/content/dam/bp/business-sites/en/global/corporate/pdfs/energy-eco
nomics/statistical-review/bp-stats-review-2019-full-report.pdf.
Brady, J. R., and J. C. Love. 2021. Alternative hosts as the missing link for equitable therapeutic
protein production. Nature Biotechnology 39(4):404-407. DOI: 10.1038/s41587-021-
00884-w.
Branger, C., W. Meouche, and A. Margaillan. 2013. Recent advances on ion-imprinted polymers.
Reactive and Functional Polymers 73(6):859-875. DOI: 10.1016/j.reactfunctpolym.
2013.03.021.
Broekhuis, R. R., R. M. Machado, and A. F. Nordquist. 2001. The ejector-driven monolith loop
reactor—Experiments and modeling. Catalysis Today 69:87-93. DOI: 10.1016/s0920-
5861(01)00358-3.
Brown, K. W., B. Gessesse, L. J. Butler, and D. L. MacIntosh. 2017. Potential effectiveness
of point-of-use filtration to address risks to drinking water in the United States.
Environmental Health Insights 11. DOI: 10.1177/1178630217746997.
Brown, S. A., B. P. Kovatchev, D. Raghinaru, J. W. Lum, B. A. Buckingham, Y. C. Kudva, L. M.
Laffel, C. J. Levy, J. E. Pinsker, R. P. Wadwa, E. Dassau, F. J. Doyle, S. M. Anderson,
M. M. Church, V. Dadlani, L. Ekhlaspour, G. P. Forlenza, E. Isganaitis, D. W. Lam, C.
Kollman, and R. W. Beck. 2019. Six-month randomized, multicenter trial of closed-loop
control in type 1 diabetes. New England Journal of Medicine 381(18):1707-1717. DOI:
10.1056/NEJMoa1907863.
Brown, T. D., K. A. Whitehead, and S. Mitragotri. 2020. Materials for oral delivery of proteins
and peptides. Nature Reviews Materials 5(2):127-148. DOI: 10.1038/s41578-019-0156-6.
Buchner, G. A., K. J. Stepputat, A. W. Zimmermann, and R. Schomäcker. 2019. Specifying
technology readiness levels for the chemical industry. Industrial & Engineering
Chemistry Research 58(17):6957-6969. DOI: 10.1021/acs.iecr.8b05693.
Bui, M., C. S. Adjiman, A. Bardow, E. J. Anthony, A. Boston, S. Brown, P. S. Fennell, S. Fuss, A.
Galindo, L. A. Hackett, J. P. Hallett, H. J. Herzog, G. Jackson, J. Kemper, S. Krevor, G.
C. Maitland, M. Matuszewski, I. S. Metcalfe, C. Petit, G. Puxty, J. Reimer, D. M. Reiner,
E. S. Rubin, S. A. Scott, N. Shah, B. Smit, J. P. M. Trusler, P. Webley, J. Wilcox, and N.
Mac Dowell. 2018. Carbon capture and storage (CCS): The way forward. Energy &
Environmental Science 11(5):1062-1176. DOI: 10.1039/C7EE02342A.
Buntz, B. 2021. Pharma's top 20 R&D spenders in 2020. Drug Discovery and Development. Re-
trieved 2021, from https://www.drugdiscoverytrends.com/pharmas-top-20-rd-spenders-
in-2020/.
Buzby, J. C., Hodan F. Wells, and J. Hyman. 2014. The estimated amount, value, and calories of
postharvest food losses at the retail and consumer levels in the United States. U.S.
Department of Agriculture, Economic Research Service EIB-121.
Byrne, C., K. M. Zahra, S. Dhaliwal, D. C. Grinter, K. Roy, W. Q. Garzon, G. Held, G. Thornton,
and A. S. Walton. 2021. A combined laboratory and synchrotron in-situ photoemission
study of the rutile TiO2 (110)/water interface. Journal of Physics D: Applied Physics
54(19). DOI: 10.1088/1361-6463/abddfb.
C&EN (Chemical and Engineering News). 2021. U.S. top 50 chemical companies of 2021. Re-
trieved December 3, 2021, from https://cen.acs.org/sections/us-top-50.html.
Caballero, J. A., J. A. Labarta, N. Quirante, A. Carrero-Parreño, and I. E. Grossmann. 2020.
Environmental and economic water management in shale gas extraction. Sustainability
12(4). DOI: 10.3390/su12041686.
Cao, L., I. K. M. Yu, Y. Liu, X. Ruan, D. C. W. Tsang, A. J. Hunt, Y. S. Ok, H. Song, and S.
Zhang. 2018a. Lignin valorization for the production of renewable chemicals: State-of-

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

270 New Directions for Chemical Engineering

the-art review and future prospects. Bioresource Technology 269:465-475. DOI:


10.1016/j.biortech.2018.08.065.
Cao, Q., M. Huang, T. H. Kuehn, L. Shen, W.-Q. Tao, J. Cao, and D. Y. H. Pui. 2018b. Urban-
scale SALSCS, Part II: A Parametric Study of System Performance. Aerosol and Air
Quality Research 18(11):2879-2894. DOI: 10.4209/aaqr.2018.06.0239.
Cao, Y., J. Romero, J. P. Olson, M. Degroote, P. D. Johnson, M. Kieferova, I. D. Kivlichan, T.
Menke, B. Peropadre, N. P. D. Sawaya, S. Sim, L. Veis, and A. Aspuru-Guzik. 2019.
Quantum chemistry in the age of quantum computing. Chemical Review 119(19):10856-
10915. DOI: 10.1021/acs.chemrev.8b00803.
Carayannis, E. G., and J. Alexander. 2004. Strategy, structure, and performance issues of
precompetitive R&D consortia: Insights and lessons learned from SEMATECH. IEEE
Transactions on Engineering Management 51(2):226-232. DOI: 10.1109/TEM.
2003.822459.
Carlson, R. 2016. Estimating the biotech sector's contribution to the US economy. Nature
Biotechnology 34(3):247-255. DOI: 10.1038/nbt.3491.
Carnahan, B., H. A. Luther, and J. O. Wilkes. 1969. Applied numerical methods. New York: Wiley.
Carter, P. J. 2011. Introduction to current and future protein therapeutics: A protein engineering
perspective. Experimental Cell Research 317(9):1261-1269. DOI: 10.1016/j.yexcr.
2011.02.013.
Catalanotti, S., V. Cuomo, G. Piro, D. Ruggi, V. Silvestrini, and G. Troise. 1975. The radiative
cooling of selective surfaces. Solar Energy 17(2):83-89. DOI: 10.1016/0038-092X(75)
90062-6.
Caudill, C. L., J. L. Perry, S. Tian, J. C. Luft, and J. M. DeSimone. 2018. Spatially controlled
coating of continuous liquid interface production microneedles for transdermal protein
delivery. Journal of Controlled Release 284:122-132. DOI: 10.1016/j.jconrel.2018.05.
042.
Cech, E. 2013. The (mis)framing of social justice: Why ideologies of depoliticization and
meritocracy hinder engineers’ ability to think about social injustices. In Engineering
education for social justice. Dordrecht, Netherlands: Springer. DOI 10.1007/978-94-007-
6350-0_4.
Cech, E. A., and M. Blair-Loy. 2010. Perceiving glass ceilings? Meritocratic versus structural
explanations of gender inequality among women in science and technology. Social
Problems 57(3):371-397. DOI: 10.1525/sp.2010.57.3.371.
Celik, G., R. M. Kennedy, R. A. Hackler, M. Ferrandon, A. Tennakoon, S. Patnaik, A. M.
LaPointe, S. C. Ammal, A. Heyden, F. A. Perras, M. Pruski, S. L. Scott, K. R.
Poeppelmeier, A. D. Sadow, and M. Delferro. 2019. Upcycling single-use polyethylene
into high-quality liquid products. ACS Central Science 5(11):1795-1803. DOI:
10.1021/acscentsci.9b00722.
Centi, G., G. Iaquaniello, and S. Perathoner. 2019. Chemical engineering role in the use of
renewable energy and alternative carbon sources in chemical production. BMC Chemical
Engineering 1. DOI: 10.1186/s42480-019-0006-8.
Cescon, M., S. Deshpande, R. Nimri, I. F. J. Doyle, and E. Dassau. 2021. Using iterative learning
for insulin dosage optimization in multiple-daily-injections therapy for people with type
1 diabetes. IEEE Transactions on Biomedical Engineering 68(2):482-491. DOI:
10.1109/TBME.2020.3005622.
Chae, D., H. Lim, S. So, S. Son, S. Ju, W. Kim, J. Rho, and H. Lee. 2021. Spectrally selective
nanoparticle mixture coating for passive daytime radiative cooling. ACS Applied
Materials & Interfaces 13(18):21119-21126. DOI: 10.1021/acsami.0c20311.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

References 271

Chanoca, A., L. de Vries, and W. Boerjan. 2019. Lignin engineering in forest trees. Frontiers in
Plant Science 10:912. DOI: 10.3389/fpls.2019.00912.
Chen, F., and R. A. Dixon. 2007. Lignin modification improves fermentable sugar yields for
biofuel production. Nature Biotechnology 25(7):759-761. DOI: 10.1038/nbt1316.
Chen, D., L. Yin, H. Wang, and P. He. 2014. Pyrolysis technologies for municipal solid waste: A
review. Waste Management 34(12):2466-2486. DOI: 10.1016/j.wasman.2014.08.004.
Cherkasov, A., E. N. Muratov, D. Fourches, A. Varnek, I. I. Baskin, M. Cronin, J. Dearden, P.
Gramatica, Y. C. Martin, R. Todeschini, V. Consonni, V. E. Kuz’min, R. Cramer, R.
Benigni, C. Yang, J. Rathman, L. Terfloth, J. Gasteiger, A. Richard, and A. Tropsha.
2014. QSAR Modeling: Where have you been? Where are you going to? Journal of
Medicinal Chemistry 57(12):4977-5010. DOI: 10.1021/jm4004285.
Chetty, R., N. Hendren, F. LIn, J. Majerovitz, and B. Scuderi. 2016. Childhood environment and
gender gaps in adulthood. Cambridge, MA: National Bureau of Economic Research.
DOI: 10.3386/w21936.
Chetty, R., N. Hendren, M. R. Jones, and S. R. Porter. 2019. Race and economic opportunity in
the United States: An intergenerational perspective. The Quarterly Journal of Economics
135(2):711-783. DOI: 10.1093/qje/qjz042.
CHiMaD (Center for Hierarchical Materials Design). 2021. Polymer property predictor and
database. Retrieved August 27, 2021, from https://pppdb.uchicago.edu.
Chmiela, S., H. E. Sauceda, K.-R. Müller, and A. Tkatchenko. 2018. Towards exact molecular
dynamics simulations with machine-learned force fields. Nature Communications 9,
Article 3887. DOI: 10.1038/s41467-018-06169-2.
Chowdhury, S., and S. S. Fong. 2020. Leveraging genome-scale metabolic models for human
health applications. Current Opinion in Biotechnology 66:267-276. DOI:
10.1016/j.copbio.2020.08.017.
Christopherson, D. A., W. C. Yao, M. Lu, R. Vijayakumar, and A. R. Sedaghat. 2020. High-
efficiency particulate air filters in the era of COVID-19: Function and efficacy.
Otolaryngology–Head and Neck Surgery 163(6):1153-1155. DOI: 10.1177/01945998
20941838.
Chung, M., G. Fortunato, and N. Radacsi. 2019. Wearable flexible sweat sensors for healthcare
monitoring: A review. Journal of the Royal Society Interface 16(159). DOI: 10.1098/
rsif.2019.0217.
CMS (Centers for Medicare & Medicaid Services). 2020. The National Health Expenditure
Accounts (NHEA). Retrieved August 19, 2021, from https://www.cms.gov/Research-
Statistics-Data-and-Systems/Statistics-Trends-and-Reports/NationalHealthExpendData/
NationalHealthAccountsHistorical.
Cocchi, M., D. D. Angelis, L. Mazzeo, P. Nardozi, V. Piemonte, R. Tuffi, and S. Vecchio Ciprioti.
2020. Catalytic pyrolysis of a residual plastic waste using zeolites produced by coal fly
ash. Catalysts 10(10). DOI: 10.3390/catal10101113.
Cohen, C. 1996. The early history of chemical engineering: A reassessment. British Journal for
the History of Science 29(2):171-194. DOI: 10.1017/s000708740003421x.
Cohen, Y., and J. Glater. 2010. A tribute to Sidney Loeb—The pioneer of reverse osmosis
desalination research. Desalination and Water Treatment 15:222-227. DOI: 10.5004/
dwt.2010.1762.
Coley, C. W., W. H. Green, and K. F. Jensen. 2018. Machine learning in computer-aided synthesis
planning. Accounts of Chemical Research 51(5):1281-1289. DOI: 10.1021/acs.
accounts.8b00087.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

272 New Directions for Chemical Engineering

Collias, D. I., A. M. Harris, V. Nagpal, I. W. Cottrell, and M. W. Schultheis. 2014. Biobased


terephthalic acid technologies: A literature review. Industrial Biotechnology 10(2):91-
105. DOI: 10.1089/ind.2014.0002.
Collias, D. I., J. E. Godlewski, and J. E. Velasquez. 2018. Method of making acrylic acid from
hydroxypropionic acid. U.S. Patent No. 9,890,102. Washington, DC: U.S. Patent and
Trademark Office.
Collias, D. I., M. I. James, and J.M. Jayman. 2021. Circular economy of polymers: Topics in
recycling technologies. ACS Symposium Series 1391. DOI: 10.1021/bk-2021-
1391.ch001.
Colman, R. J., and D. T. Rubin. 2014. Fecal microbiota transplantation as therapy for inflammatory
bowel disease: A systematic review and meta-analysis. Journal of Crohn’s and Colitis
8(12):1569-1581. DOI: 10.1016/j.crohns.2014.08.006.
Colton, C. K., ed. 1991. Advances in chemical engineering. San Diego, CA: Academic Press.
Conn, R. W., M. M. Crow, C. M. Friend, and M. McNutt. July 21, 2021. The next 75 years of US
science and innovation policy: An introduction. Issues in Science and Technology.
https://issues.org/the-next-75-years-of-us-science-and-innovation-policy-an-introduction.
Coronell, D. G., T. H.-L. Hsiung, J. Howard Paul Withers, and A. J. Woytek. 1997. Process for
nitrogen trifluoride synthesis. EP Patent No. 0787684B1. Munich, Germany: European
Patent Office.
Correa-Baena, J. P., M. Saliba, T. Buonassisi, M. Gratzel, A. Abate, W. Tress, and A. Hagfeldt.
2017. Promises and challenges of perovskite solar cells. Science 358(6364):739-744.
DOI: 10.1126/science.aam6323.
Cortes Garcia, G. E., J. van der Schaaf, and A. A. Kiss. 2017. A review on process intensification
in HiGee distillation. Journal of Chemical Technology & Biotechnology 92(6):1136-
1156. DOI: 10.1002/jctb.5206.
Coughanowr, D. R., and L. B. Koppel. 1965. Process systems analysis and control. New York:
McGraw-Hill Book Company, Inc.
Creative Energy. 2007. European roadmap for process intensification. Retrieved August 17, 2021,
from https://efce.info/efce_media/-p-531.pdf.
Crocker, J. C., M. T. Valentine, E. R. Weeks, T. Gisler, P. D. Kaplan, A. G. Yodh, and D. A. Weitz.
2000. Two-point microrheology of inhomogeneous soft materials. Physical Review
Letters 85(4):888-891. DOI: 10.1103/PhysRevLett.85.888.
Crone, B. C., T. F. Speth, D. G. Wahman, S. J. Smith, G. Abulikemu, E. J. Kleiner, and J. G.
Pressman. 2019. Occurrence of per- and polyfluoroalkyl substances (PFAS) in source
water and their treatment in drinking water. Critical Reviews in Environmental Science
and Technology 49(24):2359-2396. DOI: 10.1080/10643389.2019.1614848.
Cruz, A. C., A. L. Medel, A. C. Bianchi, V. Wong, and M. Danforth. 2021. Impact of flipped
classroom model on high-workload and low-income students in upper-division computer
science. Presented at ASEE Virtual Annual Conference.
Cui, S., Y. Wang, Y. Yang, F. M. Nave, and K. T. Harris. 2011. Connecting incoming freshmen
with engineering through hands-on projects. American Journal of Engineering Education
2(2):31-42. DOI: 10.19030/ajee.v2i2.6636.
Cybulski, A., and J. A. Moulijn. 1994. Monoliths in heterogeneous catalysis. Catalysis Reviews
36(2):179-270. DOI: 10.1080/01614949408013925.
Cywar, R., N. A. Rorrerm, C. A. Hoyt, G. T. Beckham, and E. Chen. 2021. Bio-based polymers
with performance-advantaged properties. Nature Reviews Materials. In press. DOI:
10.1038/s41578-021-00363-3.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

References 273

Daglar, H., and S. Keskin. 2020. Recent advances, opportunities, and challenges in high-
throughput computational screening of MOFs for gas separations. Coordination
Chemistry Reviews 422. DOI: 10.1016/j.ccr.2020.213470.
Dahlman, J. E., K. J. Kauffman, Y. Xing, T. E. Shaw, F. F. Mir, C. C. Dlott, R. Langer, D. G.
Anderson, and E. T. Wang. 2017. Barcoded nanoparticles for high throughput in vivo
discovery of targeted therapeutics. Proceedings of the National Academy of Sciences
114(8). DOI: 10.1073/pnas.1620874114.
Daiko, T., H. Dernis, M. Dosso, P. Gkotsis, M. Squicciarini, and A. Vezzani. 2017. World
corporate R&D investors: Industrial property strategies in the digital economy. A JRC
and OECD Common Report. Luxembourg: Publications office of the European Union.
Dan, Y., Y. Zhao, X. Li, S. Li, M. Hu, and J. Hu. 2020. Generative adversarial networks (GAN)
based efficient sampling of chemical composition space for inverse design of inorganic
materials. NPJ Computational Materials 6(1). DOI: 10.1038/s41524-020-00352-0.
Das, T., and H. Cabezas. 2018. Tools and concepts for environmental sustainability in the food-
energy-water nexus: Chemical engineering perspective. Environmental Progress &
Sustainable Energy 37(1):73-81. DOI: 10.1002/ep.12763.
Dassau, E., E. Renard, J. Place, A. Farret, M. J. Pelletier, J. Lee, L. M. Huyett, A. Chakrabarty, F.
J. Doyle 3rd, and H. C. Zisser. 2017. Intraperitoneal insulin delivery provides superior
glycaemic regulation to subcutaneous insulin delivery in model predictive control-based
fully-automated artificial pancreas in patients with type 1 diabetes: A pilot study.
Diabetes, Obesity and Metabolism 19(12):1698-1705. DOI: 10.1111/dom.12999.
Davis, R., L. Tao, E. C. D. Tan, M. J. Biddy, G. T. Beckham, C. Scarlata, J. Jacobson, K. Cafferty,
J. Ross, J. Lukas, D. Knorr, and P. Schoen. 2013. Process design and economics for the
conversion of lignocellulosic biomass to hydrocarbons: Dilute acid and enzymatic decon-
struction of biomass to sugars and biological conversion of sugars to hydrocarbons.
Golden, CO: National Renewable Energy Laboratory.
de Beer, M. P., H. L. van der Laan, M. A. Cole, R. J. Whelan, M. A. Burns, and T. F. Scott. 2019.
Rapid, continuous additive manufacturing by volumetric polymerization inhibition
patterning. Science Advances 5(1). DOI: 10.1126/sciadv.aau8723.
de Pee, A., D. Pinner, O. Roelofsen, K. Somers, E. Speelman, and M. Witteveen. 2018. Decarbon-
ization of industrial sectors: The next frontier. Amsterdam: McKinsey.
Debenedetti, P. G. 2003. Supercooled and glassy water. Journal of Physics: Condensed Matter
15(45):R1669-1726. DOI: 10.1088/0953-8984/15/45/r01.
Debenedetti, P. G., and M. L. Klein. 2017. Chemical physics of water. Proceedings of the National
Academy of Sciences 114(51):13325-13326. DOI: 10.1073/pnas.1719350115.
Decante, G., J. B. Costa, J. Silva-Correia, M. N. Collins, R. L. Reis, and J. M. Oliveira. 2021.
Engineering bioinks for 3D bioprinting. Biofabrication 13(3). DOI: 10.1088/1758-
5090/abec2c.
Denn, M. M. 1991. The identity of our profession. In Advances in chemical engineering. C. K.
Colton, ed. San Diego, CA: Academic Press.
Deri, M. A., P. Mills, and D. McGregor. 2018. Structure and evaluation of a flipped general
chemistry course as a model for small and large gateway science courses at an urban
public institution. Journal of College Science Teaching 47(3):68-77.
DHS (Department of Homeland Security). 2019. Chemical sector profile. Retrieved August 16,
2021, from https://www.cisa.gov/sites/default/files/publications/Chemical-Sector-Pro-
file_Final%20508.pdf.
DiMasi, J. A., H. G. Grabowski, and R. W. Hansen. 2016. Innovation in the pharmaceutical
industry: New estimates of R&D costs. Journal of Health Economics 47:20-33. DOI:
10.1016/j.jhealeco.2016.01.012.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

274 New Directions for Chemical Engineering

Dixon, K., and J. L. Wendt. 2021. Science motivation and achievement among minority urban high
school students: An examination of the flipped classroom model. Journal of Science
Education and Technology 30(5):642-657. DOI: 10.1007/s10956-021-09909-0.
Dobbelaere, M. R., P. P. Plehiers, R. Van de Vijver, C. V. Stevens, and K. M. Van Geem. 2021.
Machine learning in chemical engineering: Strengths, weaknesses, opportunities, and
threats. Engineering 7(9):1201-1211. DOI: 10.1016/j.eng.2021.03.019.
D’Odorico, P., K. F. Davis, L. Rosa, J. A. Carr, D. Chiarelli, J. Dell’Angelo, J. Gephart, G. K.
MacDonald, D. A. Seekell, S. Suweis, and M. C. Rulli. 2018. The global food-energy-
water nexus. Reviews of Geophysics 56(3):456-531. DOI: 10.1029/2017rg000591.
DOE (U.S. Department of Energy). 2014. The water-energy nexus: Challenges and opportunities.
Retrieved August 16, 2021, from https://www.energy.gov/articles/water-energy-nexus-
challenges-and-opportunities.
DOE. 2015. Revolution…now—The future arrives for five clean energy technologies. Update
retrieved August 16, 2021, from https://www.energy.gov/sites/prod/files/2015/11/f27/
Revolution-Now-11132015.pdf.
DOE. 2016. 2016 Billion-ton report—Advancing domestic resources for a thriving bioeconomy.
Retrieved August 18, 2021, from https://www.energy.gov/sites/default/files/2016/12/f34/
2016_billion_ton_report_12.2.16_0.pdf.
DOE. 2018a. Moving beyond drop-in replacements: Performance-advantaged biobased chemicals.
Retrieved August 16, 2021, from https://www.energy.gov/sites/prod/files/2018/06/f53/
Performance-Advantaged%20Biobased%20Chemicals%20Workshop%20Report.pdf.
DOE. 2018b. R&D Opportunities for natural gas technologies in building applications. Retrieved
August 16, 2021, from https://www.energy.gov/sites/prod/files/2018/08/f55/bto-Natural-
Gas-RD-Opportunities-082918.pdf.
DOE. 2018c. Accelerating breakthrough innovation in carbon capture, utilization, and storage.
Retrieved 2021, https://www.energy.gov/fe/downloads/accelerating-breakthrough-inno
vation-carbon-capture-utilization-and-storage.
DOE. 2021a. Solar energy technologies office multi-year program plan. Retrieved August 16,
2021, from https://www.energy.gov/eere/solar/articles/solar-energy-technologies-office-
multi-year-program-plan.
DOE. 2021b. Shale research & development. Retrieved August 16, 2021, from
https://www.energy.gov/fe/science-innovation/oil-gas-research/shale-gas-rd.
DOE. 2021c. Nuclear reactor technologies. Retrieved August 16, 2021, from
https://www.energy.gov/ne/nuclear-reactor-technologies.
Doyle, F. J., 3rd, L. M. Huyett, J. B. Lee, H. C. Zisser, and E. Dassau. 2014. Closed-loop artificial
pancreas systems: Engineering the algorithms. Diabetes Care 37(5):1191-1197. DOI:
10.2337/dc13-2108.
Du Bois, W. E. B. 1939. The negro scientist. The American Scholar 8(3):309-320.
http://www.jstor.org/stable/41204425.
Eberle, A., A. Bhatt, Y. Zhang, and G. Heath. 2017. Potential air pollutant emissions and permitting
classifications for two biorefinery process designs in the United States. Environmental
Science & Technology 51(11):5879-5888. DOI: 10.1021/acs.est.7b00229.
Edwards, D. A., D. Ausiello, J. Salzman, T. Devlin, R. Langer, B. J. Beddingfield, A. C. Fears, L.
A. Doyle-Meyers, R. K. Redmann, S. Z. Killeen, N. J. Maness, and C. J. Roy. 2021.
Exhaled aerosol increases with COVID-19 infection, age, and obesity. Proceedings of the
National Academy of Sciences 118(8). DOI: 10.1073/pnas.2021830118.
Eggleton, C. D., T.-M. Tsai, and K. J. Stebe. 2001. Tip streaming from a drop in the presence of
surfactants. Physical Review Letters 87(4). DOI: 10.1103/PhysRevLett.87.048302.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

References 275

EIA (U.S. Energy Information Administration). 2013. Few transportation fuels surpass the energy
densities of gasoline and diesel. Retrieved July 18, 2021, from https://www.eia.
gov/todayinenergy/detail.php?id=9991.
EIA. 2020a. U.S. fuel ethanol production capacity increased by 3% in 2019. Retrieved August 16,
2021, from https://www.eia.gov/todayinenergy/detail.php?id=45316.
EIA. 2020b. International energy outlook 2020. Retrieved August 16, 2021, from https://www.eia.
gov/outlooks/ieo/.
EIA. 2020c. Biomass explained: Waste-to-energy (municipal solid waste). Retrieved August 16,
2021, from https://www.eia.gov/energyexplained/biomass/waste-to-energy.php.
EIA. 2020d. U.S. coal-fired electricity generation in 2019 falls to 42-year low. Retrieved August
16, 2021, from https://www.eia.gov/todayinenergy/detail.php?id=43675#.
EIA. 2021a. Global ethanol production by country or region. Retrieved August 16, 2021, from
https://afdc.energy.gov/data/10331.
EIA. 2021b. Annual energy outlook 2021. U.S. Department of Energy, Washington, DC.
EIA. 2021c. How much shale (tight) oil is produced in the United States? Retrieved August 16,
2021, from https://www.eia.gov/tools/faqs/faq.php?id=847&t=6.
EIA. 2021d. Short-term energy outlook: Global liquid fuels. Retrieved August 9, 2021, from
https://www.eia.gov/outlooks/steo/report/global_oil.php.
EIA. 2021e. Natural gas explained: Use of natural gas. Retrieved August 31, 2021, from
https://www.eia.gov/energyexplained/natural-gas/use-of-natural-gas.php.
EIA. 2021f. Natural gas explained: Delivery and storage of natural gas. Retrieved August 31, 2021,
from https://www.eia.gov/energyexplained/natural-gas/delivery-and-storage.php.
EIA. 2021g. Natural gas explained. Retrieved September 11, 2021, from https://www.eia.gov/
energyexplained/natural-gas.
EIA. 2021h. Monthly energy review, August 2021. Washington, DC: Office of Energy Statistics,
U.S. Department of Energy.
EIA. 2021i. Frequently asked questions: How much carbon dioxide is produced when different
fuels are burned? Retrieved January 31, 2022 from https://www.eia.gov/tools/faqs/faq.
php?id=73&t=11.
EIA. 2021j. Oil and petroleum products explained: Where our oil comes from. Retrieved August
16, 2021, from https://www.eia.gov/energyexplained/oil-and-petroleum-products/where-
our-oil-comes-from.php.
EIA. 2021k. Oil and petroleum products explained: Use of oil. Retrieved August 16, 2021, from
https://www.eia.gov/energyexplained/oil-and-petroleum-products/use-of-oil.php.
Elbashier, E., A. Mussa, M. Hafiz, and A. H. Hawari. 2021. Recovery of rare earth elements from
waste streams using membrane processes: An overview. Hydrometallurgy 204. DOI:
10.1016/j.hydromet.2021.105706.
Elhacham, E., L. Ben-Uri, J. Grozovski, Y. M. Bar-On, and R. Milo. 2020. Global human-made
mass exceeds all living biomass. Nature 588(7838):442-444. DOI: 10.1038/s41586-020-
3010-5.
Elton, D. C., Z. Boukouvalas, M. D. Fuge, and P. W. Chung. 2019. Deep learning for molecular
design—A review of the state of the art. Molecular Systems Design & Engineering
4(4):828-849. DOI: 10.1039/c9me00039a.
EMF (Ellen MacArthur Foundation). 2013a. Towards the circular economy, volume 2: Opportu-
nities for the consumer goods sector. Cowes, United Kingdom. https://emf.thirdlight.
com/link/coj8yt1jogq8-hkhkq2/@/preview/1?o.
EMF. 2013b. Towards the circular economy, volume 1: Economic and business rationale for an
accelerated transition. Cowes, United Kingdom. https://emf.thirdlight.com/link/x8ay
372a3r11-k6775n/@/preview/1?o.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

276 New Directions for Chemical Engineering

EMF. 2014. Towards the circular economy, volume 3: Accelerating the scale up across global
supply chains. Cowes, United Kingdom. https://emf.thirdlight.com/link/t4gb0fs4knot-
n8nz6f/@/preview/1?o.
EMF. 2016. The new plastics economy: Rethinking the future of plastics. Retrieved September 14,
2021, 2021, from https://emf.thirdlight.com/link/faarmdpz93ds-5vmvdf/@/preview/1?o.
EMF. 2017a. What is the circular economy? Retrieved August 17, 2021, from https://archive.
ellenmacarthurfoundation.org/circular-economy/what-is-the-circular-economy.
EMF. 2017b. The circular economy in detail. Retrieved August 17, 2021, from https://archive.
ellenmacarthurfoundation.org/explore/the-circular-economy-in-detail.
EMF. 2019. Circular economy diagram. Retrieved 2021, from https://ellenmacarthurfoundation.
org/circular-economy-diagram.
EMF. 2021. Circular economy introduction. Retrieved August 17, 2021, from https://www.ellen
macarthurfoundation.org/circular-economy/concept.
EPA (U.S. Environmental Protection Agency). 2018. Indoor air quality—What are the trends in
indoor air quality and their effects on human health? Retrieved August 16, 2021, from
https://www.epa.gov/report-environment/indoor-air-quality#note1.
EPA. 2020. Health, energy efficiency and climate change. Retrieved August 16, 2021, from
https://www.epa.gov/indoor-air-quality-iaq/health-energy-efficiency-and-climate-change.
EPA. 2021a. Air pollution: Current and future challenges. Retrieved August 16, 2021, from
https://www.epa.gov/clean-air-act-overview/air-pollution-current-and-future-challenges.
EPA. 2021b. The process of unconventional natural gas production. Retrieved September 11, 2021,
2021, from https://www.epa.gov/uog/process-unconventional-natural-gas-production.
Eris, O., D. Chachra, H. L. Chen, S. Sheppard, L. Ludlow, C. Rosca, T. Bailey, and G. Toye. 2010.
Outcomes of a longitudinal administration of the persistence in engineering survey.
Journal of Engineering Education 99(4):371-395. DOI: 10.1002/j.2168-9830.2010.
tb01069.x.
Eun, C., and M. L. Berkowitz. 2011. Molecular dynamics simulation study of interaction between
model rough hydrophobic surfaces. The Journal of Physical Chemistry A 115(23):6059-
6067. DOI: 10.1021/jp110608p.
European Bioplastics. 2021. Bioplastic materials. Retrieved August 17, 2021, from
https://www.european-bioplastics.org/bioplastics/materials.
Extance, A. 2020. IBM seeks to simplify robotic chemistry. Chemistry World. From https://www.
chemistryworld.com/news/ibm-seeks-to-simplify-robotic-chemistry/4012359.article.
ExxonMobil. 2019. 2019 Outlook for energy: A perspective to 2040. Retrieved August 17, 2021,
from https://corporate.exxonmobil.com/-/media/Global/Files/outlook-for-energy/2019-
Outlook-for-Energy_v4.pdf.
Faller, C., M. Bracher, N. Dami, and R. Roguet. 2002. Predictive ability of reconstructed human
epidermis equivalents for the assessment of skin irritation of cosmetics. Toxicology in
Vitro 16(5):557-572. DOI: 10.1016/s0887-2333(02)00053-x.
FAO (Food and Agriculture Organization of the United Nations). 2020. Land use in agriculture by
the numbers. Retrieved August 16, 2021, from http://www.fao.org/sustainability/news/
detail/en/c/1274219.
FAO. 2021. Key facts and findings. Retrieved August 16, 2021, from http://www.fao.org/news/
story/en/item/197623/icode.
Farrington, R. 2020. These states offer tuition-free community college. Forbes. https://www.
forbes.com/sites/robertfarrington/2020/03/25/these-states-offer-tuition-free-community-
college.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

References 277

FDA (U.S. Food and Drug Administration). 2021. Approved cellular and gene therapy products.
Retrieved August 16, 2021, from https://www.fda.gov/vaccines-blood-biologics/cellular-
gene-therapy-products/approved-cellular-and-gene-therapy-products.
Fenton, O. S., K. N. Olafson, P. S. Pillai, M. J. Mitchell, and R. Langer. 2018. Advances in
biomaterials for drug delivery. Advanced Materials 30(29). DOI: 10.1002/adma.201
705328.
Finley, J. W., and J. N. Seiber. 2014. The nexus of food, energy, and water. Journal of Agricultural
and Food Chemistry 62(27):6255-6262. DOI: 10.1021/jf501496r.
Flaga, C. T. 2006. The process of transition for community college transfer students. Community
College Journal of Research and Practice 30(1):3-19. DOI: 10.1080/106689205
00248845.
Flavell-While, C. 2011. Arthur D Little—Dedicated to industrial progress. The Chemical Engineer.
https://www.thechemicalengineer.com/features/cewctw-arthur-d-little-dedicated-to-indu
strial-progress.
Florey, H. W. 1949. Antibiotics: A survey of penicillin, streptomycin, and other antimicrobial
substances from fungi, actinomycetes, bacteria, and plants. New York: Oxford University
Press.
Ford, M. A. 1982. Computer control of equipment and data handling. Philosophical Transactions
of the Royal Society of London. Series A, Mathematical and Physical Sciences
307(1500):491-501. DOI: https://www.jstor.org/stable/37280.
Fortune. 1985. A database of 50 years of FORTUNE’s list of America’s largest corporations.
Retrieved August 16, 2021, from https://archive.fortune.com/magazines/fortune/fortune
500_archive/full/1985/.
Fortune. 2020. Fortune 500. Retrieved August 17, 2021, from https://fortune.com/fortune500/
2020.
Franklin Associates. 2018. Life cycle impacts for postconsumer recycled resins: PET, HDPE, and
PP. Retrieved August 17, 2021, from https://plasticsrecycling.org/images/library/2018-
APR-LCI-report.pdf.
Franks, R. G. E. 1972. Modeling and simulation in chemical engineering. Hoboken, New Jersey:
Wiley-Interscience.
Frenkel, D., and A. J. C. Ladd. 1984. New Monte Carlo method to compute the free energy of
arbitrary solids. Application to the FCC and HCP phases of hard spheres. The Journal of
Chemical Physics 81(7):3188-3193. DOI: 10.1063/1.448024.
Frueh, S. 2020. Engineering a response to the COVID-19 pandemic. National Academies of Sci-
ences, Engineering, and Medicine. https://www.nationalacademies.org/news/2020/09/
engineering-a-response-to-the-covid-19-pandemic.
Funk, C., and K. Parker. 2018. Women in STEM see more gender disparities at work, especially
those in computer jobs, majority-male workplaces. In women and men in STEM often at
odds over workplace equity. Washington, DC: Pew Research Center.
Furter, W., F., ed. 1983. History of chemical engineering: Advances in chemistry, series 190.
Washington, DC: American Chemical Society.
Fuss, S., W. F. Lamb, M. W. Callaghan, J. Hilaire, F. Creutzig, T. Amann, T. Beringer, W. De
Oliveira Garcia, J. Hartmann, T. Khanna, G. Luderer, G. F. Nemet, J. Rogelj, P. Smith, J.
V. Vicente, J. Wilcox, M. Del Mar Zamora Dominguez, and J. C. Minx. 2018. Negative
emissions—Part 2: Costs, potentials and side effects. Environmental Research Letters
13(6). DOI: 10.1088/1748-9326/aabf9f.
Gagliano, E., M. Sgroi, P. P. Falciglia, F. G. A. Vagliasindi, and P. Roccaro. 2020. Removal of
poly- and perfluoroalkyl substances (PFAS) from water by adsorption: Role of PFAS

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

278 New Directions for Chemical Engineering

chain length, effect of organic matter and challenges in adsorbent regeneration. Water
Research 171. DOI: 10.1016/j.watres.2019.115381.
Galán, G., M. Martín, and I. E. Grossmann. 2021. Integrated renewable production of sorbitol and
xylitol from switchgrass. Industrial & Engineering Chemistry Research 60(15):5558-
5573. DOI: 10.1021/acs.iecr.1c00397.
Gallo, P., K. Amann-Winkel, C. A. Angell, M. A. Anisimov, F. Caupin, C. Chakravarty, E.
Lascaris, T. Loerting, A. Z. Panagiotopoulos, J. Russo, J. A. Sellberg, H. E. Stanley, H.
Tanaka, C. Vega, L. Xu, and L. G. Pettersson. 2016. Water: A tale of two liquids.
Chemical Reviews 116(13):7463-7500. DOI: 10.1021/acs.chemrev.5b00750.
Galloway, J. N., A. R. Townsend, J. W. Erisman, M. Bekunda, Z. Cai, J. R. Freney, L. A.
Martinelli, S. P. Seitzinger, and M. A. Sutton. 2008. Transformation of the nitrogen cycle:
Recent trends, questions, and potential solutions. Science 320(5878):889-892. DOI:
10.1126/science.1136674.
Gao, Z., S. Rohani, J. Gong, and J. Wang. 2017. Recent developments in the crystallization
process: Toward the pharmaceutical industry. Engineering 3(3):343-353. DOI:
10.1016/J.ENG.2017.03.022.
Gao, W., H. Ota, D. Kiriya, K. Takei, and A. Javey. 2019. Flexible electronics toward wearable
sensing. Accounts of Chemical Research 52(3):523-533. DOI: 10.1021/acs.accounts.
8b00500.
Garcia, D. J., and F. You. 2016. The water-energy-food nexus and process systems engineering: A
new focus. Computers & Chemical Engineering 91:49-67. DOI: 10.1016/j.comp
chemeng.2016.03.003.
Gargini, P., F. Balestra, and Y. Hayashi. 2020. The international roadmap for devices and systems
(IRDS). IEEE Electron Devices Society Newsletter 27(3).
Garnier, G. 2014. Grand challenges in chemical engineering. Frontiers in Chemistry 2, Article 17.
DOI: 10.3389/fchem.2014.00017.
Gartner. 2020. Gartner says worldwide PC shipments grew 2.3% in 4Q19 and 0.6% for the year.
Gartner Newsroom. Press release https://www.gartner.com/en/newsroom/press-
releases/2020-01-13-gartner-says-worldwide-pc-shipments-grew-2-point-3-percent-in-
4q19-and-point-6-per
cent-for-the-year.
Gaynes, R. 2017. The discovery of penicillin—New insights after more than 75 years of clinical
use. Emerging Infectious Diseases 23(5):849-853. DOI: 10.3201/eid2305.161556.
Geoscience News and Information. 2021. Hydraulic fracturing fluids—Composition and additives.
Retrieved August 17, 2021, from https://geology.com/energy/hydraulic-fracturing-fluids.
Gerber, C., R. Vaikmäe, W. Aeschbach, A. Babre, W. Jiang, M. Leuenberger, Z.-T. Lu, R. Mokrik,
P. Müller, V. Raidla, T. Saks, H. N. Waber, T. Weissbach, J. C. Zappala, and R.
Purtschert. 2017. Using 81Kr and noble gases to characterize and date groundwater and
brines in the Baltic Artesian Basin on the one-million-year timescale. Geochimica et
Cosmochimica Acta 205:187-210. DOI: 10.1016/j.gca.2017.01.033.
Geyer, R. 2021. Plastic: Too much of a good thing? Presented at Wallace Stegner Center 26th
Annual Symposium—The Plastics Paradox: Societal Boon or Environmental Bane?. Salt
Lake City, UT.
Geyer, R., J. R. Jambeck, and K. L. Law. 2017. Production, use, and fate of all plastics ever made.
Science Advances 3(7). DOI: 10.1126/sciadv.1700782.
Ghaemmaghami, A. M., M. J. Hancock, H. Harrington, H. Kaji, and A. Khademhosseini. 2012.
Biomimetic tissues on a chip for drug discovery. Drug Discovery Today 17:173-181.
DOI: 10.1016/j.drudis.2011.10.029.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

References 279

Ghosal, S. 2005. Electron spectroscopy of aqueous solution interfaces reveals surface enhancement
of halides. Science 307(5709):563-566. DOI: 10.1126/science.1106525.
Glasser, W. G. 2019. About making lignin great again—Some lessons from the past. Frontiers in
Chemistry 7, Article 565. DOI: 10.3389/fchem.2019.00565.
Glotzer, S. C., and M. J. Solomon. 2007. Anisotropy of building blocks and their assembly into
complex structures. Nature Materials 6(8):557-562. DOI: 10.1038/nmat1949.
Godlewski, J. E., J. Villalobos, D. I. Collias, J. E. and Velasquez. 2014. Process for production of
acrylic acid or its derivatives from hydroxypropionic acid or its derivatives. U.S. Patent
No. 8,884,050. Washington, DC: U.S. Patent and Trademark Office.
Gong, K., Y. Cheng, L. L. Daemen, and C. E. White. 2019. In situ quasi-elastic neutron scattering
study on the water dynamics and reaction mechanisms in alkali-activated slags. Physical
Chemistry Chemical Physics 21(20):10277-10292. DOI: 10.1039/c9cp00889f.
Gonzalez-Portillo, L. F., K. Albrecht, and C. K. Ho. 2021. Techno-economic optimization of CSP
plants with free-falling particle receivers. Entropy 23(1). DOI: 10.3390/e23010076.
Goto, Y., T. Hisatomi, Q. Wang, T. Higashi, K. Ishikiriyama, T. Maeda, Y. Sakata, S. Okunaka,
H. Tokudome, M. Katayama, S. Akiyama, H. Nishiyama, Y. Inoue, T. Takewaki, T.
Setoyama, T. Minegishi, T. Takata, T. Yamada, and K. Domen. 2018. A particulate
photocatalyst water-splitting panel for large-scale solar hydrogen generation. Joule
2(3):509-520. DOI: 10.1016/j.joule.2017.12.009.
Gove, J. M., J. L. Whitney, M. A. McManus, J. Lecky, F. C. Carvalho, J. M. Lynch, J. Li, P.
Neubauer, K. A. Smith, J. E. Phipps, D. R. Kobayashi, K. B. Balagso, E. A. Contreras,
M. E. Manuel, M. A. Merrifield, J. J. Polovina, G. P. Asner, J. A. Maynard, and G. J.
Williams. 2019. Prey-size plastics are invading larval fish nurseries. Proceedings of the
National Academy of Sciences 116(48):24143-24149. DOI: 10.1073/pnas.1907496116.
Goyal, H., A. Mehdad, R. F. Lobo, G. D. Stefanidis, and D. G. Vlachos. 2019. Scaleup of a single-
mode microwave reactor. Industrial & Engineering Chemistry Research 59(6):2516-
2523. DOI: 10.1021/acs.iecr.9b04491.
Grand View Research. 2021. Artificial intelligence market size, share & trends analysis report by
solution, by technology (deep learning, machine learning, natural language processing,
machine vision), by end use, by region, and segment forecasts, 2021-2028. Retrieved
2021, from https://www.researchandmarkets.com/reports/4375395/global-artificial-intel
ligence-market-size-share.
Grätzel, M., and J. Milić. 2019. The advent of molecular photovoltaics and hybrid perovskite solar
cells. Substantia 3(2):27-43. DOI: 10.13128/Substantia-697.
Green, D. W., and M. Z. Southard, eds. 2019. Perry’s chemical engineers’ handbook. New York:
McGraw Hill Education.
Greenstein, K. E., N. V. Myung, G. F. Parkin, and D. M. Cwiertny. 2019. Performance comparison
of hematite (α-Fe2O3)-polymer composite and core-shell nanofibers as point-of-use fil-
tration platforms for metal sequestration. Water Research 148:492-503. DOI: 10.1016/
j.watres.2018.10.048.
Gumpertz, M., R. Durodoye, E. Griffith, and A. Wilson. 2017. Retention and promotion of women
and underrepresented minority faculty in science and engineering at four large land grant
institutions. PLoS One 12(11). DOI: 10.1371/journal.pone.0187285.
Gupta, A. 2018. Introduction to deep learning: Part 1. CEP Magazine. From https://www.aiche.
org/resources/publications/cep/2018/june/introduction-deep-learning-part-1.
Gutowski, T. G., S. Sahni, J. M. Allwood, M. F. Ashby, and E. Worrell. 2013. The energy required
to produce materials: Constraints on energy-intensity improvements, parameters of
demand. Philosophical Transactions of the Royal Society A 371(1986):20120003. DOI:
10.1098/rsta.2012.0003.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

280 New Directions for Chemical Engineering

Haegel, N. M., H. Atwater, Jr., T. Barnes, C. Breyer, A. Burrell, Y. M. Chiang, S. De Wolf, B.


Dimmler, D. Feldman, S. Glunz, J. C. Goldschmidt, D. Hochschild, R. Inzunza, I.
Kaizuka, B. Kroposki, S. Kurtz, S. Leu, R. Margolis, K. Matsubara, A. Metz, W. K.
Metzger, M. Morjaria, S. Niki, S. Nowak, I. M. Peters, S. Philipps, T. Reindl, A. Richter,
D. Rose, K. Sakurai, R. Schlatmann, M. Shikano, W. Sinke, R. Sinton, B. J. Stanbery, M.
Topic, W. Tumas, Y. Ueda, J. van de Lagemaat, P. Verlinden, M. Vetter, E. Warren, M.
Werner, M. Yamaguchi, and A. W. Bett. 2019. Terawatt-scale photovoltaics: Transform
global energy. Science 364(6443):836-838. DOI: 10.1126/science.aaw1845.
Hafner, J., C. Wolverton, and G. Ceder. 2006. Toward computational materials design: The impact
of density functional theory on materials research. MRS Bulletin 31(9):659-668. DOI:
10.1557/mrs2006.174.
Hamilton, S. R., R. C. Davidson, N. Sethuraman, J. H. Nett, Y. Jiang, S. Rios, P. Bobrowicz, T. A.
Stadheim, H. Li, B.-K. Choi, D. Hopkins, H. Wischnewski, J. Roser, T. Mitchell, R. R.
Strawbridge, J. Hoopes, S. Wildt, and T. U. Gerngross. 2006. Humanization of yeast to
produce complex terminally sialylated glycoproteins. Science 313(5792):1441-1443.
DOI: 10.1126/science.1130256.
Han, K., C. W. Shields Iv, and O. D. Velev. 2018. Engineering of self-propelling microbots and
microdevices powered by magnetic and electric fields. Advanced Functional Materials
28(25). DOI: 10.1002/adfm.201705953.
Handle, P. H., T. Loerting, and F. Sciortino. 2017. Supercooled and glassy water: Metastable
liquid(s), amorphous solid(s), and a no-man’s land. Proceedings of the National Academy
of Sciences 114(51):13336-13344. DOI: 10.1073/pnas.1700103114.
Hanson, M. 2021. Average cost of college & tuition. Retreived 2021, from https://education
data.org/average-cost-of-college.
Hao, Y., W. Li, X. Zhou, F. Yang, and Z. Qian. 2017. Microneedles-based transdermal drug
delivery systems: A review. Journal of Biomedical Nanotechnology 13(12):1581-1597.
DOI: 10.1166/jbn.2017.2474.
Hardin, B. E., H. J. Snaith, and M. D. McGehee. 2012. The renaissance of dye-sensitized solar
cells. Nature Photonics 6(3):162-169. DOI: 10.1038/nphoton.2012.22.
Harimoto, T., and T. Danino. 2019. Engineering bacteria for cancer therapy. Emerging Topics in
Life Sciences 3(5):623-629. DOI: 10.1042/etls20190096.
Harmsen, G. J. 2007. Reactive distillation: The front-runner of industrial process intensification.
Chemical Engineering and Processing: Process Intensification 46(9):774-780. DOI:
10.1016/j.cep.2007.06.005.
Hart, J., R. M. Machado, H. Withers Jr., S. Lo, E. Cialkowski, K. Jambunathan. 2015. Electrolytic
apparatus, system and method for the safe production of nitrogen trifluoride. U.S. Patent
No. 8945367. Washington, DC: U.S. Patent and Trademark Office.
Hasanbeigi, A., and C. Springer. 2019. Deep decarbonization roadmap for the cement and con-
crete industries in California. San Fransisco, CA: Global Efficiency Intelligence.
https://www.climateworks.org/wp-content/uploads/2019/09/Decarbonization-Roadmap-
CA-Cement-Final.pdf.
Häuβler, M., M. Eck, D. Rothauer, and S. Mecking. 2021. Closed-loop recycling of polyethylene-
like materials. Nature 590(7846):423-427. DOI: 10.1038/s41586-020-03149-9.
Hautier, G., A. Jain, and S. P. Ong. 2012. From the computer to the laboratory: Materials discovery
and design using first-principles calculations. Journal of Materials Science 47(21):7317-
7340. DOI: 10.1007/s10853-012-6424-0.
Haywood, J. 2016. Chapter 27—Atmospheric aerosols and their role in climate change. In Climate
Change (Second Edition). T. M. Letcher, ed. Boston: Elsevier.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

References 281

Heinken, A., G. Acharya, D. A. Ravcheev, J. Hertel, M. Nyga, O. E. Okpala, M. Hogan, S.


Magnúsdóttir, F. Martinelli, G. Preciat, J. N. Edirisinghe, C. S. Henry, R. M. T. Fleming,
and I. Thiele. 2020. AGORA2: Large scale reconstruction of the microbiome highlights
wide-spread drug-metabolising capacities. bioRxiv. DOI: 10.1101/2020.11.09.375451.
Helgeson, M. E. 2016. Colloidal behavior of nanoemulsions: Interactions, structure, and rheology.
Current Opinion in Colloid & Interface Science 25:39-50. DOI: 10.1016/j.cocis.
2016.06.006.
Heney, P. 2020. Global R&D investments unabated in spending growth. Global funding forecast.
Cleveland, OH: R&D World. From https://www.rdworldonline.com/global-rd-invest-
ments-unabated-in-spending-growth.
Hepburn, C., E. Adlen, J. Beddington, E. A. Carter, S. Fuss, N. Mac Dowell, J. C. Minx, P. Smith,
and C. K. Williams. 2019. The technological and economic prospects for CO2 utilization
and removal. Nature 575(7781):87-97. DOI: 10.1038/s41586-019-1681-6.
Higman, C., and M. van der Burgt. 2008. Gasification. Oxford, UK: Gulf Professional Publishing.
Himmelblau, D. M., and K. B. Bischoff. 1968. Process analysis and simulation: Deterministic
systems. New York: John Wiley & Sons, Inc.
Hoit, M., and M. Ohland. 1998. The impact of a discipline-based introduction to engineering
course on improving retention. Journal of Engineering Education 87(1):79-85. DOI:
10.1002/j.2168-9830.1998.tb00325.x.
Holt, J. K., H. G. Park, Y. Wang, M. Stadermann, A. B. Artyukhin, C. P. Grigoropoulos, A. Noy,
and O. Bakajin. 2006. Fast mass transport through sub-2-nanometer carbon nanotubes.
Science 312(5776):1034-1037. DOI: 10.1126/science.1126298.
Hoornweg, D., P. Bhada-Tata, and C. Kennedy. 2013. Environment: Waste production must peak
this century. Nature 502(7473):615-617. DOI: 10.1038/502615a.
Hori, Y. 2008. Electrochemical CO2 Reduction on metal electrodes. In Modern aspects of electro-
chemistry. C. G., Vayenas, R. E. White, and M. E. Gamboa-Aldeco, eds. New York:
Springer.
Hougen, O. A., and K. M. Watson. 1943. Chemical process principles. Part I: Material and energy
balances. New York: John Wiley and Sons, Inc.
Hougen, O. A., and K. M. Watson. 1947a. Chemical process principles. Part III: Kinetics and
catalysis. New York: John Wiley and Sons, Inc.
Hougen, O. A., and K. M. Watson. 1947b. Chemical process principles. Part II: Thermodynamics.
New York: John Wiley and Sons, Inc.
Hsieh, C.-T., M.-J. Lee, and H.-m. Lin. 2006. Multiphase equilibria for mixtures containing acetic
acid, water, propylene glycol monomethyl ether, and propylene glycol methyl ether
acetate. Industrial & Engineering Chemistry Research 45(6):2123-2130. DOI: 10.1021/
ie051245t.
Hua, L., R. Zangi, and B. J. Berne. 2009. Hydrophobic interactions and dewetting between plates
with hydrophobic and hydrophilic domains. The Journal of Physical Chemistry C
113(13):5244-5253. DOI: 10.1021/jp8088758.
Hubbs, C. D., C. Li, N. V. Sahinidis, I. E. Grossmann, and J. M. Wassick. 2020. A deep
reinforcement learning approach for chemical production scheduling. Computers &
Chemical Engineering 141. DOI: 10.1016/j.compchemeng.2020.106982.
Hussain, A., Y. D. Chaniago, A. Riaz, and M. Lee. 2019. Process design alternatives for producing
ultra-high-purity electronic-grade propylene glycol monomethyl ether acetate. Industrial
& Engineering Chemistry Research 58(6):2246-2257. DOI: 10.1021/acs.iecr.8b04052.
Hwang, I. Y., H. L. Lee, J. G. Huang, Y. Y. Lim, W. S. Yew, Y. S. Lee, and M. W. Chang. 2018.
Engineering microbes for targeted strikes against human pathogens. Cellular and
Molecular Life Sciences 75(15):2719-2733. DOI: 10.1007/s00018-018-2827-7.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

282 New Directions for Chemical Engineering

Hydrogen Council. 2019. Hydrogen decarbonization pathways. Retrieved August 17, 2021, from
https://hydrogencouncil.com/en/hydrogen-decarbonization-pathways.
IChemE (Institution of Chemical Engineers). 2015. Ten ways chemical engineers can save the
world from climate change #COP21. Retrieved January 18, 2021, from
https://ichemeblog.org/2015/12/21/ten-ways-chemical-engineers-can-save-the-world-from
-climate-change-cop21.
IEA (International Energy Agency). 2012. Technology roadmap—High-efficiency, low-emissions
coal-fired power generation. Retrieved August 18, 2021, from https://www.iea.org/
reports/technology-roadmap-high-efficiency-low-emissions-coal-fired-power-generation.
IEA. 2019. The future of hydrogen. Retrieved August 16, 2021, from https://www.iea.org/
reports/the-future-of-hydrogen.
IEA. 2020a. Global EV outlook 2020. Retrieved August 16, 2021, from https://www.iea.org/
reports/global-ev-outlook-2020.
IEA. 2020b. Bio-based chemicals—A 2020 update. Retrieved August 16, 2021, from
https://www.ieabioenergy.com/blog/publications/new-publication-bio-based-chemicals-
a-2020-update.
IEA. 2020c. Coal-fired power. https://prod.iea.org/reports/coal-fired-power.
IEA. 2020d. Energy technology perspectives 2020. Retrieved August 16, 2021, from
https://www.iea.org/reports/energy-technology-perspectives-2020.
IEA. 2020e. Global hydrogen demand by sector in the Sustainable Development Scenario, 2019-
2070. Retrieved August 16, 2021, from https://www.iea.org/data-and-statistics/charts/
global-hydrogen-demand-by-sector-in-the-sustainable-development-scenario-2019-2070.
IEA. 2021a. Transport. Retrieved August 16, 2021, from https://www.iea.org/topics/transport.
IEA. 2021b. Net zero by 2050: A roadmap for the global energy sector. Retrieved August 16, 2021,
from https://www.iea.org/reports/net-zero-by-2050.
IEA. 2021c. Coal Information: Overview. Retrieved August 9, 2021, from https://www.iea.org/
reports/coal-information-overview.
IEA. 2021d. Natural gas information: Overview. Retrieved August 9, 2021, from https://www.iea.
org/reports/natural-gas-information-overview.
IEA. 2021e. World energy balances: Overview. Retrieved August 16, 2021, from https://www.iea.
org/reports/world-energy-balances-overview.
IEA, ICCA, and DECHEMA (International Council of Chemical Associations, and German Soci-
ety for Chemical Engineering and Biotechnology). 2013. Technology roadmap: Energy
and GHG Reductions in the chemical industry via catalytic processes.
https://iea.blob.core.windows.net/assets/d0f7ff3a-0612-422d-ad7d-a682091cb500/Tech
nologyRoadmapEnergyandGHGReductionsintheChemicalIndustryviaCatalyticProcesses
.pdf.
IndustryARC. 2021. Protein therapeutics market—Forecast (2021–2026). Retrieved August 17,
2021, from https://www.industryarc.com/Report/16207/protein-therapeutics-market.html.
IRENA (International Renewable Energy Agency). 2018. Hydrogen from renewable power
technology outlook for the energy transition. Retrieved August 18, 2021, from
https://www.irena.org/publications/2018/sep/hydrogen-from-renewable-power.
IRENA. 2020. Innovation outlook: Ocean energy technologies. Retrieved August 18, 2021, from
https://www.irena.org/publications/2020/Dec/Innovation-Outlook-Ocean-Energy-Techn
ologies.
IRENA. 2021. World energy transitions outlook: 1.5°C pathway. Retrieved August 16, 2021, from
https://irena.org/publications/2021/Jun/World-Energy-Transitions-Outlook.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

References 283

Irwin, D. A., and P. J. Klenow. 1996. High-tech R&D subsidies estimating the effects of Sematech.
Journal of International Economics 40(3):323-344. DOI: 10.1016/0022-1996(95)01408-
X.
ISO (International Standards Organization). 2006. Environmental management—Life cycle
assessment—Principles and framework. https://www.iso.org/standard/37456.html.
Iulianelli, A., S. Liguori, J. Wilcox, and A. Basile. 2016. Advances on methane steam reforming
to produce hydrogen through membrane reactors technology: A review. Catalysis
Reviews 58(1):1-35. DOI: 10.1080/01614940.2015.1099882.
Iyare, P. U., S. K. Ouki, and T. Bond. 2020. Microplastics removal in wastewater treatment plants:
A critical review. Environmental Science: Water Research & Technology 6(10):2664-
2675. DOI: 10.1039/d0ew00397b.
Jagschies, G. 2020. Hierarchy of high impact improvements in biomanufacturing. Presented at
Workshop on Innovations in Pharmaceutical Manufacturing, National Academies of
Sciences, Engineering, and Medicine. Washington, DC.
Jain, A., S. P. Ong, G. Hautier, W. Chen, W. D. Richards, S. Dacek, S. Cholia, D. Gunter, D.
Skinner, G. Ceder, and K. A. Persson. 2013. The materials project: A materials genome
approach to accelerating materials innovation. APL Materials 1. DOI: 10.1063/1.481
2323.
Jama-Rodzenska, A., A. Bialowiec, J. A. Koziel, and J. Sowinski. 2021. Waste to phosphorus: A
transdisciplinary solution to P recovery from wastewater based on the TRIZ approach.
Journal of Environmental Economics and Management 287. DOI: 10.1016/j.
jenvman.2021.112235.
Jambeck, J. R., R. Geyer, C. Wilcox, T. R. Siegler, M. Perryman, A. Andrady, R. Narayan, and K.
L. Law. 2015. Plastic waste inputs from land into the ocean. Science 347(6223):768-771.
DOI: 10.1126/science.1260352.
Janes, K. A., P. L. Chandran, R. M. Ford, M. J. Lazzara, J. A. Papin, S. M. Peirce, J. J. Saucerman,
and D. A. Lauffenburger. 2017. An engineering design approach to systems biology.
Integrative Biology 9(7):574-583. DOI: 10.1039/c7ib00014f.
Jayapal, K. P., K. F. Wlaschin, W. S. Hu, and M. G. S. Yap. 2007. Recombinant protein
therapeutics from CHO cells—20 years and counting. Chemical Engineering Progress
103(10):40-47.
Jiang, H., A. A. Horwitz, C. Wright, A. Tai, E. A. Znameroski, Y. Tsegaye, H. Warbington, B. S.
Bower, C. Alves, C. Co, K. Jonnalagadda, D. Platt, J. M. Walter, V. Natarajan, J. A.
Ubersax, J. R. Cherry, and J. C. Love. 2019. Challenging the workhorse: Comparative
analysis of eukaryotic micro-organisms for expressing monoclonal antibodies.
Biotechnology and Bioengineering 116(6):1449-1462. DOI: 10.1002/bit.26951.
Johnson, E. F. 1967. Automatic process control. New York: McGraw Hill.
Jones, P. T., P. H. Dear, J. Foote, M. S. Neuberger, and G. Winter. 1986. Replacing the
complementarity-determining regions in a human antibody with those from a mouse.
Nature 321(6069):522-525. DOI: 10.1038/321522a0.
Jungwirth, P., and D. J. Tobias. 2006. Specific ion effects at the air/water interface. Chemical
Reviews 106(4):1259-1281. DOI: 10.1021/cr0403741.
Jyothi, R. K., T. Thenepalli, J. W. Ahn, P. K. Parhi, K. W. Chung, and J.-Y. Lee. 2020. Review of
rare earth elements recovery from secondary resources for clean energy technologies:
Grand opportunities to create wealth from waste. Journal of Cleaner Production 267.
DOI: 10.1016/j.jclepro.2020.122048.
Kamm, R. D., R. Bashir, N. Arora, R. D. Dar, M. U. Gillette, L. G. Griffith, M. L. Kemp, K.
Kinlaw, M. Levin, A. C. Martin, T. C. McDevitt, R. M. Nerem, M. J. Powers, T. A. Saif,
J. Sharpe, S. Takayama, S. Takeuchi, R. Weiss, K. Ye, H. G. Yevick, and M. H. Zaman.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

284 New Directions for Chemical Engineering

2018. Perspective: The promise of multi-cellular engineered living systems. APL


Bioengineering 2(4). DOI: 10.1063/1.5038337.
Kang, F., D. Wang, Y. Pu, X.-F. Zeng, J.-X. Wang, and J.-F. Chen. 2018. Efficient preparation of
monodisperse CaCO3 nanoparticles as overbased nanodetergents in a high-gravity
rotating packed bed reactor. Powder Technology 325:405-411. DOI: 10.1016/j.powtec.
2017.11.036.
Kantor, E. D., C. D. Rehm, J. S. Haas, A. T. Chan, and E. L. Giovannucci. 2015. Trends in
prescription drug use among adults in the United States From 1999-2012. The Journal of
the American Medical Association 314(17):1818-1830. DOI: 10.1001/jama.2015.13766.
Kapoor, R. 2012. Collaborative innovation in the global semiconductor industry: A report on the
findings from the 2010 Wharton-GSA semiconductor ecosystem survey. https://faculty.
wharton.upenn.edu/wp-
content/uploads/2012/05/SemiconductorEcosystemStudyFinal.pdf.
Kapteijn, F., and J. A. Moulijn. 2020. Structured catalysts and reactors—Perspectives for
demanding applications. Catalysis Today 383:5-14. DOI: 10.1016/j.cattod.2020.09.026.
Karabasz, A., M. Bzowska, and K. Szczepanowicz. 2020. Biomedical applications of
multifunctional polymeric nanocarriers: A review of current literature. International
Journal of Nanomedicine 15:8673-8696. DOI: 10.2147/IJN.S231477.
Karsch-Mizrachi, I., Y. Nakamura, and G. Cochrane. 2012. The international nucleotide sequence
database collaboration. Nucleic Acids Research 40(D1):D33-37. DOI: 10.1093/nar/
gkr1006.
Kass, I., C. F. Reboul, and A. M. Buckle. 2011. Chapter 14—Computational methods for studying
serpin conformational change and structural plasticity. In Methods in enzymology. J.C.
Whisstock and P. I. Bird, eds. San Diego, CA: Academic Press.
Kastner, M. 2018. Philanthropy: A critical player in supporting scientific research. Science
Philanthropy Alliance News. https://sciencephilanthropyalliance.org/philanthropy-a-
critical-player-in-supporting-scientific-research-alliance-blog.
Katz, D. L. 1966. Computers in engineering design education. Ann Arbor, MI: University of
Michigan.
Kaya, M. 2016. Recovery of metals and nonmetals from electronic waste by physical and chemical
recycling processes. Waste Management 57:64-90. DOI: 10.1016/j.wasman.2016.08.004.
Kaz, D. M., R. McGorty, M. Mani, M. P. Brenner, and V. N. Manoharan. 2012. Physical ageing
of the contact line on colloidal particles at liquid interfaces. Nature Materials 11(2):138-
142. DOI: 10.1038/nmat3190.
Kaza, S., L. Yao, P. Bhada-Tata, and F. Van Woerden. 2018. What a waste 2.0—A global snapshot
of solid waste management to 2050. Washington, DC: The World Bank.
Kekicheff, P., J. Iss, P. Fontaine, and A. Johner. 2018. Direct measurement of lateral correlations
under controlled nanoconfinement. Physical Review Letters 120(11). DOI:
10.1103/PhysRevLett.120.118001.
Kellett, P. J., and D. I. Collias. 2016. Catalysts and processes for the production of aromatic
compounds from lignin. U.S. Patent No. 9452422. Washington, DC: U.S. Patent and
Trademark Office.
Kevrekidis, I. G., and G. Samaey. 2009. Equation-free multiscale computation: Algorithms and
applications. Annual Review of Physical Chemistry 60(1):321-344. DOI: 10.1146/
annurev.physchem.59.032607.093610.
Khalilpour, R., and I. A. Karimi. 2012. Evaluation of utilization alternatives for stranded natural
gas. Energy 40(1):317-328. DOI: 10.1016/j.energy.2012.01.068.
Kim, A. J., P. L. Biancaniello, and J. C. Crocker. 2006. Engineering DNA-mediated colloidal
crystallization. Langmuir 22(5):1991-2001. DOI: 10.1021/la0528955.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

References 285

Kim, S. B., J. C. Palmer, and P. G. Debenedetti. 2015. A computational study of the effect of matrix
structural order on water sorption by TRP-cage miniproteins. The Journal of Physical
Chemistry B 119(5):1847-1856. DOI: 10.1021/jp510172w.
Kim, T. H., Y. Wang, C. R. Oliver, D. H. Thamm, L. Cooling, C. Paoletti, K. J. Smith, S. Nagrath,
and D. F. Hayes. 2019. A temporary indwelling intravascular aphaeretic system for in
vivo enrichment of circulating tumor cells. Nature Communications 10(1). DOI:
10.1038/s41467-019-09439-9.
Kinney, M. A., L. T. Vo, J. M. Frame, J. Barragan, A. J. Conway, S. Li, K. K. Wong, J. J. Collins,
P. Cahan, T. E. North, D. A. Lauffenburger, and G. Q. Daley. 2019. A systems biology
pipeline identifies regulatory networks for stem cell engineering. Nature Biotechnology
37(7):810-818. DOI: 10.1038/s41587-019-0159-2.
Kleiner, L. W., J. C. Wright, and Y. Wang. 2014. Evolution of implantable and insertable drug
delivery systems. Journal of Controlled Release 181:1-10. DOI: 10.1016/j.jconrel.2014.
02.006.
Knight, D., J. Sullivan, and L. Carlson. 2003. Staying in engineering: Effects of a hands on, team
based, first year projects course on student retention. Presented at American Society for
Engineering Education Annual Conference, 2003. Nashville, TN.
Kofke, D. A. 1993. Direct evaluation of phase coexistence by molecular simulation via integration
along the saturation line. The Journal of Chemical Physics 98(5):4149-4162. DOI:
10.1063/1.465023.
König, A., W. Marquardt, A. Mitsos, J. Viell, and M. Dahmen. 2020. Integrated design of
renewable fuels and their production processes: Recent advances and challenges. Current
Opinion in Chemical Engineering 27:45-50. DOI: 10.1016/j.coche.2019.11.001.
Koros, W. J., and R. P. Lively. 2012. Water and beyond: Expanding the spectrum of large-scale
energy efficient separation processes. American Institute of Chemical Engineers Journal
58(9):2624-2633. DOI: 10.1002/aic.13888.
Koros, W. J., and C. Zhang. 2017. Materials for next-generation molecularly selective synthetic
membranes. Nature Materials 16(3):289-297. DOI: 10.1038/nmat4805.
Kriegler, E., J. P. Weyant, G. J. Blanford, V. Krey, L. Clarke, J. Edmonds, A. Fawcett, G. Luderer,
K. Riahi, R. Richels, S. K. Rose, M. Tavoni, and D. P. van Vuuren. 2014. The role of
technology for achieving climate policy objectives: Overview of the EMF 27 study on
global technology and climate policy strategies. Climatic Change 123(3):353-367. DOI:
10.1007/s10584-013-0953-7.
Kromdijk, J., K. Głowacka, L. Leonelli, S. T. Gabilly, M. Iwai, K. K. Niyogi, and S. P. Long.
2016. Improving photosynthesis and crop productivity by accelerating recovery from
photoprotection. Science 354(6314):857-861. DOI: 10.1126/science.aai8878.
Krouse, S., R. M. Machado, J. Hart, and J. Nehlsen. 2016. Electrolytic apparatus, system, and
method for the efficient production of nitrogen trifluoride. U.S. Patent 9528191.
Washington, DC: U.S. Patent and Trademark Office.
Kumar, A., H. Fukuda, T. A. Hatton, and J. H. Lienhard. 2019. Lithium recovery from oil and gas
produced water: A need for a growing energy industry. ACS Energy Letters 4(6):1471-
1474. DOI: 10.1021/acsenergylett.9b00779.
Lachance, J. C., D. Matteau, J. Brodeur, C. J. Lloyd, N. Mih, Z. A. King, T. F. Knight, A. M. Feist,
J. M. Monk, B. O. Palsson, P. E. Jacques, and S. Rodrigue. 2021. Genome-scale metabolic
modeling reveals key features of a minimal gene set. Molecular Systems Biology 17(7).
DOI: 10.15252/msb.202010099.
Landsman, M. R., R. Sujanani, S. H. Brodfuehrer, C. M. Cooper, A. G. Darr, R. J. Davis, K. Kim,
S. Kum, L. K. Nalley, S. M. Nomaan, C. P. Oden, A. Paspureddi, K. K. Reimund, L. S.
Rowles 3rd, S. Yeo, D. F. Lawler, B. D. Freeman, and L. E. Katz. 2020. Water treatment:

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

286 New Directions for Chemical Engineering

Are membranes the panacea? Annual Review of Chemical and Biomolecular Engineering
11:559-585. DOI: 10.1146/annurev-chembioeng-111919-091940.
Langhoff, S. R., ed. 1995. Quantum mechanical electronic structure calculations with chemical
accuracy. Dordrecht, Netherlands: Springer.
Lapidus, L. 1962. Digital computation for chemical engineers. New York: McGraw-Hill.
Lapointe, M., J. M. Farner, L. M. Hernandez, and N. Tufenkji. 2020. Understanding and improving
microplastic removal during water treatment: Impact of coagulation and flocculation.
Environmental Science & Technology 54(14):8719-8727. DOI: 10.1021/acs.est.0c00712.
Laramy, C. R., M. N. O’Brien, and C. A. Mirkin. 2019. Crystal engineering with DNA. Nature
Reviews Materials 4(3):201-224. DOI: 10.1038/s41578-019-0087-2.
Larson, R. G. 1999. The structure and rheology of complex fluids. New York: Oxford University
Press.
Layman, J. M., D. I. Collias, H. Schonemann, and K. Williams. 2019a. Method for purifying
reclaimed polypropylene. U.S. Patent No. 10450436. Washington, DC: U.S. Patent and
Trademark Office.
Layman, J. M., D. I. Collias, H. Schonemann, and K. Williams. 2019b. Method for purifying
reclaimed polymers. U.S. Patent No. 10465058. Washington, DC:U.S. Patent and
Trademark Office.
Lazard. 2019. Lazard’s levelized cost of energy analysis—Version 13.0. Retrieved August 18,
2021, from https://www.lazard.com/media/451086/lazards-levelized-cost-of-energy-
version-130-vf.pdf.
Leader, B., Q. J. Baca, and D. E. Golan. 2008. Protein therapeutics: A summary and
pharmacological classification. Nature Reviews Drug Discovery 7(1):21-39. DOI: 10.
1038/nrd2399.
Lebreton, L., B. Slat, F. Ferrari, B. Sainte-Rose, J. Aitken, R. Marthouse, S. Hajbane, S. Cunsolo,
A. Schwarz, A. Levivier, K. Noble, P. Debeljak, H. Maral, R. Schoeneich-Argent, R.
Brambini, and J. Reisser. 2018. Evidence that the Great Pacific Garbage Patch is rapidly
accumulating plastic. Scientific Reports 8(1). DOI: 10.1038/s41598-018-22939-w.
Lenaburg, L., O. Aguirre, F. Goodchild, and J.-U. Kuhn. 2012. Expanding pathways: A summer
bridge program for community college STEM students. Community College Journal of
Research and Practice 36(3):153-168. DOI: 10.1080/10668921003609210.
Leslie, S.-J., A. Cimpian, M. Meyer, and E. Freeland. 2015. Expectations of brilliance underlie
gender distributions across academic disciplines. Science 347(6219):262-265. DOI:
10.1126/science.1261375.
Lewis, N. S. 2016. Research opportunities to advance solar energy utilization. Science 351(6271).
DOI: 10.1126/science.aad1920.
Li, Z., T. R. Klein, D. H. Kim, M. Yang, J. J. Berry, M. F. A. M. van Hest, and K. Zhu. 2018.
Scalable fabrication of perovskite solar cells. Nature Reviews Materials 3(4). DOI:
10.1038/natrevmats.2018.17.
Li, L., J. Zhong, Y. Yan, J. Zhang, J. Xu, J. S. Francisco, and X. C. Zeng. 2020a. Unraveling
nucleation pathway in methane clathrate formation. Proceedings of the National Academy
of Sciences 117(40):24701-24708. DOI: 10.1073/pnas.2011755117.
Li, Y., Y. Sun, Y. Qin, W. Zhang, L. Wang, M. Luo, H. Yang, and S. Guo. 2020b. Recent advances
on water‐splitting electrocatalysis mediated by noble‐metal‐based nanostructured
materials. Advanced Energy Materials 10(11). DOI: 10.1002/aenm.201903120.
Li, X., J. Peoples, P. Yao, and X. Ruan. 2021a. Ultrawhite BaSO4 paints and films for remarkable
daytime subambient radiative cooling. ACS Applied Materials & Interfaces 13(18):
21733-21739. DOI: 10.1021/acsami.1c02368.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

References 287

Li, C., D. L. Ramasamy, M. Sillanpää, and E. Repo. 2021b. Separation and concentration of rare
earth elements from wastewater using electrodialysis technology. Separation and Purifi-
cation Technology 254, Article 117442. DOI: 10.1016/j.seppur.2020.117442.
Liang, R., H. Xu, Y. Shen, S. Sun, J. Xu, S. Meng, Y. R. Shen, and C. Tian. 2019. Nucleation and
dissociation of methane clathrate embryo at the gas-water interface. Proceedings of the
National Academy of Sciences 116(47):23410-23415. DOI: 10.1073/pnas.1912592116.
Lin, H., S. Lee, L. Sun, M. Spellings, M. Engel, C. Glotzer Sharon, and A. Mirkin Chad. 2017.
Clathrate colloidal crystals. Science 355(6328):931-935. DOI: 10.1126/science.aal3919.
Lipani, L., B. G. R. Dupont, F. Doungmene, F. Marken, R. M. Tyrrell, R. H. Guy, and A. Ilie.
2018. Non-invasive, transdermal, path-selective and specific glucose monitoring via a
graphene-based platform. Nature Nanotechnology 13(6):504-511. DOI: 10.1038/s41565-
018-0112-4.
Litzler, E., and J. Young. 2012. Understanding the risk of attrition in undergraduate engineering:
Results from the project to assess climate in engineering. Journal of Engineering
Education 101(2):319-345. DOI: 10.1002/j.2168-9830.2012.tb00052.x.
Logar, N., L. D. Anadon, and V. Narayanamurti. 2014. Semiconductor research corporation: A
case study in cooperative innovation partnerships. Minerva 52(2):237-261. DOI:
10.1007/s11024-014-9253-2.
Low, L. A., C. Mummery, B. R. Berridge, C. P. Austin, and D. A. Tagle. 2021. Organs-on-chips:
Into the next decade. Nature Reviews Drug Discovery 20(5):345-361. DOI: 10.1038/
s41573-020-0079-3.
Lu, R.-M., Y.-C. Hwang, I. J. Liu, C.-C. Lee, H.-Z. Tsai, H.-J. Li, and H.-C. Wu. 2020.
Development of therapeutic antibodies for the treatment of diseases. Journal of
Biomedical Science 27(1). DOI: 10.1186/s12929-019-0592-z.
Luo, Y., P. R. Westmoreland, D. Alkaya, R.V. Alves da Cruz, I.E. Grossmann, W.D. Provine, D.L.
Silverstein, R.J. Steininger II, J.B. Talbot, A. Varma, T. McCreight, K. Chin, D. Schuster.
2015. Chemical engineering academia-industry alignment: Expectations about new grad-
uates. American Institute of Chemical Engineers. https://www.aiche.org/sites/de-
fault/files/docs/conferences/2015che_academicindustryalignmentstudy.compressed.pdf.
Luo, X., B. Guo, J. Luo, F. Deng, S. Zhang, S. Luo, and J. Crittenden. 2015. Recovery of lithium
from wastewater using development of li ion-imprinted polymers. ACS Sustainable
Chemistry & Engineering 3(3):460-467. DOI: 10.1021/sc500659h.
Luo, S., T. Li, X. Wang, M. Faizan, and L. Zhang. 2020. High‐throughput computational materials
screening and discovery of optoelectronic semiconductors. WIREs Computational
Molecular Science 11(1). DOI: 10.1002/wcms.1489.
Lutz, J.-F., M. Ouchi, D. R. Liu, and M. Sawamoto. 2013. Sequence-controlled polymers. Science
341(6146). DOI: 10.1126/science.1238149.
Lynd, L. R., J. H. Cushman, R. J. Nichols, and C. E. Wyman. 1991. Fuel ethanol from cellulosic
biomass. Science 251(4999):1318-1323. DOI: 10.1126/science.251.4999.1318.
Ma, C. D., C. Wang, C. Acevedo-Velez, S. H. Gellman, and N. L. Abbott. 2015. Modulation of
hydrophobic interactions by proximally immobilized ions. Nature 517(7534):347-350.
DOI: 10.1038/nature14018.
Machado, R. M., and R. R. Broekhuis. 2003. Gas-liquid reaction process including ejector and
monolith catalyst. U.S. Patent No. 6506361. Washington, DC: U.S. Patent and Trademark
Office.
Machado, R. M., R. R. Broekhuis, A. F. Nordquist, B. P. Roy, and S. R. Carney. 2005. Applying
monolith reactors for hydrogenations in the production of specialty chemicals—Process
and economic considerations. Catalysis Today 105:305-317. DOI: 10.1016/j.cattod.
2005.06.036.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

288 New Directions for Chemical Engineering

Macrotrends. 2021a. Exxon research and development expenses 2006-2021. Retrieved December
3, 2021, from https://www.macrotrends.net/stocks/charts/XOM/exxon/research-develop
ment-expenses.
Macrotrends. 2021b. Corteva research and development expenses 2018-2021. Retrieved December
3, 2021, from https://www.macrotrends.net/stocks/charts/CTVA/corteva/research-devel
opment-expenses.
Maeurer, A., M. Schlummer, and O. Beck. 2012. Method for recycling plastic materials and use
thereof. U.S. Patent No. 8138232. Washington, DC: U.S. Patent and Trademark Office.
Magnusdottir, S., A. Heinken, L. Kutt, D. A. Ravcheev, E. Bauer, A. Noronha, K. Greenhalgh, C.
Jager, J. Baginska, P. Wilmes, R. M. Fleming, and I. Thiele. 2017. Generation of genome-
scale metabolic reconstructions for 773 members of the human gut microbiota. Nature
Biotechnology 35(1):81-89. DOI: 10.1038/nbt.3703.
Mahmood, A., M. Eqan, S. Pervez, H. A. Alghamdi, A. B. Tabinda, A. Yasar, K. Brindhadevi, and
A. Pugazhendhi. 2020. COVID-19 and frequent use of hand sanitizers: Human health and
environmental hazards by exposure pathways. Science of the Total Environment 742.
DOI: 10.1016/j.scitotenv.2020.140561.
Makurvet, F. D. 2021. Biologics vs. small molecules: Drug costs and patient access. Medicine in
Drug Discovery 9. DOI: 10.1016/j.medidd.2020.100075.
Malcom, S. M. 1996. Science and diversity: A compelling national interest. Science.
271(5257):1817-1819. http://www.jstor.org/stable/2889362.
Malcom, S., and L. Malcom-Piqueux. 2020. Institutional transformation: Supporting equity and
excellence in STEMM. Change: The Magazine of Higher Learning 52(2):79-82. DOI:
10.1080/00091383.2020.1732792.
Mandal, J., Y. Fu, A. C. Overvig, M. Jia, K. Sun, N. N. Shi, H. Zhou, X. Xiao, N. Yu, and Y. Yang.
2018. Hierarchically porous polymer coatings for highly efficient passive daytime
radiative cooling. Science 362(6412):315-319. DOI: 10.1126/science.aat9513.
Marbach, S., D. S. Dean, and L. Bocquet. 2018. Transport and dispersion across wiggling
nanopores. Nature Physics 14(11):1108-1113. DOI: 10.1038/s41567-018-0239-0.
Marchetti, M. C., J. F. Joanny, S. Ramaswamy, T. B. Liverpool, J. Prost, M. Rao, and R. A. Simha.
2013. Hydrodynamics of soft active matter. Reviews of Modern Physics 85(3):1143-1189.
DOI: 10.1103/RevModPhys.85.1143.
Marshall, J. 2007. Who needs oil? New Scientist 195(2611):28-31. DOI: https://doi.org/
10.1016/S0262-4079(07)61712-6.
Martin, A. B., M. Hartman, D. Lassman, A. Catlin, and T. National Health Expenditure Accounts.
2021. National health care spending in 2019: Steady growth for the fourth consecutive
year. Health Affairs 40(1):14-24. DOI: 10.1377/hlthaff.2020.02022.
Masias, A., J. Marcicki, and W. A. Paxton. 2021. Opportunities and challenges of lithium ion
batteries in automotive applications. ACS Energy Letters 6(2):621-630. DOI:
10.1021/acsenergylett.0c02584.
Matthews, A. A., P. L. R. Ee, and R. Ge. 2020. Developing inhaled protein therapeutics for lung
diseases. Molecular Biomedicine 1. DOI: 10.1186/s43556-020-00014-z.
Matthews, C. B., C. Wright, A. Kuo, N. Colant, M. Westoby, and J. C. Love. 2017. Reexamining
opportunities for therapeutic protein production in eukaryotic microorganisms.
Biotechnology and Bioengineering 114(11):2432-2444. DOI: 10.1002/bit.26378.
Maxson, A., and J. Phillips. 2011. Research and development for future coal generation. Retrieved
August 16, 2021, from https://www.powermag.com/research-and-development-for-
future-coal-generation.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

References 289

Maziarka, Ł., A. Pocha, J. Kaczmarczyk, K. Rataj, T. Danel, and M. Warchoł. 2020. Mol-
CycleGAN: A generative model for molecular optimization. Journal of Cheminformatics
12(1). DOI: 10.1186/s13321-019-0404-1.
Mbow, C., C. Rosenzweig, L.G. Barioni, T.G. Benton, M. Herrero, M. Krishnapillai, E. Liwenga,
P. Pradhan, M.G. Rivera-Ferre,, and F. N. T. T. Sapkota, and Y. Xu. 2019. Food security.
In Climate change and land: An IPCC special report on climate change, desertification,
land degradation, sustainable land management, food security, and greenhouse gas
fluxes in terrestrial ecosystems. P.R. Shukla, J. S., E. Calvo Buendia, V. Masson-
Delmotte, H.-O. Pörtner, D.C. Roberts, P. Zhai, R. Slade, S. Connors, R. van Diemen, M.
Ferrat, E. Haughey, S. Luz, S. Neogi, M. Pathak, J. Petzold, J. Portugal Pereira, P. Vyas,
E. Huntley, K. Kissick, M. Belkacemi, J. Malley, eds. Geneva: Intergovernmental Panel
on Climate Change.
McBride, J., and A. Chatzky. 2019. Is ‘Made in China 2025’ a threat to global trade? Retrieved
September 13, 2021, from https://www.cfr.org/backgrounder/made-china-2025-threat-
global-trade.
McCarty, N. S., and R. Ledesma-Amaro. 2019. Synthetic biology tools to engineer microbial
communities for biotechnology. Trends in Biotechnology 37(2):181-197. DOI: 10.1016/
j.tibtech.2018.11.002.
McCullough, M. B. A., and K. Williams. 2018. STEM researchers are needed to advance multi-
level interventions for health disparities. Journal of Public Health Policy and Planning
2(1):71-73. https://www.alliedacademies.org/download.php?download=articles/stem-
researchers-are-needed-to-advance-multilevel-interventions-for-health-disparities.pdf.
McGann, P. T., and C. Hoppe. 2017. The pressing need for point-of-care diagnostics for sickle cell
disease: A review of current and future technologies. Blood Cells, Molecules and
Diseases 67:104-113. DOI: 10.1016/j.bcmd.2017.08.010.
McGee, E. O. 2020. Interrogating structural racism in STEM higher education. Educational
Researcher 49(9):633-644. DOI: 10.3102/0013189X20972718.
McKinsey Sustainability. 2016. The circular economy: Moving from theory to practice. Retrieved
August 18, 2021, from https://www.mckinsey.com/business-functions/sustainability/our-
insights/the-circular-economy-moving-from-theory-to-practice.
McMichael, S., P. Fernández-Ibáñez, and J. A. Byrne. 2021. A review of photoelectrocatalytic
reactors for water and wastewater treatment. Water 13(9). DOI: 10.3390/w13091198.
McMurtrie, B. 2018. This is what Georgia Tech thinks college will look like in 2040. The
Chronicle of Higher Education. https://www.chronicle.com/article/this-is-what-georgia-
tech-thinks-college-will-look-like-in-2040.
McNeill, V. F. 2020. COVID-19 and the air we breathe. ACS Earth and Space Chemistry 4(5):674-
675. DOI: 10.1021/acsearthspacechem.0c00093.
Mester, Z., and A. Z. Panagiotopoulos. 2015. Temperature-dependent solubilities and mean ionic
activity coefficients of alkali halides in water from molecular dynamics simulations. The
Journal of Chemical Physics 143(4). DOI: 10.1063/1.4926840.
Miandad, R., M. A. Barakat, A. S. Aburiazaiza, M. Rehan, and A. S. Nizami. 2016. Catalytic
pyrolysis of plastic waste: A review. Process Safety and Environmental Protection
102:822-838. DOI: 10.1016/j.psep.2016.06.022.
Middleton, C. T., P. Marek, P. Cao, C. C. Chiu, S. Singh, A. M. Woys, J. J. de Pablo, D. P. Raleigh,
and M. T. Zanni. 2012. Two-dimensional infrared spectroscopy reveals the complex
behaviour of an amyloid fibril inhibitor. Nature Chemistry 4(5):355-360. DOI:
10.1038/nchem.1293.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

290 New Directions for Chemical Engineering

Milbrandt, A., T. Seiple, D. Heimiller, R. Skaggs, and A. Coleman. 2018. Wet waste-to-energy
resources in the United States. Resources, Conservation and Recycling 137:32-47. DOI:
10.1016/j.resconrec.2018.05.023.
Miller, K. K., and H. S. Alper. 2019. Yarrowia lipolytica: More than an oleaginous workhorse.
Applied Microbiology and Biotechnology 103:9251-9262. DOI: 10.1007/s00253-019-
10200-x.
MIT (Massachusetts Institute of Technology). 2021. Department of chemical engineering—
History. Retrieved August 17, 2021, from https://cheme.mit.edu/about/history.
Mitchell, M. J., M. M. Billingsley, R. M. Haley, M. E. Wechsler, N. A. Peppas, and R. Langer.
2021. Engineering precision nanoparticles for drug delivery. Nature Reviews Drug
Discovery 20(2):101-124. DOI: 10.1038/s41573-020-0090-8.
Monroe, J., M. Barry, A. DeStefano, P. Aydogan Gokturk, S. Jiao, D. Robinson-Brown, T.
Webber, E. J. Crumlin, S. Han, and M. S. Shell. 2020. Water structure and properties at
hydrophilic and hydrophobic surfaces. Annual Review of Chemical and Biomolecular
Engineering 11:523-557. DOI: 10.1146/annurev-chembioeng-120919-114657.
Montoya, J. H., L. C. Seitz, P. Chakthranont, A. Vojvodic, T. F. Jaramillo, and J. K. Nørskov.
2017. Materials for solar fuels and chemicals. Nature Materials 16(1):70-81. DOI:
10.1038/nmat4778.
Morishita, M., and N. A. Peppas. 2006. Is the oral route possible for peptide and protein drug
delivery? Drug Discovery Today 11:905-910. DOI: 10.1016/j.drudis.2006.08.005.
Morrow, W. R., J. Marano, A. Hasanbeigi, E. Masanet, and J. Sathaye. 2015. Efficiency
improvement and CO2 emission reduction potentials in the United States petroleum
refining industry. Energy 93:95-105. DOI: 10.1016/j.energy.2015.08.097.
Moss, B., O. Babacan, A. Kafizas, and A. Hankin. 2021. A review of inorganic photoelectrode
developments and reactor scale‐up challenges for solar hydrogen production. Advanced
Energy Materials 11(13). DOI: 10.1002/aenm.202003286.
Moya, X., and N. D. Mathur. 2020. Caloric materials for cooling and heating. Science
370(6518):797-803. DOI: 10.1126/science.abb0973.
Mozur, P., and S. L. Myers. 2021. Xi’s gambit: China plans for a world without American
technology. New York Times. https://www.nytimes.com/2021/03/10/business/china-us-
tech-rivalry.html.
Mullin, R. 2021. Cell and gene therapy: The next frontier in pharmaceutical services. Chemical
and Engineering News Digital Magazine 99(14). https://cen.acs.org/business/out
sourcing/Cell-and-gene-therapy-The-next-frontier-in-pharmaceutical-services/99/i14.
Muralikrishna, I. V., and V. Manickam. 2017. Chapter 5—Life cycle assessment. In Environmental
management. Oxford: Butterworth-Heinemann.
Murphy, S. V., and A. Atala. 2014. 3D bioprinting of tissues and organs. Nature Biotechnology
32(8):773-785. DOI: 10.1038/nbt.2958.
NAE (National Academy of Engineering). 2008. Changing the conversation: Messages for
improving public understanding of engineering. Washington, DC: The National
Academies Press. DOI: 10.17226/12187.
NAE. 2018. Understanding the educational and career pathways of engineers. Washington, DC:
The National Academies Press. DOI: 10.17226/25284.
Naidu, G., S. Ryu, R. Thiruvenkatachari, Y. Choi, S. Jeong, and S. Vigneswaran. 2019. A critical
review on remediation, reuse, and resource recovery from acid mine drainage.
Environmental Pollution 247:1110-1124. DOI: 10.1016/j.envpol.2019.01.085.
Napoli, M. T., E. Sciaky, D. J. Arya, and N. Balos. 2017. PIPELINES: Fostering university-
community college partnerships and STEM professional success for underrepresented
populations. Presented at ASEE Annual Conference & Exposition. Columbus, Ohio.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

References 291

NASEM (National Academies of Sciences, Engineering, and Medicine). 2016. Barriers and
opportunities for 2-year and 4-year STEM degrees: Systemic change to support students’
diverse pathways. Washington, DC: The National Academies Press. DOI:
10.17226/21739.
NASEM. 2017. Communities in action: Pathways to health equity. Washington, DC: The National
Academies Press. DOI: 10.17226/24624.
NASEM. 2018. Sexual harassment of women: Climate, culture, and consequences in academic
sciences, engineering, and medicine. Washington, DC: The National Academies Press.
DOI: 10.17226/24994.
NASEM. 2019a. Gaseous carbon waste streams utilization: Status and research needs.
Washington, DC: The National Academies Press. DOI: 10.17226/25232.
NASEM. 2019b. Negative emissions technologies and reliable sequestration: A research agenda.
Washington, DC: The National Academies Press. DOI: 10.17226/25259.
NASEM. 2019c. Environmental engineering for the 21st century: Addressing grand challenges.
Washington, DC: The National Academies Press. DOI: 10.17226/25121.
NASEM. 2019d. Frontiers of materials research: A decadal survey. Washington, DC: The
National Academies Press. DOI: 10.17226/25244.
NASEM. 2019e. The science of effective mentorship in STEMM. Washington, DC: The National
Academies Press. DOI: 10.17226/25568.
NASEM. 2019f. Deployment of deep decarbonization technologies: Proceedings of a workshop.
Washington, DC: The National Academies Press. DOI: 10.17226/25656.
NASEM. 2020a. Promising practices for addressing the underrepresentation of women in science,
engineering, and medicine: Opening doors. Washington, DC: The National Academies
Press. DOI: 10.17226/25585.
NASEM. 2020b. The impacts of racism and bias on Black people pursuing careers in science,
engineering, and medicine: Proceedings of a workshop. Washington, DC: The National
Academies Press. DOI: 10.17226/25849.
NASEM. 2021a. Accelerating decarbonization of the U.S. energy system. Washington, DC: The
National Academies Press. DOI: 10.17226/25932.
NASEM. 2021b. Innovations in pharmaceutical manufacturing on the horizon: Technical
challenges, regulatory issues, and recommendations. Washington, DC: The National
Academies Press. DOI: 10.17226/26009.
NASEM. 2021c. Diversity, equity, and inclusion in chemistry and chemical engineering:
Proceedings of a workshop-in-brief. Washington, DC: The National Academies Press.
DOI: 10.17226/26334.
Nature. 2019. Human Microbiome Project, part 2. Retrieved August 16, 2021, from
https://www.nature.com/collections/fiabfcjbfj.
NCES (National Center for Education Statistics). 2021. Expenditures. Retrieved March 18, 2021,
2021, from https://nces.ed.gov/fastfacts/display.asp?id=75.
NCSES (National Center for Science and Engineering Statistics). 2018. Science and engineering
degrees, by race and ethnicity of receipts: 2008-18. Alexandria, VA: NCSES.
https://ncsesdata.nsf.gov/sere/2018/index.html.
NCSES. 2020. National patterns of R&D resources: 2017-18 data update. Alexandria, VA:
NCSES. https://ncses.nsf.gov/pubs/nsf20307/#general-notes.
NCSES. 2021. Explore data. Retreived 2021, from https://www.nsf.gov/statistics/data.cfm.
NIMS (National Institute for Materials Science). 2021. Polymer database (polyinfo). Retrieved
August 27, 2021, from https://polymer.nims.go.jp/en.
NIST (National Institute of Standards and Technology). 2021. Thermophysical properties of fluid
systems. Retrieved August 27, 2021, from https://webbook.nist.gov/chemistry/fluid.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

292 New Directions for Chemical Engineering

NOAA (National Oceanic and Atmospheric Administration). 2020. Climate change: Atmospheric
carbon dioxide. Retrieved September 11, 2021, from https://www.climate.gov/news-
features/understanding-climate/climate-change-atmospheric-carbon-dioxide.
NOAA. 2021a. Deepwater Horizon. Retrieved August 16, 2021, from https://darrp.noaa.gov/oil-
spills/deepwater-horizon.
NOAA. 2021b. A guide to plastic in the ocean. Retrieved August 16, 2021, from
https://oceanservice.noaa.gov/hazards/marinedebris/plastics-in-the-ocean.html.
NREL (National Renewable Energy Laboratory). 2017. Concentrating solar power gen3
demonstration roadmap. Retrieved August 16, 2021, from https://www.nrel.gov/docs/
fy17osti/67464.pdf.
NREL. 2020. NREL’s top 2020 wind program accomplishments demonstrate a clear vision for
wind energy advancement. Retrieved August 17, 2021, from https://www.nrel.gov/
news/program/2020/2020-top-wind-accomplishments.html.
NRC (National Research Council). 1988. Frontiers in chemical engineering: Research needs and
opportunities. Washington, DC: The National Academies Press. DOI: 10.17226/1095.
NRC. 2007. International benchmarking of U.S. chemical engineering research competitiveness.
Washington, DC: The National Academies Press. DOI: 10.17226/11867.
NRC. 2012. Discipline-based education research: Understanding and improving learning in un-
dergraduate science and engineering. Washington, DC: The National Academies Press.
DOI: 10.17226/13362.
NRC. 2015. Cost, effectiveness, and deployment of fuel economy technologies for light-duty
vehicles. Washington, DC: The National Academies Press. DOI: 10.17226/21744.
NSB (National Science Board). 2018. Science and engineering indicators NSB-2018-1.
Alexandria, VA: National Science Foundation.
NSB. 2020. Merit review process fiscal year 2019 digest. Retrieved August 16, 2021, from
https://www.nsf.gov/nsb/publications/2020/merit_review/FY-2019/nsb202038.pdf.
NSF (National Science Foundation). 2016. U.S. science and technology leadership increasingly
challenged by advances in Asia. Retrieved 2021, from https://nsf.gov/news/news_
summ.jsp?cntn_id=137394&org=NSF&from=news.
NSF. 2020. NSF & Congress. Retrieved August 16, 2021, from https://www.nsf.gov/about/
congress/117/highlights/cu20.jsp.
Nelson, M. J., G. Nakhla, and J. Zhu. 2017. Fluidized-bed bioreactor applications for biological
wastewater treatment: A review of research and developments. Engineering 3(3):330-
342. DOI: 10.1016/j.Eng.2017.03.021.
Neri, G., P. M. Donaldson, and A. J. Cowan. 2017. The role of electrode-catalyst interactions in
enabling efficient CO2 reduction with Mo(bpy)(CO)4 as revealed by vibrational sum-
frequency generation spectroscopy. Journal of the American Chemical Society
139(39):13791-13797. DOI: 10.1021/jacs.7b06898.
Ng, A. 2016. What artificial intelligence can and can’t do right now. Harvard Business Review.
https://hbr.org/2016/11/what-artificial-intelligence-can-and-cant-do-right-now.
Ng, W. L., C. K. Chua, and Y.-F. Shen. 2019. Print me an organ! Why we are not there yet.
Progress in Polymer Science 97. DOI: 10.1016/j.progpolymsci.2019.101145.
Nielsen. 2015. The sustainability imperative—New insights on consumer expectations. Retrieved
August 18, 2021, from https://www.nielsen.com/wp-content/uploads/sites/3/2019/04/
Global20Sustainability20Report_October202015.pdf.
Nietzel, M. T. 2021. Georgia Tech’s online MS in computer science continues to thrive. Why that’s
important for the future of MOOCs. Retrieved September 8, 2021, from
https://www.forbes.com/sites/michaeltnietzel/2021/07/01/georgia-techs-online-ms-in-co

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

References 293

mputer-science-continues-to-thrive-what-that-could-mean-for-the-future-of-moocs/?sh=
750c608aa277.
Nikolau, B. J., M. A. D. N. Perera, L. Brachova, and B. Shanks. 2008. Platform biochemicals for
a biorenewable chemical industry. The Plant Journal 54(4):536-545. DOI: 10.1111/
j.1365-313X.2008.03484.x.
Nimpuno, N., and C. Scruggs. 2011. Information on chemicals in electronic products: A study of
needs, gaps, obstacles and solutions to provide and access information on chemicals in
electronic products. Copenhagen: Nordic Council of Ministers.
Nisbet, E. G., M. R. Manning, E. J. Dlugokencky, R. E. Fisher, D. Lowry, S. E. Michel, C. L.
Myhre, S. M. Platt, G. Allen, P. Bousquet, R. Brownlow, M. Cain, J. L. France, O.
Hermansen, R. Hossaini, A. E. Jones, I. Levin, A. C. Manning, G. Myhre, J. A. Pyle, B.
H. Vaughn, N. J. Warwick, and J. W. C. White. 2019. Very strong atmospheric methane
growth in the 4 years 2014–2017: Implications for the Paris Agreement. Global
Biogeochemical Cycles 33(3):318-342. DOI: 10.1029/2018gb006009.
Nitopi, S., E. Bertheussen, S. B. Scott, X. Liu, A. K. Engstfeld, S. Horch, B. Seger, I. E. L.
Stephens, K. Chan, C. Hahn, J. K. Norskov, T. F. Jaramillo, and I. Chorkendorff. 2019.
Progress and perspectives of electrochemical CO2 Reduction on copper in aqueous
electrolyte. Chemical Reviews 119(12):7610-7672. DOI: 10.1021/acs.chemrev.8b00705.
Nnodu, O., H. Isa, M. Nwegbu, C. Ohiaeri, S. Adegoke, R. Chianumba, N. Ugwu, B. Brown, J.
Olaniyi, E. Okocha, J. Lawson, A. A. Hassan, I. Diaku-Akinwumi, A. Madu, O.
Ezenwosu, Y. Tanko, U. Kangiwa, A. Girei, Y. Israel-Aina, A. Ladu, P. Egbuzu, U.
Abjah, A. Okolo, N. Akbulut-Jeradi, M. Fernandez, F. B. Piel, and A. Adekile. 2019.
HemoTypeSC, a low-cost point-of-care testing device for sickle cell disease: Promises
and challenges. Blood Cells, Molecules and Diseases 78:22-28. DOI: 10.1016/j.bcmd.
2019.01.007.
Noé, F., A. Tkatchenko, K. R. Müller, and C. Clementi. 2020. Machine learning for molecular
simulation. Annual Review of Physical Chemistry 71:361-390. DOI: 10.1146/annurev-
physchem-042018-052331.
Nordquist, A. F., F. C. Wilhelm, F. J. Waller, and R. M. Machado. 2002. Hydrogenation with
monolith reactor under conditions of immiscible liquid phases. U.S. Patent No. 6479704.
Washington, DC: U.S. Patent and Trademark Office.
Nuss, P. 2015. Book review: Life cycle assessment handbook: A guide for environmentally
sustainable products. M. A. Curran, ed. Hoboken, NJ: PB - John Wiley & Sons, Inc., and
Salem, MA: Scrivener Publishing LLC.
O’Brien, E. J., J. M. Monk, and B. O. Palsson. 2015. Using genome-scale models to predict
biological capabilities. Cell 161(5):971-987. DOI: 10.1016/j.cell.2015.05.019.
Ochedi, F. O., D. Liu, J. Yu, A. Hussain, and Y. Liu. 2020. Photocatalytic, electrocatalytic and
photoelectrocatalytic conversion of carbon dioxide: A review. Environmental Chemistry
Letters 19(2):941-967. DOI: 10.1007/s10311-020-01131-5.
OECD (Organisation for Economic Co-operation and Development). 2012. OECD environmental
outlook to 2050. OECD Publishing. DOI: 10.1787/9789264122246-en.
Ogden, J., L. Fulton, and D. Sperling. 2016. Making the transition to light-duty electric-drive
vehicles in the U.S.: Costs in perspective to 2035. Davis: University of California, Davis.
https://trid.trb.org/view/1441689.
Ogunnaike, B. A. 2019. 110th anniversary: Process and systems engineering perspectives on
personalized medicine and the design of effective treatment of diseases. Industrial &
Engineering Chemistry Research 58(44):20357-20369. DOI: 10.1021/acs.iecr.9b04228.
Ohio History Central. 2021. Standard oil company. Retrieved August 17, 2021, from
https://ohiohistorycentral.org/w/Standard_Oil_Company.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

294 New Directions for Chemical Engineering

Olivetti, E. A., J. M. Cole, E. Kim, O. Kononova, G. Ceder, T. Y.-J. Han, and A. M. Hiszpanski.
2020. Data-driven materials research enabled by natural language processing and
information extraction. Applied Physics Reviews 7(4). DOI: 10.1063/5.0021106.
Olson, J., Y. Cao, J. Romero, P. Johnson, P.-L. Dallaire-Demers, N. Sawaya, P. Narang, I.
Kivlichan, M. Wasielewski, and A. Aspuru-Guzik. 2016. Quantum information and
computation for chemistry, report of an NSF workshop. arXiv. DOI: 1706.05413.
Oluwole, E. O., T. A. Adeyemo, G. E. Osanyin, O. O. Odukoya, P. J. Kanki, and B. B. Afolabi.
2020. Feasibility and acceptability of early infant screening for sickle cell disease in
Lagos, Nigeria—A pilot study. PLoS One 15(12). DOI: 10.1371/journal.pone.0242861.
O’Neill, J., and J.-F. Zheng. 2019. The holistic approach to materials and processing for new and
scaled devices. Semiconductor Digest. https://www.semiconductor-digest.com/the-holis
tic-approach-to-materials-and-processing-for-new-and-scaled-devices.
O’Regan, B., and M. Grätzel. 1991. A low-cost, high-efficiency solar cell based on dye-sensitized
colloidal TiO2 films. Nature 353(6346):737-740. DOI: 10.1038/353737a0.
Our World in Data. 2019. Number of deaths by risk factor, world, 2017. Retrieved August 17,
2021 from https://ourworldindata.org/grapher/number-of-deaths-by-risk-factor.
Our World in Data. 2021a. Emissions of air pollutants, United States, 1970 to 2016. Retrieved
August 17, 2021, from https://ourworldindata.org/grapher/emissions-of-air-pollutants?
country=~USA.
Our World in Data. 2021b. Food: Greenhouse gas emissions across the supply chain. Retrieved
August 17, 2021, from https://ourworldindata.org/grapher/food-emissions-supply-
chain?country=Beef+%28beef+herd%29~Cheese~Poultry+Meat~Milk~Eggs~Rice~Pig
+Meat~Peas~Bananas~Wheat+%26+Rye~Fish+%28farmed%29~Lamb+%26+Mutton~
Beef+%28dairy+herd%29~Shrimps+%28farmed%29~Tofu~Maize.
Owoseni, O., E. Nyankson, Y. Zhang, S. J. Adams, J. He, G. L. McPherson, A. Bose, R. B. Gupta,
and V. T. John. 2014. Release of surfactant cargo from interfacially-active halloysite clay
nanotubes for oil spill remediation. Langmuir 30(45):13533-13541. DOI: 10.1021/la
503687b.
Pacala, S., and R. Socolow. 2004. Stabilization wedges: Solving the climate problem for the next
50 years with current technologies. Science 305(5686):968-972. DOI: 10.1126/science.
1100103.
Padervand, M., E. Lichtfouse, D. Robert, and C. Wang. 2020. Removal of microplastics from the
environment. A review. Environmental Chemistry Letters 18(3):807-828. DOI: 10.1007/
s10311-020-00983-1.
Pal, S., and K. A. Fichthorn. 1999. Accelerated molecular dynamics of infrequent events. Chemical
Engineering Journal 74(1):77-83. DOI: 10.1016/S1385-8947(99)00055-8.
Palmer, J. C., P. H. Poole, F. Sciortino, and P. G. Debenedetti. 2018. Advances in computational
studies of the liquid-liquid transition in water and water-like models. Chemical Reviews
118(18):9129-9151. DOI: 10.1021/acs.chemrev.8b00228.
Panagiotopoulos, A. Z. 1987. Direct determination of phase coexistence properties of fluids by
Monte Carlo simulation in a new ensemble. Molecular Physics 61(4):813-826. DOI:
10.1080/00268978700101491.
Panuwatsuk, W., and N. A. Da Silva. 2003. Application of a gratuitous induction system in
Kluyveromyces lactis for the expression of intracellular and secreted proteins during fed-
batch culture. Biotechnology and Bioengineering 81(6):712-718. DOI: 10.1002/bit.
10518.
Pappa, G., C. Boukouvalas, C. Giannaris, N. Ntaras, V. Zografos, K. Magoulas, A. Lygeros, and
D. Tassios. 2001. The selective dissolution/precipitation technique for polymer recycling:

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

References 295

A pilot unit application. Resources, Conservation and Recycling 34(1):33-44. DOI:


10.1016/s0921-3449(01)00092-1.
Park, H., and K. Park. 1996. Biocompatibility issues of implantable drug delivery systems.
Pharmaceutical Research 13(12):1770-1776. DOI: 10.1023/a:1016012520276.
Park, D. S., K. E. Joseph, M. Koehle, C. Krumm, L. Ren, J. N. Damen, M. H. Shete, H. S. Lee, X.
Zuo, B. Lee, W. Fan, D. G. Vlachos, R. F. Lobo, M. Tsapatsis, and P. J. Dauenhauer.
2016. Tunable oleo-furan surfactants by acylation of renewable furans. ACS Central
Science 2(11):820-824. DOI: 10.1021/acscentsci.6b00208.
Park, S.-Y., C.-H. Park, D.-H. Choi, J. K. Hong, and D.-Y. Lee. 2021. Bioprocess digital twins of
mammalian cell culture for advanced biomanufacturing. Current Opinion in Chemical
Engineering 33. DOI: 10.1016/j.coche.2021.100702.
Parsons, S., S. Raikova, and C. J. Chuck. 2020. The viability and desirability of replacing palm oil.
Nature Sustainability 3(6):412-418. DOI: 10.1038/s41893-020-0487-8.
Paul, R., and R. Brennan. 2019. Discipline-based education research (DBER)—What is it, and why
should engineering education research scholars be talking about it more? Presented at
Canadian Engineering Education Association Conference. Ottawa, Ontario.
Peng, L., H. Dai, Y. Wu, Y. Peng, and X. Lu. 2018. A comprehensive review of the available
media and approaches for phosphorus recovery from wastewater. Water, Air, & Soil Pol-
lution 229(4). DOI: 10.1007/s11270-018-3706-4.
Pereao, O., C. Bode-Aluko, O. Fatoba, L. Petrik, and K. Laatikainen. 2018. Rare earth elements
removal techniques from water/wastewater: A review. Desalination and Water Treatment
130:71-86. DOI: 10.5004/dwt.2018.22844.
Perlmutter, D. D. 1975. Introduction to chemical process control. Malabar, FL: Robert E. Krieger
Publishing Company.
Peters, M., K. Timmerhaus, R. West, and M. Peters. 2002. Plant design and economics for
chemical engineers. New York: McGraw-Hill Education.
Plastics Hall of Fame. 2021. Daniel Wayne Fox. Retrieved August 17, 2021, from
https://www.plasticshof.org/members/daniel-wayne-fox.
PlasticsEurope. 2019. Plastics—The facts 2019: An analysis of European plastics production,
demand and waste data. Retrieved 2021, from https://plasticseurope.org/wp-content/
uploads/2021/10/2019-Plastics-the-facts.pdf.
Pozrikidis, C. 1997. Numerical studies of singularity formation at free surfaces and fluid interfaces
in two-dimensional Stokes flow. Journal of Fluid Mechanics 331:145-167. DOI:
10.1017/S0022112096003813.
Prather, K. A., C. C. Wang, and R. T. Schooley. 2020. Reducing transmission of SARS-CoV-2.
Science 368(6498):1422-1424. DOI: 10.1126/science.abc6197.
Prausnitz, M. R. 2017. Engineering microneedle patches for vaccination and drug delivery to skin.
Annual Review of Chemical and Biomolecular Engineering 8:177-200. DOI: 10.1146/
annurev-chembioeng-060816-101514.
Prausnitz, M. R., and R. Langer. 2008. Transdermal drug delivery. Nature Biotechnology
26(11):1261-1268. DOI: 10.1038/nbt.1504.
PRISMS Center (Center for Predictive Integrated Structural Materials Science). 2021. Materials
common 2.0. Retrieved August 27, 2021, from https://materialscommons.org.
Ragauskas, A. J., G. T. Beckham, M. J. Biddy, R. Chandra, F. Chen, M. F. Davis, B. H. Davison,
R. A. Dixon, P. Gilna, M. Keller, P. Langan, A. K. Naskar, J. N. Saddler, T. J.
Tschaplinski, G. A. Tuskan, and C. E. Wyman. 2014. Lignin valorization: Improving
lignin processing in the biorefinery. Science 344(6185). DOI: 10.1126/science.1246843.
Rahimi, A., and J. M. García. 2017. Chemical recycling of waste plastics for new materials
production. Nature Reviews Chemistry 1(6). DOI: 10.1038/s41570-017-0046.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

296 New Directions for Chemical Engineering

Raišienė, A. G., V. Rapuano, and K. Varkulevičiūtė. 2021. Sensitive men and hardy women: How
do millennials, xennials and gen X manage to work from home? Journal of Open
Innovation: Technology, Market, and Complexity 7(2). DOI: 10.3390/joitmc7020106.
Ramadan, Q., and M. Zourob. 2020. Organ-on-a-chip engineering: Toward bridging the gap
between lab and industry. Biomicrofluidics 14(4). DOI: 10.1063/5.0011583.
Raman, A. P., M. A. Anoma, L. Zhu, E. Rephaeli, and S. Fan. 2014. Passive radiative cooling
below ambient air temperature under direct sunlight. Nature 515(7528):540-544. DOI:
10.1038/nature13883.
Rao, D. P. 2015. The Story of “HIGEE”. Indian Chemical Engineer 57(3-4):282-299. DOI:
10.1080/00194506.2015.1026946.
Ratnasari, D. K., M. A. Nahil, and P. T. Williams. 2017. Catalytic pyrolysis of waste plastics using
staged catalysis for production of gasoline range hydrocarbon oils. Journal of Analytical
and Applied Pyrolysis 124:631-637. DOI: 10.1016/j.jaap.2016.12.027.
Renzulli, K. A. 2019. Women reach leadership roles earlier than men do—But fewer make it to
the top, according to LinkedIn. Retrieved September 8, 2021, from https://www.cnbc.
com/2019/06/24/women-reach-leadership-roles-1point4-years-earlier-than-men-says-link
edin.html.
Rephaeli, E., A. Raman, and S. Fan. 2013. Ultrabroadband photonic structures to achieve high-
performance daytime radiative cooling. Nano Letters 13(4):1457-1461. DOI:
10.1021/nl4004283.
Reynaert, S., C. F. Brooks, P. Moldenaers, J. Vermant, and G. G. Fuller. 2008. Analysis of the
magnetic rod interfacial stress rheometer. Journal of Rheology 52(1):261-285. DOI:
10.1122/1.2798238.
Rezaiyan, J., and N. P. Cheremisinoff. 2005. Gasification technologies: A primer for engineers
and scientists. Boca Raton, FL: CRC Press.
Richards, G. 1979. Third age of quantum chemistry. Nature 278(5704):507-507. DOI:
10.1038/278507a0.
Richardson, J., and J. Dantzler. 2002. Effect of a freshman engineering program on retention and
academic performance. Presented at 32nd Annual Frontiers in Education. Boston, MA..
Ritger, P. L., and N. A. Peppas. 1987. A simple equation for description of solute release I. Fickian
and non-fickian release from non-swellable devices in the form of slabs, spheres,
cylinders or discs. Journal of Controlled Release 5(1):23-36. DOI: 10.1016/0168-
3659(87)90034-4.
Roh, S., A. H. Williams, R. S. Bang, S. D. Stoyanov, and O. D. Velev. 2019. Soft dendritic
microparticles with unusual adhesion and structuring properties. Nature Materials
18(12):1315-1320. DOI: 10.1038/s41563-019-0508-z.
Rong, Y., Y. Hu, A. Mei, H. Tan, M. I. Saidaminov, S. I. Seok, M. D. McGehee, E. H. Sargent,
and H. Han. 2018. Challenges for commercializing perovskite solar cells. Science
361(6408). DOI: 10.1126/science.aat8235.
Roque, B. M., M. Venegas, R. D. Kinley, R. de Nys, T. L. Duarte, X. Yang, and E. Kebreab. 2021.
Red seaweed (Asparagopsis taxiformis) supplementation reduces enteric methane by over
80 percent in beef steers. PLoS One 16(3). DOI: 10.1371/journal.pone.0247820.
Rorrer, J. E., G. T. Beckham, and Y. Román-Leshkov. 2021. Conversion of polyolefin waste to
liquid alkanes with Ru-based catalysts under mild conditions. Journal of the American
Chemical Society: Au 1(1):8-12. DOI: 10.1021/jacsau.0c00041.
Rosales, A. M., and K. S. Anseth. 2016. The design of reversible hydrogels to capture extracellular
matrix dynamics. Nature Reviews Materials 1(2). DOI: 10.1038/natrevmats.2015.12.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

References 297

Rosales, A. M., R. A. Segalman, and R. N. Zuckermann. 2013. Polypeptoids: A model system to


study the effect of monomer sequence on polymer properties and self-assembly. Soft
Matter 9(35):8400-8414. DOI: 10.1039/C3SM51421H.
Ross, I., J. McDonough, J. Miles, P. Storch, P. Thelakkat Kochunarayanan, E. Kalve, J. Hurst, S.
S. Dasgupta, and J. Burdick. 2018. A review of emerging technologies for remediation of
PFASs. Remediation Journal 28(2):101-126. DOI: 10.1002/rem.21553.
Ruiz-Lopez, M. F., J. S. Francisco, M. T. C. Martins-Costa, and J. M. Anglada. 2020. Molecular
reactions at aqueous interfaces. Nature Reviews Chemistry 4(9):459-475. DOI:
10.1038/s41570-020-0203-2.
Rzhetsky, A., J. G. Foster, I. T. Foster, and J. A. Evans. 2015. Choosing experiments to accelerate
collective discovery. Proceedings of the National Academy of Sciences 112(47):14569-
14574. DOI: 10.1073/pnas.1509757112.
Samet, J. M., K. Prather, G. Benjamin, S. Lakdawala, J. M. Lowe, A. Reingold, J. Volckens, and
L. Marr. 2021. Airborne transmission of SARS-CoV-2: What we know. Clinical
Infectious Diseases 73(10):1924-1926. DOI: 10.1093/cid/ciab039.
Samuel, M., D. Polson, D. Graham, W. Kordziel, T. Waite, G. Waters, P. S. Vinod, D. Fu, and R.
Downey. 2000. Viscoelastic surfactant fracturing fluids: Applications in low permeability
reservoirs. Presented at SPE Rocky Mountain Regional/Low-Permeability Reservoirs
Symposium and Exhibition. Denver, CO.
Sánchez, A. 2019. The current role of chemical engineering in solving environmental problems.
Frontiers in Chemical Engineering 1, Article 1. DOI: 10.3389/fceng.2019.00001.
Sanchez-Lengeling, B., and A. Aspuru-Guzik. 2018. Inverse molecular design using machine
learning: Generative models for matter engineering. Science 361(6400):360-365. DOI:
10.1126/science.aat2663.
Sanderson, K. 2019. Automation: Chemistry shoots for the moon. Nature 568(7753):577-579.
DOI: 10.1038/d41586-019-01246-y.
Santiesteban, J. G. and T. F. Degnan., Jr. 2021. Catalysis and the future of transportation fuels. The
Bridge 50th Anniversary Issue. https://www.nae.edu/244855/Catalysis-and-the-Future-
of-Transportation-Fuels.
Sargent & Lundy LLC Consulting Group. 2003. Assessment of parabolic trough and power tower
solar technology cost and performance forecasts NREL/SR-550-34440. Chicago, IL:
National Renewable Energy Laboratory (NREL).
Sarikurt, S., T. Kocabaş, and C. Sevik. 2020. High-throughput computational screening of 2D
materials for thermoelectrics. Journal of Materials Chemistry A 8(37):19674-19683.
DOI: 10.1039/d0ta04945j.
Sarsons, H. 2017. Gender differences in recognition for group work. https://scholar.harvard.
edu/files/sarsons/files/full_v6.pdf.
Saygili, E., E. Yildiz-Ozturk, M. J. Green, A. M. Ghaemmaghami, and O. Yesil-Celiktas. 2021.
Human lung-on-chips: Advanced systems for respiratory virus models and assessment of
immune response. Biomicrofluidics 15(2). DOI: 10.1063/5.0038924.
Scanlon, B. R., B. L. Ruddell, P. M. Reed, R. I. Hook, C. Zheng, V. C. Tidwell, and S. Siebert.
2017. The food-energy-water nexus: Transforming science for society. Water Resources
Research 53(5):3550-3556. DOI: 10.1002/2017wr020889.
Schepers, A., C. Li, A. Chhabra, B. T. Seney, and S. Bhatia. 2016. Engineering a perfusable 3D
human liver platform from iPS cells. Lab on a Chip 16(14):2644-2653. DOI:
10.1039/c6lc00598e.
Schiffer, Z. J., and K. Manthiram. 2017. Electrification and decarbonization of the chemical
industry. Joule 1(1):10-14. DOI: 10.1016/j.joule.2017.07.008.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

298 New Directions for Chemical Engineering

Schöttker, B., K.-U. Saum, D. C. Muhlack, L. K. Hoppe, B. Holleczek, and H. Brenner. 2017.
Polypharmacy and mortality: New insights from a large cohort of older adults by detection
of effect modification by multi-morbidity and comprehensive correction of confounding
by indication. European Journal of Clinical Pharmacology 73(8):1041-1048. DOI:
10.1007/s00228-017-2266-7.
Schowalter, W. R. 2003. The equations (of change) don’t change: But the profession of engineering
does. Chemical Engineering Education 37(4):242-247.
Schwaller, P., T. Laino, T. Gaudin, P. Bolgar, C. A. Hunter, C. Bekas, and A. A. Lee. 2019.
Molecular Transformer: A model for uncertainty-calibrated chemical reaction prediction.
ACS Central Science 5(9):1572-1583. DOI: 10.1021/acscentsci.9b00576.
Scriven, L. E. 1960. Dynamics of a fluid interface equation of motion for Newtonian surface fluids.
Chemical Engineering Science 12(2):98-108. DOI: 10.1016/0009-2509(60)87003-0.
Scriven, L. E. 1991. On the emergence and evolution of chemical engineering. In Advances in
chemical engineering. Colton, C. K., ed. San Diego, CA: Academic Press.
Seader, J. D., and E. J. Henley. 1998. Separation process principles. New York: Wiley.
Searchinger, T., R. Heimlich, R. A. Houghton, F. Dong, A. Elobeid, J. Fabiosa, S. Tokgoz, D.
Hayes, and T.-H. Yu. 2008. Use of U.S. croplands for biofuels increases greenhouse gases
through emissions from land-use change. Science 319(5867):1238-1240. DOI:
10.1126/science.1151861.
Seinfeld, J. H. 1991. Environmental chemical engineering. In Advances in chemical engineering.
Colton, C. K., ed. San Diego, CA: Academic Press.
Seinfeld, J. H., and S. N. Pandis. 2016. Atmospheric chemistry and physics: From air pollution to
climate change. New York: John Wiley & Sons, Inc.
Sharma, S. 2013. Ferrolectric nanofibers: Principle, processing and applications. Advanced
Materials Letters 4(7):522-533. DOI: 10.5185/amlett.2012.9426.
Shashvatt, U., F. Amurrio, C. Portner, and L. Blaney. 2021. Phosphorus recovery by Donnan dial-
ysis: Membrane selectivity, diffusion coefficients, and speciation effects. Chemical En-
gineering Journal 419. DOI: 10.1016/j.cej.2021.129626.
Shen, L., E. Worrell, and M. K. Patel. 2010. Open-loop recycling: A LCA case study of PET bottle-
to-fibre recycling. Resources, Conservation and Recycling 55(1):34-52. DOI:
10.1016/j.resconrec.2010.06.014.
Sherman, Z. M., M. P. Howard, B. A. Lindquist, R. B. Jadrich, and T. M. Truskett. 2020. Inverse
methods for design of soft materials. The Journal of Chemical Physics 152(14). DOI:
10.1063/1.5145177.
Shi, D., E. Dassau, and F. J. Doyle, 3rd. 2019. Multivariate learning framework for long-term
adaptation in the artificial pancreas. Bioengineering & Translational Medicine 4(1):61-
74. DOI: 10.1002/btm2.10119.
Shi, C., L. T. Reilly, V. S. Phani Kumar, M. W. Coile, S. R. Nicholson, L. J. Broadbelt, G. T.
Beckham, and E. Y. X. Chen. 2021. Design principles for intrinsically circular polymers
with tunable properties. Chem 7(11):2896-2912. DOI: 10.1016/j.chempr.2021.10.004.
Shim, S. H., R. Gupta, Y. L. Ling, D. B. Strasfeld, D. P. Raleigh, and M. T. Zanni. 2009. Two-
dimensional IR spectroscopy and isotope labeling defines the pathway of amyloid
formation with residue-specific resolution. Proceedings of the National Academy of
Sciences 106(16):6614-6619. DOI: 10.1073/pnas.0805957106.
Sidorova, J., and M. Anisimova. 2014. NLP-inspired structural pattern recognition in chemical
application. Pattern Recognition Letters 45:11-16. DOI: 10.1016/j.patrec.2014.02.012.
Silva, R. A., K. Hawboldt, and Y. Zhang. 2018. Application of resins with functional groups in the
separation of metal ions/species—A review. Mineral Processing and Extractive
Metallurgy Review 39(6):395-413. DOI: 10.1080/08827508.2018.1459619.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

References 299

Simpson, G. B., and G. P. W. Jewitt. 2019. The development of the water-energy-food nexus as a
framework for achieving resource security: A review. Frontiers in Environmental Science
7, Article 8. DOI: 10.3389/fenvs.2019.00008.
Sims, R. E. H., R. N. S., A. Adegbululgbe, J. Fenhann, I. Konstantinaviciute, W. Moomaw, H.B.
Nimir, B. Schlamadinger, J. Torres-Martínez, C. Turner, Y. Uchiyama, S. J.V. Vuori, N.
Wamukonya, X. Zhang 2007. Chapter 4: Energy supply. In IPCC Fourth Assessment
Report: Climate Change 2007: Contribution of Working Group III: Mitigation of Climate
Change. New York: Cambridge University Press.
Singh, A., N. A. Rorrer, S. R. Nicholson, E. Erickson, J. S. DesVeaux, A. F. T. Avelino, P. Lamers,
A. Bhatt, Y. Zhang, G. Avery, L. Tao, A. R. Pickford, A. C. Carpenter, J. E. McGeehan,
and G. T. Beckham. 2021. Techno-economic, life-cycle, and socioeconomic impact anal-
ysis of enzymatic recycling of poly(ethylene terephthalate). Joule 5(9):2479-2503. DOI:
10.1016/j.joule.2021.06.015.
Sinha, S., U. D. Irani, V. Manchaiah, and M. S. Bhamla. 2020. LoCHAid: An ultra-low-cost
hearing aid for age-related hearing loss. PLoS One 15(9). DOI: 10.1371/journal.
pone.0238922.
Siracusa, V., and I. Blanco. 2020. Bio-polyethylene (Bio-PE), bio-polypropylene (Bio-PP) and
bio-poly (ethylene terephthalate) (Bio-PET): Recent developments in bio-based polymers
analogous to petroleum-derived ones for packaging and engineering applications.
Polymers 12(8). DOI: 10.3390/polym12081641.
Sivakumar, S., K. L. Wark, J. K. Gupta, N. L. Abbott, and F. Caruso. 2009. Liquid crystal
emulsions as the basis of biological sensors for the optical detection of bacteria and
viruses. Advanced Functional Materials 19(14):2260-2265. DOI: 10.1002/adfm.2009
00399.
Skoulidas, A. I., D. M. Ackerman, J. K. Johnson, and D. S. Sholl. 2002. Rapid transport of gases
in carbon nanotubes. Physical Review Letters 89(18). DOI: 10.1103/PhysRevLett.89.
185901.
Smith, R. C., R. K. Taggart, J. C. Hower, M. R. Wiesner, and H. Hsu-Kim. 2019. Selective recov-
ery of rare earth elements from coal fly ash leachates using liquid membrane processes.
Environmental Science & Technology 53(8):4490-4499. DOI: 10.1021/acs.est.9b00539.
Smith, C., A. K. Hill, and L. Torrente-Murciano. 2020. Current and future role of Haber–Bosch
ammonia in a carbon-free energy landscape. Energy & Environmental Science 13(2):331-
344. DOI: 10.1039/c9ee02873k.
Socorro, I. M., K. Taylor, and J. M. Goodman. 2005. ROBIA: A reaction prediction program.
Organic Letters 7(16):3541-3544. DOI: 10.1021/ol0512738.
Song, X., Y. Guo, J. Zhang, N. Sun, G. Shen, X. Chang, W. Yu, Z. Tang, W. Chen, W. Wei, L.
Wang, J. Zhou, X. Li, X. Li, J. Zhou, and Z. Xue. 2019. Fracturing with carbon dioxide:
From microscopic mechanism to reservoir application. Joule 3(8):1913-1926. DOI:
10.1016/j.joule.2019.05.004.
Spielman, L. A. 1977. Particle capture from low-speed laminar flows. Annual Review of Fluid
Mechanics 9(1):297-319. DOI: 10.1146/annurev.fl.09.010177.001501.
Spiker, M. L., H. A. B. Hiza, S. M. Siddiqi, and R. A. Neff. 2017. Wasted food, wasted nutrients:
Nutrient loss from wasted food in the United States and comparison to gaps in dietary
intake. Journal of the Academy of Nutrition and Dietetics 117(7):1031-1040.e22. DOI:
10.1016/j.jand.2017.03.015.
Squires, T. M., and T. G. Mason. 2009. Fluid mechanics of microrheology. Annual Review of Fluid
Mechanics 42(1):413-438. DOI: 10.1146/annurev-fluid-121108-145608.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

300 New Directions for Chemical Engineering

Stankiewicz, A. I., and J. A. Moulijn. 2000. Process intensification: Transforming chemical


engineering. CEP Magazine. https://www.aiche.org/sites/default/files/docs/news/010022
_cep_stankiewicz.pdf.
Statista. 2021. Forecast number of mobile devices worldwide from 2020 to 2025 (in billions).
Retrieved September 11, 2021, from https://www.statista.com/statistics/245501/multiple-
mobile-device-ownership-worldwide.
Staudinger, H. 1920. Über polymerisation. Berichte der deutschen chemischen Gesellschaft (A and
B Series) 53(6):1073-1085. DOI: 10.1002/cber.19200530627.
Stebe, K. J., S. Y. Lin, and C. Maldarelli. 1991. Remobilizing surfactant retarded fluid particle
interfaces. I. Stress‐free conditions at the interfaces of micellar solutions of surfactants
with fast sorption kinetics. Physics of Fluids A: Fluid Dynamics 3(1):3-20. DOI:
10.1063/1.857862.
Stone, H. A., and L. G. Leal. 1989. Relaxation and breakup of an initially extended drop in an
otherwise quiescent fluid. Journal of Fluid Mechanics 198:399-427. DOI:
10.1017/S0022112089000194.
Striolo, A. 2006. The mechanism of water diffusion in narrow carbon nanotubes. Nano Letters
6(4):633-639. DOI: 10.1021/nl052254u.
Su, Z., J. D. Hostert, and J. N. Renner. 2020. Phosphate recovery by a surface-immobilized cerium
affinity peptide. ACS ES&T Water 1(1):58-67. DOI: 10.1021/acsestwater.0c00001.
Sun, T., A. Dasgupta, Z. Zhao, M. Nurunnabi, and S. Mitragotri. 2020. Physical triggering
strategies for drug delivery. Advanced Drug Delivery Reviews 158:36-62. DOI: 10.1016/
j.addr.2020.06.010.
Swartz, J. R. 2012. Transforming biochemical engineering with cell-free biology. American
Institute of Chemical Engineers Journal 58(1):5-13. DOI: 10.1002/aic.13701.
Takata, T., J. Jiang, Y. Sakata, M. Nakabayashi, N. Shibata, V. Nandal, K. Seki, T. Hisatomi, and
K. Domen. 2020. Photocatalytic water splitting with a quantum efficiency of almost unity.
Nature 581(7809):411-414. DOI: 10.1038/s41586-020-2278-9.
Takatori, S. C., and J. F. Brady. 2015. Towards a thermodynamics of active matter. Physical
Review E 91(3). DOI: 10.1103/PhysRevE.91.032117.
Takeuchi, I., J. Lauterbach, and M. J. Fasolka. 2005. Combinatorial materials synthesis. Materials
Today 8(10):18-26. DOI: 10.1016/S1369-7021(05)71121-4.
Talmadge, M. S., R. M. Baldwin, M. J. Biddy, R. L. McCormick, G. T. Beckham, G. A. Ferguson,
S. Czernik, K. A. Magrini-Bair, T. D. Foust, P. D. Metelski, C. Hetrick, and M. R. Nimlos.
2014. A perspective on oxygenated species in the refinery integration of pyrolysis oil.
Green Chemistry 16(2):407-453. DOI: 10.1039/c3gc41951g.
Talvitie, J., A. Mikola, A. Koistinen, and O. Setala. 2017. Solutions to microplastic pollution—
Removal of microplastics from wastewater effluent with advanced wastewater treatment
technologies. Water Research 123:401-407. DOI: 10.1016/j.watres.2017.07.005.
Tan, D. H. S., A. Banerjee, Z. Chen, and Y. S. Meng. 2020a. From nanoscale interface
characterization to sustainable energy storage using all-solid-state batteries. Nature
Nanotechnology 15(3):170-180. DOI: 10.1038/s41565-020-0657-x.
Tan, Y., J. Shen, T. Si, C. L. Ho, Y. Li, and L. Dai. 2020b. Engineered live biotherapeutics:
Progress and challenges. Biotechnology Journal 15(10). DOI: 10.1002/biot.202000155.
TARSC (Training and Research Support Centre). 2015. Innovations for health: Use of appropriate
technologies in primary health care in Zimbabwe—Report of an assessment. Retrieved
August 18, 2021, from https://www.tarsc.org/publications/documents/AppTech%20PHC
%20Zim%20rep%20April2015.pdf.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

References 301

Taylor, G. I. 1934. The formation of emulsions in definable fields of flow. Proceedings of the
Royal Society of London. Series A, Containing Papers of a Mathematical and Physical
Character 146(858):501-523. DOI: 10.1098/rspa.1934.0169.
Taylor, R., and R. Krishna. 2000. Modelling reactive distillation. Chemical Engineering Science
55(22):5183-5229. DOI: 10.1016/s0009-2509(00)00120-2.
Teesalu, T., K. N. Sugahara, and E. Ruoslahti. 2012. Mapping of vascular ZIP codes by phage
display. Methods in Enzymology 503:35-56. DOI: 10.1016/B978-0-12-396962-0.00002-
1.
Tesar, J. E. 1996. The Macmillan visual almanac. B. S. Glassman, ed. London: Macmillan General
Reference.
Teuten, E. L., J. M. Saquing, D. R. U. Knappe, M. A. Barlaz, S. Jonsson, A. Björn, S. J. Rowland,
R. C. Thompson, T. S. Galloway, R. Yamashita, D. Ochi, Y. Watanuki, C. Moore, P. H.
Viet, T. S. Tana, M. Prudente, R. Boonyatumanond, M. P. Zakaria, K. Akkhavong, Y.
Ogata, H. Hirai, S. Iwasa, K. Mizukawa, Y. Hagino, A. Imamura, M. Saha, and H.
Takada. 2009. Transport and release of chemicals from plastics to the environment and to
wildlife. Philosophical Transactions of the Royal Society B: Biological Sciences
364(1526):2027-2045. DOI: doi:10.1098/rstb.2008.0284.
The Pew Charitable Trusts. 2019. Two decades of change in federal and state higher education
funding. Retrieved September 8, 2021, from https://www.pewtrusts.org/en/research-and-
analysis/issue-briefs/2019/10/two-decades-of-change-in-federal-and-state-higher-educa
tion-funding.
The Royal Society. 2018a. Options for producing low-carbon hydrogen at scale. Retrieved August
17, 2021, from https://royalsociety.org/~/media/policy/projects/hydrogen-production/
energy-briefing-green-hydrogen.pdf.
The Royal Society. 2018b. Greenhouse gas removal. Retrieved August 18, 2021, from
https://royalsociety.org/-/media/policy/projects/greenhouse-gas-removal/royal-society-
greenhouse-gas-removal-report-2018.pdf.
Thollander, P., M. Karlsson, P. Rohdin, W. Johan, and J. Rosenqvist. 2020. Introduction to
industrial energy efficiency. Cambridge, MA: Academic Press.
Thoman, D. B., E. R. Brown, A. Z. Mason, A. G. Harmsen, and J. L. Smith. 2015. The role of
altruistic values in motivating underrepresented minority students for biomedicine.
BioScience 65(2):183-188. DOI: 10.1093/biosci/biu199.
Thomassen, G., M. Van Dael, S. Van Passel, and F. You. 2019. How to assess the potential of
emerging green technologies? Towards a prospective environmental and techno-
economic assessment framework. Green Chemistry 21(18):4868-4886. DOI: 10.1039/
C9GC02223F.
Tian, Y., S. E. Demirel, M. M. F. Hasan, and E. N. Pistikopoulos. 2018. An overview of process
systems engineering approaches for process intensification: State of the art. Chemical
Engineering and Processing—Process Intensification 133:160-210. DOI: 10.1016/j.cep.
2018.07.014.
Tian, X., S. D. Stranks, and F. You. 2020a. Life cycle energy use and environmental implications
of high-performance perovskite tandem solar cells. Science Advances 6(31). DOI:
10.1126/sciadv.abb0055.
Tian, Y., J. R. Lhermitte, L. Bai, T. Vo, H. L. Xin, H. Li, R. Li, M. Fukuto, K. G. Yager, J. S.
Kahn, Y. Xiong, B. Minevich, S. K. Kumar, and O. Gang. 2020b. Ordered three-
dimensional nanomaterials using DNA-prescribed and valence-controlled material
voxels. Nature Materials 19(7):789-796. DOI: 10.1038/s41563-019-0550-x.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

302 New Directions for Chemical Engineering

Tilman, D., C. Balzer, J. Hill, and B. L. Befort. 2011. Global food demand and the sustainable
intensification of agriculture. Proceedings of the National Academy of Sciences
108(50):20260-20264. DOI: 10.1073/pnas.1116437108.
Tolle, K. M., D. S. W. Tansley, and A. J. G. Hey. 2011. The fourth paradigm: Data-intensive
scientific discovery. Proceedings of the IEEE 99(8):1334-1337. DOI: 10.1109/JPROC.
2011.2155130.
Tong, Y., I. Y. Zhang, and R. K. Campen. 2018. Experimentally quantifying anion polarizability
at the air/water interface. Nature Communications 9. DOI: 10.1038/s41467-018-03598-
x.
Tremblay, J.-F. 2018. Golden times for electronic materials suppliers. Chemical & Engineering
News 96(28). https://cen.acs.org/materials/electronic-materials/Golden-times-electronic-
materials-suppliers/96/i28.
Tsurushita, N., P. R. Hinton, and S. Kumar. 2005. Design of humanized antibodies: From anti-Tac
to Zenapax. Methods 36(1):69-83. DOI: 10.1016/j.ymeth.2005.01.007.
Tuck, C. O., E. Perez, I. T. Horvath, R. A. Sheldon, and M. Poliakoff. 2012. Valorization of
biomass: Deriving more value from waste. Science 337(6095):695-699. DOI:
10.1126/science.1218930.
Tumbleston, J. R., D. Shirvanyants, N. Ermoshkin, R. Janusziewicz, A. R. Johnson, D. Kelly, K.
Chen, R. Pinschmidt, J. P. Rolland, A. Ermoshkin, E. T. Samulski, and J. M. DeSimone.
2015. Additive manufacturing: Continuous liquid interface production of 3D objects.
Science 347(6228):1349-1352. DOI: 10.1126/science.aaa2397.
Tunuguntla, R. H., R. Y. Henley, Y. C. Yao, T. A. Pham, M. Wanunu, and A. Noy. 2017. Enhanced
water permeability and tunable ion selectivity in subnanometer carbon nanotube porins.
Science 357(6353):792-796. DOI: 10.1126/science.aan2438.
Tursi, A., T. F. Mastropietro, R. Bruno, M. Baratta, J. Ferrando-Soria, A. I. Mashin, F. P. Nicoletta,
E. Pardo, G. De Filpo, and D. Armentano. 2021. Synthesis and enhanced capture proper-
ties of a new BioMOF@SWCNT-BP: Recovery of the endangered rare-earth elements
from aqueous systems. Advanced Materials Interfaces 8(16). DOI: 10.1002/admi.
202100730.
Turton, R., R. C. Bailie, W. B. Whiting, J. A. Shaeiwitz, and D. Bhattacharyya. 2012. Analysis,
synthesis, and design of chemical processes. Upper Saddle River, NJ: Prentice Hall.
Uekert, T., M. A. Bajada, T. Schubert, C. M. Pichler, and E. Reisner. 2020. Scalable photocatalyst
panels for photoreforming of plastic, biomass and mixed waste in flow. ChemSusChem
14(19):4190-4197. DOI: 10.1002/cssc.202002580.
UN (United Nations). 2017. World population projected to reach 9.8 billion in 2050, and 11.2
billion in 2100. Retrieved August 16, 2021, from https://www.un.org/development/
desa/en/news/population/world-population-prospects-2017.html.
UNEP (United Nations Environment Programme). 2020. Handbook for the Montreal protocol on
substances that deplete the ozone: 14th edition. Nairobi, Kenya: Ozone Secretariat.
United Nations Environment Programme. 2021. Our planet is drowning in plastic pollution—It’s
time for change! https://www.unep.org/interactive/beat-plastic-pollution.
UNFCC (United Nations Framework Convention on Climate Change). 2015. Adoption of the Paris
Agreement. 21st Conference of the Parties. Paris, United Nations. https://unfccc.int/sites/
default/files/english_paris_agreement.pdf.
University of Hertfordshire. 2021. PPDB: Pesticide properties database. Retrieved August 27,
2021, from http://sitem.herts.ac.uk/aeru/ppdb.
University of Minnesota. 2003. Professor L. E. (Skip) Scriven. Retrieved August 17, 2021, from
http://www.chemeng.ntua.gr/dep/boudouvis/U%20of%20M%20CEMS-scriven.htm.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

References 303

U.S. Census Bureau. 2018. An aging nation: Projected number of children and older adults.
Retrieved March 18, 2021, from https://www.census.gov/library/visualizations/2018/
comm/historic-first.html.
U.S. Census Bureau. 2019a. Percent of population 25 Years and over by detailed attainment level:
2000-2019. Retrieved March 18, 2021, 2021, from https://www.census.gov/content/
dam/Census/library/visualizations/time-series/demo/fig11.png.
U.S. Census Bureau. 2019b. Percent of population age 25 and over by educational attainment.
Retrieved March 18, 2021, 2021, from https://www.census.gov/content/dam/Census/
library/visualizations/time-series/demo/fig2.png.
USPTO (U.S. Patent and Trademark Office) 2021. U.S. colleges and universities—Utility patent
grants 1969-2012. Retrieved September 8, 2021, from https://www.uspto.gov/web/
offices/ac/ido/oeip/taf/univ/doc/doc_info_2012.htm.
van Anders, G., D. Klotsa, A. S. Karas, P. M. Dodd, and S. C. Glotzer. 2015. Digital alchemy for
materials design: Colloids and beyond. ACS Nano 9(10):9542-9553. DOI: 10.1021/
acsnano.5b04181.
Vargason, A. M., A. C. Anselmo, and S. Mitragotri. 2021. The evolution of commercial drug
delivery technologies. Nature Biomedical Engineering 5:951-967. DOI: 10.1038/s41551-
021-00698-w.
Vaucher, A. C., F. Zipoli, J. Geluykens, V. H. Nair, P. Schwaller, and T. Laino. 2020. Automated
extraction of chemical synthesis actions from experimental procedures. Nature
Communications 11(1):3601. DOI: 10.1038/s41467-020-17266-6.
Vazquez, M. 2018. Engaging biomedical engineering in health disparities challenges. Journal of
Community Medicine and Health Education 8(2). DOI: 10.4172/2161-0711.1000595.
Veers, P., K. Dykes, E. Lantz, S. Barth, C. L. Bottasso, O. Carlson, A. Clifton, J. Green, P. Green,
H. Holttinen, D. Laird, V. Lehtomäki, J. K. Lundquist, J. Manwell, M. Marquis, C.
Meneveau, P. Moriarty, X. Munduate, M. Muskulus, J. Naughton, L. Pao, J. Paquette, J.
Peinke, A. Robertson, J. Sanz Rodrigo, A. M. Sempreviva, J. C. Smith, A. Tuohy, and R.
Wiser. 2019. Grand challenges in the science of wind energy. Science 366(6464). DOI:
10.1126/science.aau2027.
Venkatasubramanian, V. 2019. The promise of artificial intelligence in chemical engineering: Is it
here, finally? AIChE Journal 65(2):466-478. DOI: 10.1002/aic.16489.
Verhougstraete, M. P., J. K. Gerald, C. P. Gerba, and K. A. Reynolds. 2019. Cost-benefit of point-
of-use devices for lead reduction. Environmental Research 171:260-265. DOI:
10.1016/j.envres.2019.01.016.
Villagomez-Salas, S., P. Manikandan, S. F. Acuna Guzman, and V. G. Pol. 2018. Amorphous
carbon chips li-ion battery anodes produced through polyethylene waste upcycling. ACS
Omega 3(12):17520-17527. DOI: 10.1021/acsomega.8b02290.
Voiland, A., and R. Simmon. 2010. Aerosols: Tiny particles, big impact. Retrieved August 16,
2021, from https://earthobservatory.nasa.gov/features/Aerosols.
Wahman, D. G., M. D. Pinelli, M. R. Schock, and D. A. Lytle. 2021. Theoretical equilibrium
lead(II) solubility revisited: Open source code and practical relationships. AWWA Water
Science 3(5). DOI: 10.1002/aws2.1250.
Walker, S., and R. Rothman. 2020. Life cycle assessment of bio-based and fossil-based plastic: A
review. Journal of Cleaner Production 261. DOI: 10.1016/j.jclepro.2020.121158.
Walker, T. W., N. Frelka, Z. Shen, A. K. Chew, J. Banick, S. Grey, M. S. Kim, J. A. Dumesic, R.
C. Van Lehn, and G. W. Huber. 2020. Recycling of multilayer plastic packaging materials
by solvent-targeted recovery and precipitation. Science Advances 6(47). DOI:
10.1126/sciadv.aba7599.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

304 New Directions for Chemical Engineering

Wang, R. E., S. A. Wu, J. A. Evans, J. B. Tenenbaum, D. C. Parkes, and M. Kleiman-Weiner.


2020. Too many cooks: Bayesian inference for coordinating multi-agent collaboration.
arXiv. DOI: arXiv:2003.11778.
Wang, L., D. Rehman, P.-F. Sun, A. Deshmukh, L. Zhang, Q. Han, Z. Yang, Z. Wang, H.-D. Park,
J. H. Lienhard, and C. Y. Tang. 2021a. Novel positively charged metal-coordinated nan-
ofiltration membrane for lithium recovery. ACS Applied Materials & Interfaces
13(14):16906-16915. DOI: 10.1021/acsami.1c02252.
Wang, L. L., M. E. Janes, N. Kumbhojkar, N. Kapate, J. R. Clegg, S. Prakash, M. K. Heavey, Z.
Zhao, A. C. Anselmo, and S. Mitragotri. 2021b. Cell therapies in the clinic.
Bioengineering & Translational Medicine 6(2). DOI: 10.1002/btm2.10214.
Wang, T., Y. Wu, L. Shi, X. Hu, M. Chen, and L. Wu. 2021c. A structural polymer for highly
efficient all-day passive radiative cooling. Nature Communications 12(1). DOI: 10.1038/
s41467-020-20646-7.
Waring, M. J., J. Arrowsmith, A. R. Leach, P. D. Leeson, S. Mandrell, R. M. Owen, G. Pairaudeau,
W. D. Pennie, S. D. Pickett, J. Wang, O. Wallace, and A. Weir. 2015. An analysis of the
attrition of drug candidates from four major pharmaceutical companies. Nature Reviews
Drug Discovery 14(7):475-486. DOI: 10.1038/nrd4609.
Warnock, S. J., R. Sujanani, E. S. Zofchak, S. Zhao, T. J. Dilenschneider, K. G. Hanson, S.
Mukherjee, V. Ganesan, B. D. Freeman, M. M. Abu-Omar, and C. M. Bates. 2021. Engi-
neering Li/Na selectivity in 12-Crown-4–functionalized polymer membranes. Proceed-
ings of the National Academy of Sciences 118(37). DOI: 10.1073/pnas.2022197118.
Washington State Department of Ecology. 2021. Monitoring Hanford’s groundwater and
protecting the Columbia River. Retrieved August 17, 2021, from https://ecology.wa.gov/
Waste-Toxics/Nuclear-waste/Hanford-cleanup/Protecting-air-water/Groundwater-moni
toring.
Wei, J. 1991. Centennial symposium of chemical engineering: Opening remarks. In Advances in
chemical engineering. Colton, C. K., ed. San Diego, CA: Academic Press.
Weisbrod, A., A. Bjork, D. McLaughlin, T. Federle, K. McDonough, J. Malcolm, and R. Cina.
2016. Framework for evaluating sustainably sourced renewable materials. Supply Chain
Forum: An International Journal 17(4):259-272. DOI: 10.1080/16258312.2016.
1258895.
Wells, E., and A. S. Robinson. 2017. Cellular engineering for therapeutic protein production:
Product quality, host modification, and process improvement. Biotechnology Journal
12(1). DOI: 10.1002/biot.201600105.
Welp, K., A. Cartolano, D. Parrillo, R. Boehme, R. Machado, and S. Caram. 2006. Monolith
catalytic reactor coupled to static mixer. U.S. Patent No. 7109378. Washington, DC: U.S.
Patent and Trademark Office.
Welp, K., A. Cartolano, D. Parrillo, R. Boehme, R. Machado, and S. Caram 2009. Monolith
catalytic reactor coupled to static mixer. U.S. Patent No. 7595029. Washington, DC: U.S.
Patent and Trademark Office.
Westmoreland, P. R., and C. McCabe. 2018. Revisiting the future of chemical engineering. CEP
Magazine. https://www.aiche.org/resources/publications/cep/2018/october/revisiting-fu
ture-chemical-engineering.
Whitehead, T. A., S. Banta, W. E. Bentley, M. J. Betenbaugh, C. Chan, D. S. Clark, C. A. Hoesli,
M. C. Jewett, B. Junker, M. Koffas, R. Kshirsagar, A. Lewis, C. T. Li, C. Maranas, E.
Terry Papoutsakis, K. L. J. Prather, S. Schaffer, L. Segatori, and I. Wheeldon. 2020. The
importance and future of biochemical engineering. Biotechnology and Bioengineering
117(8):2305-2318. DOI: 10.1002/bit.27364.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

References 305

Whiteside, A. 2019. Language-based software’s accurate predictions translate to benefits for


chemists. Chemistry World News. https://www.chemistryworld.com/news/language-
based-softwares-accurate-predictions-translate-to-benefits-for-chemists/4010437.article.
WHO (World Health Organization). 2016. Ambient air pollution: A global assessment of exposure
and burden of disease. Geneva: WHO. https://apps.who.int/iris/rest/bitstreams/1061179/
retrieve.
WHO. 2018. Ambient (outdoor) air pollution. Geneva: WHO. https://www.who.
int/news-room/fact-sheets/detail/ambient-(outdoor)-air-quality-and-health.
Wilkerson, C. G., S. D. Mansfield, F. Lu, S. Withers, J.-Y. Park, S. D. Karlen, E. Gonzales-Vigil,
D. Padmakshan, F. Unda, J. Rencoret, and J. Ralph. 2014. Monolignol ferulate transferase
introduces chemically labile linkages into the lignin backbone. Science 344(6179):90-93.
DOI: 10.1126/science.1250161.
Wilkes, J. 2002. A century of chemical engineering at the University of Michigan. Ann Arbor, MI:
University of Michigan.
Willson, V. L., T. Monogue, and C. Malave. 1995. First year comparative evaluation of the Texas
A&M freshman integrated engineering program. Proceedings Frontiers in Education
1995 25th Annual Conference. Engineering Education for the 21st Century. DOI:
10.1109/FIE.1995.483114.
Winter, E., M. C. Clark, and C. Burns. 2021. Team-based learning brings academic rigor,
collaboration, and community to online learning. In Resilient pedagogy: Practical
teaching strategies to overcome distance, disruption, and distraction. Thurston, T. N., K.
Lundstrom and C. Gonzalesz, eds. Logan: Utah State University.
Wischmeyer, P. E., D. McDonald, and R. Knight. 2016. Role of the microbiome, probiotics, and
“dysbiosis therapy” in critical illness. Current Opinion in Critical Care 22(4):347-353.
DOI: 10.1097/MCC.0000000000000321.
Wittrup, K. D. 2001. Protein engineering by cell-surface display. Current Opinion in
Biotechnology 12(4):395-399. DOI: 10.1016/s0958-1669(00)00233-0.
WNA (World Nuclear Association). 2021. Nuclear power in the USA. Retrieved August 17, 2021,
from https://www.world-nuclear.org/information-library/country-profiles/countries-t-
z/usa-nuclear-power.aspx.
Wolkowicz, K. L., E. M. Aiello, E. Vargas, H. Teymourian, F. Tehrani, J. Wang, J. E. Pinsker, F.
J. Doyle, 3rd, M. E. Patti, L. M. Laffel, and E. Dassau. 2021. A review of biomarkers in
the context of type 1 diabetes: Biological sensing for enhanced glucose control.
Bioengineering & Translational Medicine 6(2). DOI: 10.1002/btm2.10201.
World Economic Forum. 2011. Water security: The water-food-energy-climate nexus.
Washington, DC: Island Press.
Wu, F., J. Maier, and Y. Yu. 2020. Guidelines and trends for next-generation rechargeable lithium
and lithium-ion batteries. Chemical Society Reviews 49(5):1569-1614. DOI:
10.1039/c7cs00863e.
Wu, L., D. Wang, and J. A. Evans. 2019. Large teams develop and small teams disrupt science and
technology. Nature 566(7744):378-382. DOI: 10.1038/s41586-019-0941-9.
Xin Yu, J., V. M. Hubbard-Lucey, and J. Tang. 2019. Immuno-oncology drug development goes
global. Nature Reviews Drug Discovery 18(12):899-900. DOI: 10.1038/d41573-019-
00167-9.
Xu, Z., P. Lei, R. Zhai, Z. Wen, and M. Jin. 2019. Recent advances in lignin valorization with
bacterial cultures: microorganisms, metabolic pathways, and bio-products. Biotechnology
for Biofuels 12. DOI: 10.1186/s13068-019-1376-0.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

306 New Directions for Chemical Engineering

Yan, Q., and J. J. de Pablo. 1999. Hyper-parallel tempering Monte Carlo: Application to the
Lennard-Jones fluid and the restricted primitive model. The Journal of Chemical Physics
111(21):9509-9516. DOI: 10.1063/1.480282.
Yang, H. C., Y. Xie, J. Hou, A. K. Cheetham, V. Chen, and S. B. Darling. 2018. Janus membranes:
Creating asymmetry for energy efficiency. Advanced Materials 30(43). DOI:
10.1002/adma.201801495.
Yang, M., N. R. Baral, B. A. Simmons, J. C. Mortimer, P. M. Shih, and C. D. Scown. 2020.
Accumulation of high-value bioproducts in planta can improve the economics of
advanced biofuels. Proceedings of the National Academy of Sciences 117(15):8639-8648.
DOI: 10.1073/pnas.2000053117.
Ye, R.-P., J. Ding, W. Gong, M. D. Argyle, Q. Zhong, Y. Wang, C. K. Russell, Z. Xu, A. G.
Russell, Q. Li, M. Fan, and Y.-G. Yao. 2019. CO2 hydrogenation to high-value products
via heterogeneous catalysis. Nature Communications 10(1). DOI: 10.1038/s41467-019-
13638-9.
Yates, F. E.; Urquhart, J.1962. Control of plasma concentrations of adrenocortical hormones.
Physiological Reviews 42, 359−443. DOI: 10.1152/physrev.1962.42.3.359.
Yeh, Y.-C., T.-H. Huang, S.-C. Yang, C.-C. Chen, and J.-Y. Fang. 2020. Nano-based drug delivery
or targeting to eradicate bacteria for infection mitigation: A review of recent advances.
Frontiers in Chemistry 8(286). DOI: 10.3389/fchem.2020.00286.
Yin, Y., Y. Wang, J. A. Evans, and D. Wang. 2019. Quantifying the dynamics of failure across
science, startups and security. Nature 575(7781):190-194. DOI: 10.1038/s41586-019-
1725-y.
Zakzeski, J., P. C. Bruijnincx, A. L. Jongerius, and B. M. Weckhuysen. 2010. The catalytic
valorization of lignin for the production of renewable chemicals. Chemical Reviews
110(6):3552-3599. DOI: 10.1021/cr900354u.
Zero Waste Scotland. 2013. Plastics to oil products—Final report. Retrieved August 18, 2021,
from https://www.zerowastescotland.org.uk/research-evidence/plastic-oil-report.
Zhai, Y., Y. Ma, S. N. David, D. Zhao, R. Lou, G. Tan, R. Yang, and X. Yin. 2017. Scalable-
manufactured randomized glass-polymer hybrid metamaterial for daytime radiative
cooling. Science 355(6329):1062-1066. DOI: 10.1126/science.aai7899.
Zhang, Z., X. Sui, P. Li, G. Xie, X. Y. Kong, K. Xiao, L. Gao, L. Wen, and L. Jiang. 2017. Ultrathin
and ion-selective Janus membranes for high-performance osmotic energy conversion.
Journal of the American Chemical Society 139(26):8905-8914. DOI: 10.1021/
jacs.7b02794.
Zhang, M., P. T. Corona, N. Ruocco, D. Alvarez, P. Malo de Molina, S. Mitragotri, and M. E.
Helgeson. 2018a. Controlling complex nanoemulsion morphology using asymmetric
cosurfactants for the preparation of polymer nanocapsules. Langmuir 34(3):978-990.
DOI: 10.1021/acs.langmuir.7b02843.
Zhang, X., M. Fevre, G. O. Jones, and R. M. Waymouth. 2018b. Catalysis as an enabling science
for sustainable polymers. Chemical Reviews 118(2):839-885. DOI: 10.1021/acs.chem
rev.7b00329.
Zhang, C., D. Hu, R. Yang, and Z. Liu. 2020a. Effect of sodium alginate on phosphorus recovery
by vivianite precipitation. Journal of Environmental Sciences (China) 93:164-169. DOI:
10.1016/j.jes.2020.04.007.
Zhang, F., M. Zeng, R. D. Yappert, J. Sun, Y.-H. Lee, A. M. LaPointe, B. Peters, M. M. Abu-
Omar, and S. L. Scott. 2020b. Polyethylene upcycling to long-chain alkylaromatics by
tandem hydrogenolysis/aromatization. Science 370(6515):437-441. DOI: 10.1126/
science.abc5441.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

References 307

Zhao, Y. B., X. D. Lv, and H. G. Ni. 2018. Solvent-based separation and recycling of waste
plastics: A review. Chemosphere 209:707-720. DOI: 10.1016/j.chemosphere.2018.06.
095.
Zhu, X., J. Hao, B. Bao, Y. Zhou, H. Zhang, J. Pang, Z. Jiang, and L. Jiang. 2018. Unique ion
rectification in hypersaline environment: A high-performance and sustainable power
generator system. Science Advances 4(10). DOI: 10.1126/sciadv.aau1665.
Zimmermann, A. W., and R. Schomäcker. 2017. Assessing early-stage CO2 utilization
technologies—Comparing apples and oranges? Energy Technology 5(6):850-860. DOI:
10.1002/ente.201600805.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Appendix A
List of Acronyms

AA acrylic acid
ACS American Chemical Society
AGORA assembly of gut organisms through reconstruction and analysis
AI artificial intelligence
AIChE American Institute of Chemical Engineers
AM additive manufacturing
API active pharmaceutical ingredient
API American Petroleum Institute
ASEE American Society for Engineering Education
ASSB all-solid-state batteries
AWEA American Wind Energy Association

BAU business as usual


BEV battery electric vehicle
BF blast furnace
BIO Biotechnology Innovation Organization
BIPOC Black, Indigenous, and People of Color
BLS Bureau of Labor Statistics
BTX benzene, toluene, xylene

CAR chimeric antigen receptor


CCUS carbon capture, use, and storage (or carbon capture, utilization, and
sequestration)
CFC chlorofluorocarbon
CHO Chinese hamster ovary
CHP combined heat and power
CMS U.S. Centers for Medicare & Medicaid Services
CNN convolutional neural network
COBRA constraints-based reconstruction and analysis
CPU central processing unit
CRISPR clustered regularly interspaced short palindromic repeats

DAC direct air capture


DBER discipline-based education research
DMF N,N-dimethylformamide
DNA deoxyribonucleic acid
DNN deep neural networks
DOE U.S. Department of Energy

308

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Appendix A 309

DRI direct reduction of iron

ECM extracellular matrix


ED electrodialysis
EG ethylene glycol
EMF Energy Modeling Forum
EOR enhanced oil recovery
EPA U.S. Environmental Protection Agency
EV electric vehicle

Fab fragment antigen-binding


FAO Food and Agriculture Organization of the United Nations
Fc fragment crystallizable
FCC fluid catalytic cracking
FCEV fuel cell electric vehicle
FDA U.S. Food and Drug Administration
FDCA furan-2,5-dicarboxylic acid
FMT fecal matter transplant
FORTRAN Formula Translation
FOSSI Future of STEM Scholars Initiative

GAN generative adversarial networks


GDP gross domestic product
GE General Electric
GEM genome-scale metabolic model
GGE gallon gasoline equivalent
GHG greenhouse gas
GI gastrointestinal
GM genetically modified
GPU graphics processing unit

HCFC hydrochlorofluorocarbon
HDA high-density amorphous
HIV human immunodeficiency viruses
HMP Human Microbiome Project
HT high temperature

IAM integrated assessment management


IBD inflammatory bowel disease
IC integrated circuit
ICE internal combustion engine
IDM independent device manufacturer
IEA International Energy Agency
ILI Institute for Learning and Innovation

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

310 New Directions for Chemical Engineering

IoT Internet of Things


IR infrared
ISO International Standards Organization
IUPAC International Union of Pure and Applied Chemistry

LAB linear alkyl benzene


LCA life-cycle assessment
LDA low-density amorphous
LNG liquefied natural gas
LPG liquefied propane gas
LPN liquid nanoparticle

mAb monoclonal antibody


MD molecular dynamics
MEA membrane electrode assembly
MEG monoethylene glycol
MF microfiltration
MIT Massachusetts Institute of Technology
ML machine learning
MPI multiple principal investigators
mRNA messenger ribonucleic acid
MS mass spectrometry
MSW municipal solid waste
MW molecular weight

NAE National Academy of Engineering


NASEM National Academies of Sciences, Engineering, and Medicine
NET negative emissions technologies
NF nanofiltration
NG natural gas
NGL natural gas liquids
NHL N-acyl-L-homoserine lactones
NIH National Institutes of Health
NMR nuclear magnetic resonance
NN neural networks
NOAA U.S. National Oceanic and Atmospheric Administration
NOM natural organic matter
NRC National Research Council
NREL National Renewable Energy Laboratory
NSF National Science Foundation

OC organic carbon
OCM oxidative coupling of methane
ORNL Oak Ridge National Laboratory

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Appendix A 311

PAH polycyclic aromatic hydrocarbon


PAM polyacrylamide
PBAT polybutyrate adipate terephthalate
PBS polybutylene succinate
PCB polychlorinated biphenyl
PCL polycaprolactone
PCR polymerase chain reaction
PDRC passive daytime radiative cooling
PE polyethylene
PEC photoelectrochemical cells
PEF polyethylene furanoate
PEG polyethylene glycol
PEM proton exchange membrane
PET polyethylene terephthalate
PGMEA propylene glycol methyl ether acetate
PHA polyhydroxy alkanoate
PI process intensification
PLA polylactic acid
PM particulate matter
PMMA polymethylmethacrylate
PNAS Proceedings of the National Academy of Sciences
PP polypropylene
PS polystyrene
PSC perovskite solar cells
PTA purified terephthalic acid
PTT polytrimethylene terephthalate
PUR polyurethane
PV photovoltaic
PVC polyvinyl chloride
PVOH polyvinyl alcohol

QSAR quantitative structure–property relationships


QSPR quantitative structure–activity relationships

R&D research and development


RD&D research, development, and demonstration
REU Research Experience for Undergraduates
RNA ribonucleic acid
RNN recurrent neural network
RO reverse osmosis

scFv single-chain variable fragment


SCM supplementary cementitious materials

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

312 New Directions for Chemical Engineering

SCR selective catalytic reduction


SEM strategic energy management
SFG sum frequency generation
siRNA small interfering RNA
SOEC solid oxide electrolysis cell
SMARTS SMILES arbitrary target specification
SMILES simplified molecular-input line-entry system
SPI single principal investigator
STEM science, technology, engineering, and math

TEA technoeconomic assessment


THF tetrahydrofuran
TRISO TRi-structural ISOtropic
TRL technology readiness level

UF ultrafiltration
UN United Nations
USD United States dollar

WEF water–energy–food
WHO World Health Organization
WHP waste heat to power
WtE waste-to-energy

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Appendix B
Journals Used in International Benchmarking

The following list of journals was compiled for the analysis of publications by
chemical engineers discussed in Chapter 10. The original list, minus journals that have
gone out of print, was taken from Appendix 3B of the National Academies report
International Benchmarking of U.S. Chemical Engineering Research Competitiveness
(NRC, 2007). Journals that began publication or that have risen in prominence since 2007
were added by the committee and are italicized in the list below.

313

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

314 New Directions for Chemical Engineering

ACS Applied Energy Materials Bioinformatics


ACS Catalysis Biomacromolecules
ACS Central Science Biomaterials
ACS Omega Biomicrofluidics
ACS Sensors Biomimetics
ACS Sustainable Chemistry & BioNanoScience
Engineering Bioprocess and Biosystems
Acta Materialia Engineering
Advanced Drug Delivery Reviews BioResources
Advanced Functional Materials Biotechnology and Bioengineering
Advanced Materials Biotechnology and Genetic Engineering
Advanced Science Reviews
Advanced Structural and Chemical Biotechnology Progress
Imaging Canadian Journal of Chemical
Advances in Nano Research Engineering
Aerosol Science and Technology Carbon
AIChE Journal Carbon Letters
Analytical Methods Case Studies in Thermal Engineering
Angewandte Chemie Catalysis Science and Technology
Annals of Biomedical Engineering Catalysis Today
Annals of Operations Research Catalysis, Structure and Reactivity
Annual Review of Biophysics Catalysts
Annual Review of Chemical and Chem
Biomolecular Engineering ChemCatChem
Applied and Environmental ChemElectroChem
Microbiology Chemical Engineering and Processing -
Applied Bionics and Biomechanics Process Intensification
Applied Catalysis A: General Chemical Engineering and Technology
Applied Catalysis B: Environmental Chemical Engineering Research and
Applied Sciences (Switzerland) Design
Arabian Journal of Chemistry Chemical Engineering Science
Atmospheric Chemistry and Physics Chemie-Ingenieur-Technik
Atmospheric Environment Chemistry of Materials
Automatica Chemosphere
Avicenna Journal of Medical ChemSusChem
Biotechnology Cogent Engineering
Biocatalysis and Agricultural Colloids and Interface Science
Biotechnology Communications
Biochip Journal Colloids and Surfaces A:
Bioengineered Physicochemical and Engineering
Bioengineering Aspects
Biofabrication Colloids and Surfaces B: Biointerfaces
Biofuel Research Journal Combustion and Flame
Biofuels, Bioproducts and Biorefining Combustion Science and Technology

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Appendix B 315

Composite Structures Ground Water


Composites Science and Technology Indonesian Journal of Science and
Computational Optimization and Technology
Applications Industrial & Engineering Chemistry
Computational Particle Mechanics Research
Computers and Chemical Engineering INFORMS Journal on Computing
Cosmetics Inorganic Chemistry
Crystals Inorganic Materials
Current Opinion in Biomedical Interface Focus
Engineering International Journal of Air-
Current Opinion in Green and Conditioning and Refrigeration
Sustainable Chemistry International Journal of Chemical
CYTA - Journal of Food Kinetics
Dose-Response International Journal of Corrosion
Ecological Economics International Journal of Industrial
Egyptian Journal of Petroleum Chemistry
Electrochimica Acta International Journal of Multiphase
Energy & Fuels Flow
Engineering International Review of Aerospace
Engineering in Agriculture, Engineering
Environment and Food International Review on Modelling and
Engineering Science and Technology, Simulations
an International Journal Iranian Journal of Catalysis
Environmental Progress and Johnson Matthey Technology Review
Sustainable Energy Journal of Aerosol Science
Environmental Science & Technology Journal of Agricultural Engineering
Environmental Toxicology and Journal of Analytical Methods in
Chemistry Chemistry
Enzyme and Microbial Technology Journal of Applied Biomaterials and
European Journal of Pharmaceutical Functional Materials
Sciences Journal of Applied Electrochemistry
Express Polymer Letters Journal of Biomaterials Science,
Extreme Mechanics Letters Polymer Edition
Fluid Phase Equilibria Journal of Biomedical Materials
Food and Bioprocess Technology Research
Food Structure Journal of Biomedical Nanotechnology
Frontiers in Bioengineering and Journal of Catalysis
Biotechnology Journal of Chemical Physics
Frontiers of Chemical Science and Journal of Chemical Thermodynamics
Engineering Journal of CO2 Utilization
Fuel Journal of Colloid and Interface
Fuel Processing Technology Science
Granular Matter Journal of Combustion
Green Processing and Synthesis Journal of Contaminant Hydrology

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

316 New Directions for Chemical Engineering

Journal of Controlled Release Journal of the European Ceramic


Journal of Diabetes Science and Society
Technology Journal of the Taiwan Institute of
Journal of Engineering Chemical Engineers
Journal of Environmental Chemical Journal of Thermal Science and
Engineering Engineering Applications
Journal of Flow Chemistry Journal of Vacuum Science and
Journal of Fluid Mechanics Technology B: Nanotechnology and
Journal of Food Measurement and Microelectronics
Characterization Journal of Water Process Engineering
Journal of Geophysical Research Journal of Water Reuse and
Journal of Global Optimization Desalination
Journal of Hazardous, Toxic, and KONA Powder and Particle Journal
Radioactive Waste Langmuir
Journal of Materials Research Macromolecular Reaction Engineering
Journal of Materials Science Macromolecules
Journal of Materials Science: Materials Materials Horizons
in Medicine Materials Research Bulletin
Journal of Membrane Science Materials Today Chemistry
Journal of Membrane Science and Mathematical Programming, Series B
Research Membranes
Journal of Nanofluids Metabolic Engineering
Journal of Nanoparticle Research Microbial Biotechnology
Journal of Non-Newtonian Fluid Molecular Catalysis
Mechanics Molecular Simulation
Journal of Optimization Theory and Molecular Systems Design and
Applications Engineering
Journal of Orthopaedic Research Nano Futures
Journal of Physical Chemistry B Nano Letters
Journal of Polymer Engineering Nanomaterials
Journal of Polymer Science, Part A: Nanotechnology for Environmental
Polymer Chemistry Engineering
Journal of Polymer Science, Part B: Nanotechnology Reviews
Polymer Physics Nanotechnology, Science and
Journal of Power Sources Applications
Journal of Process Control Nature
Journal of Rheology Nature Biomedical Engineering
Journal of the Air and Waste Nature Biotechnology
Management Association Nature Catalysis
Journal of the American Ceramic Nature Chemistry
Society Nature Materials
Journal of the American Chemical Nature Reviews Chemistry
Society New Biotechnology
Journal of the Electrochemical Society Nonlinear Engineering

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Appendix B 317

Optimization and Engineering SPE Journal


Particuology Studies in Surface Science and
Pharmaceutical Research Catalysis
Physical Review Fluids Surface and Interface Analysis
Physical Review Letters Surface Innovations
Physicochemical Problems of Mineral Surface Topography: Metrology and
Processing Properties
Polymer Synthetic Biology
Polymer Chemistry Tellus, Series B: Chemical and Physical
Polymer Composites Meteorology
Polymer Engineering and Science Thermal Science and Engineering
Polymer-Plastics Technology and Progress
Materials Tissue Engineering
Powder Technology Tissue Engineering - Part A
Proceedings of the Combustion Institute Tissue Engineering - Part B: Reviews
Proceedings of the National Academy Tissue Engineering - Part C: Methods
of Sciences of the United States of Toxics
America Vodohospodarsky Casopis/Journal of
Process Biochemistry Hydrology & Hydromechanics
Process Safety Progress Water Research
Progress in Energy and Combustion Water Resources Research
Science Wiley Interdisciplinary Reviews:
Progress in Polymer Science Nanomedicine and
Propulsion and Power Research Nanobiotechnology
Protein Science
Reaction Chemistry and Engineering
Reaction Kinetics, Mechanisms and
Catalysis
Rheologica Acta
RSC Advances
Safety and Health at Work
Science
Science and Technology for the Built
Environment
Separation and Purification Reviews
Separation and Purification Technology
Separation Science and Technology
Separations
SIAM Journal of Scientific Computing
SIAM Journal on Optimization
Solar Energy Materials and Solar Cells
Solid State Ionics
South African Journal of Chemical
Engineering

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Appendix C
Summary of Results of the Chemical Engineering
Community Questionnaire

The committee originally planned to hold workshops and other meetings in con-
junction with the annual meetings of relevant professional societies to solicit input for this
study from the broader chemical engineering community. Unfortunately, the COVID-19
pandemic precluded such in-person gatherings. As an alternative, in spring 2021 the com-
mittee distributed a questionnaire to members of the chemical engineering community to
gather broad input on challenges and opportunities for the discipline, as well as key needs
in education and training.
The web link to the online questionnaire was distributed via email to subscribers
to the mailing list for the National Academies’ Board on Chemical Sciences and Technol-
ogy (BCST). The questionnaire was open to anyone who wished to provide input. Addi-
tionally, members of the committee shared the invitation with their own professional net-
works. There were 43 complete responses and 249 partial responses. All questions were
optional, including basic demographic information, which was collected only to help en-
sure an appropriate range of perspectives.
This appendix summarizes the input provided. Some responses have been edited
or condensed for clarity. Note that the ideas and suggestions summarized here are those
of the anonymous respondents to the online questionnaire and do not represent the views
of the committee or the National Academies.

1. Across all of chemical engineering, what three fields of chemical engineering will
be the most intellectually exciting/promising in the next decade?

 Energy and sustainable/renewable  Electrochemistry, energy storage,


energy – 15 batteries – 4
 Process automation/control  Materials
/design/safety/data analytics – 8 (engineering/processing/computat
 Environmental ional design) – 4
sustainability/engineering – 7  Agriculture (artificial
 Pharmaceuticals, therapeutics, photosynthesis, safe food) – 3
personalized medicine, health  Bioprocess and process
care – 7 engineering – 3
 Data science/AI – 6  Biotechnology – 3
 Biomedical engineering  Catalysis – 3
applications – 4  Food process engineering – 3
 Decarbonization (energy,  Green chemistry – 3
transportation, etc.) – 4  Hydrogen (synthesis, use, storage)
–3

318

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Appendix C 319

 Bio/Biochemical engineering – 2
 Biomolecular engineering – 2
 Environmentally friendly
materials (i.e., biodegradable end
products) – 2
 Modeling/Simulations – 2
 Multiscale modeling in biological
systems – 2
 Nanotechnology – 2
 Plastics alternatives – 2
 Carbon, capture, utilization,
storage – 1
 Chemical technology – 1
 Cosmetics – 1
 Educational technology, training
–1
 Entropy vs enthalpy-controlled
systems – 1
 Global engineering – 1
 Green materials – 1
 Manufacturing of complex, non-
Newtonian fluids - 1
 Membrane science – 1
 Reaction engineering - 1
 Recycling plastics and critical
materials – 1
 Research and development – 1
 Water resources management – 1

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

320 New Directions for Chemical Engineering

2. Thinking about the next 10 to 30 years, what three fields of chemical engineering
will have the greatest impact on emerging technologies, national needs, and/or the
wider science and engineering enterprise?

 Energy/Renewable energy (its  Health care/testing for public


availability & processes) – 11 health – 2
 Green/environmental  Rheology studies – 2
engineering & sustainability –  Automation – 1
8  Biochemistry/Biochemical
 Catalysis – 6 engineering – 1
 Electrochemistry/Batteries/  Bio-defense – 1
Electronic materials – 5  Bioprocess engineering – 1
 Materials  Carbon capture – 1
science/processing/discovery/  Chemical technology – 1
engineering – 5  Cost-effective biodegradable
 Biotechnology/Biopharmaceuti end products – 1
cals – 4  Food waste/loss – 1
 Circular economy/LCA  Hydrogen economy – 1
(recycling plastics/products) –  Manufacturing – 1
4  Multiscale simulation – 1
 Decarbonization – 4  Nanotechnology – 1
 Pharmaceuticals (i.e., additive  Novel sensing – 1
manufacturing, biopharma) – 4
 Nuclear – 1
 Biomolecular/Bioengineering
 Plastics alternatives – 1
–3
 Process energy optimization –
 Food security/waste – 3
1
 Process control/data analytics/
 Research and development – 1
development and scale-up – 3
 Separations – 1
 Transport phenomena – 3
 Systems integration – 1
 Biomedical engineering – 2
 Clean water resources – 2
 Computing & AI (in all facets)
–2

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Appendix C 321

3. Within your personal field of research or professional focus, what are the major
goals for the next 10 to 30 years, and what are the major barriers to getting there?

Goal(s) Barrier(s)
Decarbonization & Reducing GHG
Emissions

Decarbonization of global economy by 2050 Lack of political consensus—because of this,


other countries will develop winning
Massive reductions in GHG emissions with technologies at scale
politically acceptable impacts of standard of
living Adoption of alternative technology

In the last decades, chemical processes were


based on petrochemistry. Not only fuels, but
also most of the chemical commodities. To
replace this fossil carbon by renewable
carbon (biomass, CO2) is still a major
challenge. Although in the last 15–20 years
a big effort was paid to these processes, they
are far from being competitive with
petrochemical. Similar situation applies to
hydrogen energy.

50% reduction in greenhouse gases by 2030,


net zero by 2050

Low cost and scalable technology for


carbon capture and storage, point source and
direct air; Novel tech for low emission fuels
for hard to decarbonize sectors, biofuels and
hydrogen; Scalable and affordable negative
emission biomass based technologies

Sustainability, carbon capture, circular


economy

Circular Economy

We are targeting to engage in circular The major barrier is the know-how to engage
economy. in it.

Sustainability and next-gen manufacturing Lack of public understanding/appreciation for


tech what goes into developing products

Sustainability Adoption of alternative technology

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

322 New Directions for Chemical Engineering

Sustainability, carbon capture, circular Cost effectiveness, matching


economy performance/purity of recycled and materials

Plastics alternatives or recycling

Improve agricultural practices

Development of energy storage systems to Fragmented research across many disciplines


make renewable energy more economically that don’t speak the same language
and practically feasible

Engineering of molecules that can Proper mechanistic understanding of the


interrogate and modulate physiological system and the ability to design molecules to
environments with greater precision and achieve the precise goal without off-target
sensitivity than current diagnostics and effects.
therapeutics.

Biopharmaceutical Processes

Move biopharmaceutical process Individuals who control the funding are trial-
development from being largely trial-and- and-error experimentalists and have a vested
error experimentation to becoming a interest in maximizing the research funding to
systematic technology based on mechanistic their own approach and activities
models, data analytics, and process control
Powerful vested interests want to continue to
Biopharmaceutical manufacturing processes have the research funding go overwhelmingly
that have the potential to replace the current into the existing established processes rather
processes while having major increases in than to competing processes. Another barrier is
quality, development time, and/or cost a strong resistance to any new technology

Connecting industry to academia in a robust, Things that have kept the groups apart over the
respectful, and collaborative manner; years—including elitism, skeptical colleagues,
connecting the various engineering and seemingly separate goals.
disciplines to act on common problems that
require consensus and convergence thinking

Robust engineering of cells for therapeutic Understanding/defining principles for


delivery and tissue repair/regeneration engineering cells for robust transgene
expression and control
(i) Mathematical modeling of phase changes Major barriers for both are the proper problem
in flowing soft-matter systems (ii) formulation and computational resources
Mathematical modeling of particle-laden
interfaces

Digitization and automation of industry

Industrial waste water recycling Cost effectiveness and disposal options of


reject streams

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Appendix C 323

Breaking silos and creating partnerships to


address grand challenges

Lost-cost biodegradable products

Materials de novo synthesis, analysis, then Shareholder short-term optimization in


delivery to systems-level in context of industry, see-sawing values for federal
specific applications research grants

Fundamental understanding of Absence of adequate tools for atomic-level


physicochemical properties of catalysts and images of catalyst structure and composition
electrocatalysts relationship with activity under working conditions, multiscale
and selectivity simulation of performance of electrochemical
systems

Sustaining and expanding manufacturing Competing with off-shore sites that do not
capability in the United States have the same labor and environmental
requirements and more government subsidies
and indirect government involvement

Innovative discoveries representing future Movement away from curiosity-driven


directions of technology research and tightening of funding with a focus
on “deliverables”

Make oil and gas space more attractive to Make it greener with suitable applications for
consumers CO2 emissions

Development based on green chemistry new


processes that account their environmental
impact

Good teaching schools for young chemical


engineers

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

324 New Directions for Chemical Engineering

4. Again, thinking about your personal field of research or professional focus, what
is the societal relevance of your work, and what are the barriers to translation
and/or scale-up?

Societal relevance Barrier


Decarbonization, GHG Emission Reduction,
Reducing Climate Change

Decarbonization of the global economy by Lack of political consensus


2050
Huge fossil fuel infrastructure makes it hard
Zero-carbon technology to sustain humanity; to scale renewable energy practically and
energy storage economically

Meeting the growing energy needs for the Science-based, technology-neutral policy that
world as economies prosper while mitigating incentivizes all relevant technologies to meet
environmental impact including emissions the dual challenge; ecosystems that promote
partnerships and collaborations; skills and
Mitigating the impacts of climate change competencies for novel process development
and scale-up

Suitable regulatory policies

Improve Human Health & Affordable Health


Care
Physiological complexity
Human health (antibiotic-resistant infectious
disease, cancer, cardiovascular disease, etc.) Need to understand and develop cheap,
simple quality control to ensure quality
Well-controlled cell-based therapeutics could production of cells that are specific to patients
reshape how we treat disease to make it scalable

Improved biopharmaceutical manufacturing Acquisition of funding. Biopharma doesn’t


would result in higher quality products at want to spend money to translate a
lower cost and shorter time to market – more technology that will help competitors, new
promising drugs make it to market tech draws resources from existing tech,
incentivizing deemphasis on new tech.
Cancer drugs, drugs for COVID, COVID Academic reviewers also have vested interest
vaccine in preventing translation of tech of other
researchers

Time for the heavily regulated industry to


accept new manufacturing methods that will
meet regulatory scrutiny and general risk
aversion; continuous manufacturing is being
implemented in biologics drug substance
manufacturing

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Appendix C 325

Sustainability/Circular Economy

It will create employment and make our Availability of funds


planet cleaner (circular economy)
Cost effectiveness, performance of the
Environmental sustainability alternatives or recycled options and trade-off
with energy

Connecting diverse groups of STEM learners


and practitioners to utilize all parts of our
communities

Political partisanship. Technoeconomic


challenges can be solved, politics cannot

Materials processing is necessary to turn Chemical engineering research has shifted


materials into useful products away from process-related questions toward
chemical-/molecular-level details. Need
fundamental research in materials processing,
such as fluid mechanics and heat/mass
transport. This needs industrial interaction,
should be encouraged through establishment
and generous support of programs like NSF
GOALI

Catalytic upgrading of bioplatform molecules Processes are not competitive, but regulations
are pulling to phase out use of oil and even
natural gas

Food engineering and sustainable


technologies—innovated food technologies
geared toward maternal and child health for
local and global development supporting
UNSDGs 2&17

Reducing water use across industries Unit ops and processes in water are 100 years
or older with little innovation. Need to
rethink priorities at national education level to
revitalize critical thinking and research in
water use and unit operations/processes

Ensure a better future Changing long established practices and


adopting new approaches and technologies

Safe, efficient, environmentally sound


petrochemicals production

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

326 New Directions for Chemical Engineering

Mitigate the human damage to the Free market capitalism and engineering
ecosystems triumphalism

Electrocatalysts will play an increasingly


important role for fuel cells and electrolyzers
for water-splitting to generate hydrogen and
reducing CO2. Developing means for the
efficient capture of CO2 and its conversion to
chemicals holds immense opportunity

Manufacturing provides more GDP impact Need government to actively maintain a level
and better salaries overall compared to playing field and use tariffs, etc., to enforce it,
service and other industries. Necessary for rather than lowering standards
national security and not rely on outside
sources excessively for strategic materials
and capabilities

Economic, social, food, and water systems Barriers are political


are going to be disrupted by climate change if
goals are not met

Isolation of industrial scientist and engineers


reduces the connection between academic
knowledge and advances from industry
knowledge and practice; industrial scientists
are not able to participate in meaningful
research projects and meetings, creating
among other things, echo chamber for
academics

Plastics manufacturing, creating low-cost


products that will be applied to food
preservation and health security

Oil and gas space is getting a lot of negative More consumer friendly public relation
publicity, but it provides a cheap and reliable pointers so that society as a whole realizes the
source of energy importance of oil and gas energy space. Also
challenge to mitigate CO2 emissions and find
suitable applications/conversion technologies
for CO2

To improve the processes and making them New mentality


more friendly toward the environment.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Appendix C 327

5. What are some of the key ideas/principles/drivers that have rotated out over the
last 10 years in your field of research or professional focus? In other words, as new
capabilities emerge in the field, which areas are making way for the new concepts?

Theme Individual Responses


AI/Machine learning/Data/ Traditional experimental phase equilibrium and property
Computational tools measurement is nearly nonexistent in academia or industry. New
AI and machine learning tools need data for training—where are
students going to learn it?

Developing biopharmaceutical processes based only on trial-


and-error experimentation is being rotated out in the last 10
years in industry, as new capabilities emerge in process data
analytics and machine learning, mechanistical modeling, and
process control. Academia with rare exceptions have been slow
to make way for these approaches which are not new in all of
chemical engineering but have become increasingly important in
biopharmaceutical process development as practiced in the
leading biopharma companies and equipment vendors. The
changes are happening in industry, whether trial-and-error
experimentalists like it or not. Either academia can lead and
contribute to these developments and contribute to society,
translation, and scale-up, or they can keep gripping tighter and
tighter on the past while harming U.S. competitiveness directly
and indirectly by not producing the trained people needed by
these changes in the industry.

Computational tools are revolutionizing our understanding of


gene regulation

Manufacturing process, introduction of technology in industry

Circular economy Material balance, recycling, the chemistry for producing certain
materials

Reliance on recycling is replacing only use of raw materials

The production of commodity fuels from hydrocarbon sources is


fading and will become, at best, like paper production today.
Sustainability and life-cycle metrics will drive innovation, and
transportation fuels from HC fails these tests. Similarly with
current plastics production. Circular polymers with minimal
waste is the new concept

Recycling waste material (lithium batteries, plastic, municipal


raw material, agricultural waste material)

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

328 New Directions for Chemical Engineering

Traditional Traditional process engineering has suffered in the quest for


processes/ChemE federal R&D funding at the “leading edge.” It has materially
fundamentals are being harmed our national capability to produce competent engineers
lost—Losing depth in the at the B.S. level for industries that exist today.
field
None. The fundamentals (applied mathematics, fluid mechanics,
heat and mass transport, thermodynamics, and reaction
engineering) are all still very relevant.

I studied polymers in graduate school and ended up in


biotechnology. Our graduate programs are teaching people how
to become hyper experts in a field. We live in an
interdisciplinary world, and we have to build skillsets in diverse
areas. Traditional chemical engineering core competencies must
be strengthened, and certainly not abandoned. However, instead
of forcing undergrads to take two semesters of organic
chemistry, how about encouraging them to take a course in
biochemistry and another course in biochemical engineering
instead? Teach chemical engineers the science underlying
carbon capture and clean energy technologies. The chip shortage
is a reminder that there is room for innovation and beefing up
the supply chain.

The field has shifted away from fundamentals and analysis and
more toward empirical “gee-whiz” results. We are losing depth.

Bioengineering is rotating Biochemical engineering, biomedical engineering are two key


out; technology is taking areas that have rotated out of the way.
over
The field of nanotechnology, green-chemistry, and bio-
molecular engineering would be some of them. Some of the
more industry-oriented technologies, such as the divided-wall-
column, do show green technology potential in the near future.

Other Benefits and limitations of remote education and remote


teamwork.

No need for all the irrelevant/redundant chemistry classes that


are in the canonical curriculum.

Computational/informatic/theoretical approaches are being


hybridized with experimental approaches, both high-
throughput/midfidelity and low-throughput/high fidelity.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Appendix C 329

In my opinion, one of the key problems observed nowadays is


the very different speeds of scientific development (very fast)
and the aspects related to the scaling up and process engineering
(traditionally a core of chemical engineering). We observe the
fast development of new and very active catalysts for different
processes (OER, HER, hydrogenation, photocatalysts, single-
atom catalysts, MOFs), but there are very few attempts of
scaling up.

UN Sustainable Development Goals Hidden Hunger


(micronutrient deficiencies); public–private partnership to
address community needs.

In cosmetics where I am working currently, trend is toward skin-


friendly organic products.

Six Sigma and Lean are over-blown concepts that have become
a distraction and their own industry.

The idea that climate science is questionable is rotating out, as


climate change becomes more obvious.

Heavy oil conversion process and catalyst development;


hydroprocessing and hydroconversion catalyst and materials for
conventional fuels; conventional petrochemical process R&D.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

330 New Directions for Chemical Engineering

6. What are the five most important areas of technical knowledge for chemical
engineers to learn during an undergraduate degree?

 Thermodynamics – 16  Chemistry – 3
 Computing, Machine Learning,  Fact-based Analysis – 3
Statistics, Data Science – 12  Science of next generation
 Mass Transfer/Mass and clean energy generation and
Energy Balances – 9 technologies – 3
 Transport Phenomena – 9  Catalysis – 2
 Kinetics – 8  Chemical Reactor Engineering
 Reaction Engineering – 8 –2
 Applied Mathematics – 7  Green Chemistry/Technologies
 Economics (Manufacturing and –2
Scale-up) – 7  Heat & Material Balances – 2
 Process Design/Engineering/  Nanotechnology – 2
Simulation – 7  Biochemistry & Biochemical
 Process Control – 6 Engineering – 1
 Fluid Mechanics/Dynamics – 5  Entrepreneurship – 1
 Unit Operation Principles – 5  Food & Nutrition Security – 1
 Biology/Earth  Humanities – 1
Sciences/Geology – 4  Leadership – 1
 Circular  Manufacturing – 1
Economy/Sustainability – 4  Momentum Transfer – 1
 Heat Transfer – 4  R&D – 1
 Systems Thinking/Engineering  Reach Methodology – 1
(i.e. Connecting engineering –  Physics – 1
techno-economic-social-  Separations – 1
cultural-geographical models  Synthesis of new materials) – 1
for process flow ) – 4  Vaccine Development – 1

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Appendix C 331

7. What are the five most important areas of technical knowledge for chemical
engineers to learn during a graduate degree?

 Numerical Methods/Statistics/  Self-Awareness, Emotional


Mathematical Modeling – 12 Intelligence, Conflict
 Thermodynamics – 7 Resolution – 2
 Advanced Transport/Transport  Systems Thinking (i.e.
Phenomena – 6 connecting engineering-
 Modeling & Simulation – 5 techno-economic-social-
 Circular Economy & cultural-geographical models
Sustainability (Hydrogen for process flow) – 2
Economy, Decarbonization,  Time Management &
etc.) – 4 Prioritization/ Project
 Reaction Engineering – 4 Management – 2
 Heat & Mass Transfer – 3  Water Resources
 Kinetics – 3 Management/Treatment – 2
 Process Engineering/Systems/  Active Materials and Low
Design – 3 Energy Separations – 1
 Research Methodology &  Advanced Manufacturing – 1
Experimental Design – 3  Advanced Chemical Synthesis
 Specialized areas relevant to –1
Thesis – 3  Biodegradable Science – 1
 Advanced Reactors/Reactor  Chemistry – 1
Design – 2  Communicating difficult
 AI/Computer Science – 2 concepts to a wide audience – 1
 Applied Mathematics – 2  Convergence Technologies – 1
 Biology – 2  Digitalization – 1
 Biomolecular  Drug Development – 1
Engineering/Bioengineering – 2  Environmental Impact
 Bio and Biomedical Assessment – 1
Applications (Vaccine  Fluid Mechanics – 1
Development) – 2  Food and Nutrition Security – 1
 Catalysis – 2  Heat and Material Balances – 1
 Data Science/Big Data  Industry Standards – 1
Analytics – 2  Instrumentation, Control,
 Economics/Cost Estimation – 2 Digital Signals and Control – 1
 Good Writing Skills/Technical  Leadership – 1
Writing – 2  Materials Science – 1
 Green Processes/Technologies  Model Discrimination &
–2 Parameter Estimation – 1
 Nanotechnology – 2  Momentum Transfer – 1
 Renewable Energy – 2  Operations Research – 1
 Problem Solving – 1

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

332 New Directions for Chemical Engineering

 Process Innovation and Scale-


up – 1
 Process Intensification – 1
 Physics – 1
 Rheology (non-Newtonian) – 1
 Thermofluids – 1
 Unit Operations – 1
 Unit Operations in Mars – 1

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Appendix C 333

8. Based on your own experience or those of your recent hires, what skills are
missing from chemical engineering undergraduate and/or graduate education
(please include skills that are important, but you/your employees have needed to
learn outside of a standard chemical engineering curriculum)?

 Computing, Data Science and Analytics, Statistics – 9


 Effective Writing & Communication Skills – 9
 Humanities & Social Sciences (all students benefit from these courses, benefits
to interdisciplinary learning) – 3
 Creativity – 2
 Economics – 2
 Holistic Process Development (Need students to be good at lab,
experimentation/simulation, and modeling) – 2
 Understanding phenomena and their relationship to complex processes – 2
 Advanced Math (Linear Algebra) – 1
 Analytical Thinking – 1
 Basic General Knowledge of Experimental Work – 1
 Biology from an engineering perspective – 1
 Climate Change, Sustainability & Circularity – 1
 Entrepreneurship – 1
 Fundamental Chemistry – 1
 Green Chemistry – 1
 Hands-on Experience – 1
 Independence – 1
 Industrial Exposure/Industry Standards – 1
 Leadership – 1
 Operation Research Skills – 1
 Overall Familiarity with Instrumentation and Control Systems – 1
 Particle Technology and Solids Handling – 1
 Practical Experience – 1
 Process Control/Process Dynamics/Process Simulation – 1
 Project Flowsheet – 1
 Societal Impacts of Technology – 1
 Transport Phenomena – 1

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

334 New Directions for Chemical Engineering

9. In what way(s) might interdisciplinary or emerging topics (for example biology,


sustainability, data science, polymers, nanomaterials, etc.) be better integrated into
chemical engineering undergraduate and/or graduate education?

 Circular Economy/Sustainability – 8
 Data Science, AI, Robotics – 6
 Incorporate broader array of real-world problems within the ChE core – 5
 Community Engagement and Practical Experience (Project, Internships, Field
Trips) – 3
 Interdisciplinary topics are useful as a way of synthesizing knowledge from core
classes, but should not be a substitute for acquiring fundamental numerical
skills that are only learned in college – 3
 Polymer Chemistry – 3
 Biology/Biological Engineering – 2
 Collaborating with and Exposure to Industry – 2
 Integrate emerging topics into core curriculum rather than creating specialized
ones – 2
 Nanotechnology – 2
 New Requirement: ChEs to take classes outside of the ChE Department – 2
 Applied ChE – 1
 Applied Statistics – 1
 Establish Forums at Universities to Encourage Entrepreneurship Around
Innovative Technologies – 1
 Focused seminars/courses (grad) – 1
 Geosciences – 1
 Guest lectures (undergrad) – 1
 Project-based work with other Engineers – 1
 Quantum Computing – 1
 Separate modules (grad) – 1
 Transversal issues can be introduced in classical undergrad modules (undergrad)
–1

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Appendix C 335

10. What are the biggest barriers to increasing the diversity of the chemical
engineering workforce?

 Current pool of undergrads is not very diverse; Diversity in STEM needs to start
at a younger age (i.e., high school) – 5
 Perception/Marketing issue—positive aspects (financial benefits, sustainability)
of ChE not emphasized enough and negative perceptions (too hard or boring,
not forward looking enough) of ChE need to be squashed – 5
 Lack of resources to attract and retain high-quality people (including
scholarships) – 3
 Negative American political ethos (anti-immigration policy, prejudice, bigotry,
discrimination, sexism, homophobia, racism) – 3
 Not enough mentors or role models to seek guidance – 3
 Develop STEM programs that support URM throughout college – 2
 Better involvement of women in STEM – 1
 Declining higher education enrollments – 1
 Increase opportunities for a broader community – 1
 Lack of diversity within faculty/educational institutions are the barrier – 1
 Lack of incentives for encouraging diversity – 1
 Lack of outreach to URM students – 1
 Lack of time to invest in this activity – 1
 Limited job and growth opportunities in conventional ChE fields – 1
 Myth of Meritocracy – 1
 Narrow-minded thinking – 1
 Other fields such as health care, medicine, bioengineering, computer and data
science, finance, etc., are deemed to be more exciting than conventional ChE – 1
 URM aren’t attracted to ChE in part because of its lack of diversity – 1

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

336 New Directions for Chemical Engineering

11. How does the U.S. remain competitive and at the cutting edge in chemical
engineering? And how do we establish and maintain important international
collaborations?

 Attract more foreign students to study in the U.S. (and reforming U.S.
immigration policy) – 5
 Modify ChE education (i.e., need more innovation in undergrad curriculum
(ABET should be radically modified); we need to reconnect academia with the
actual practice of engineering; need to increase STEM education at a younger
age) – 5
 International collaborations should be with countries that are committed to our
same values of democracy, intellectual property, and openness – 3
 AIChE growing a global network of student chapters – 2
 ChE not prioritized in the U.S.; funding agencies do not provide incentives for
U.S. faculty to work on unresolved ChE problems – 2
 Diversity of opinion and looking at the problems from multiple perspectives are
key to innovative solutions – 2
 Identify the wide breadth of opportunities where chemical engineers can be a
driving force; expose the community (at all career stages) to these opportunities
–2
 Systems-thinking approach coupled with a quest for innovation; make ChE
more interdisciplinary – 2
 To establish and maintain important international collaborations, parties should
ensure mutual benefit by protecting intellectual property – 2
 U.S. universities would need to increase its hiring/teaching in chemical
engineering topical areas of importance to growing high-tech industries (topical
areas include particle technology, biopharmaceutical manufacturing, advanced
industrial polymers, energy challenges and lead in discovery, development and
deployment of new breakthrough technologies for low-carbon future, etc.) – 2
 By engaging with different entities in our professional societies, research
endeavors and transparent interactions in the international/global realm – 1
 Establish ecosystems to further partnership and collaboration with academia,
government, and industry to accelerate advancement of new technologies – 1
 Incentivize domestic students to pursue ChE – 1
 International collaborations are less important, build expertise and facilities here
in the U.S. – 1
 Maintenance of an interactive academia and industry focus; along with support
of industrial development into the future – 1
 More global collaboration to share best practices and ideas – 1
 Processing of advanced materials has not received nearly the amount of focus it
should relative to its importance; this problem could be addressed by providing
more research support for the fundamentals that play a key role in materials
processing, such as fluid mechanics and heat/mass transport – 1

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Appendix C 337

 U.S. universities should be encouraged to set up their campuses in other


countries, and offer the same enriched engineering curriculum experience and
quality of education as it would on their main campuses in the mainland USA – 1

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

338 New Directions for Chemical Engineering

12. Please use this box for any additional input related to the committee’s charge or
the questions above.

 The focus needs to capture nonresearch elements.


 Chemical engineering should include basic general knowledge like who is the
president/prime minister of their country and US, UK, China, etc. Should have
their field training during their course of work. Sound grasp on computer tools
like Excel, Word, etc. And also chemical engineering or any other course should
have the capacity to generate research skills in students.
 Focus on non-hyped, fact-based analysis.
 The hallmark of chemical engineers is their ability to blend in a seamless manner
knowledge from the fields of chemistry, biology, physics, and mathematics.
 Chemical engineering input is needed to assess proposed solutions to climate
change and other modern challenges, so that solutions that do not make sense
can be ruled out early, rather than wasting time and money on them. An
example is government funding of research to use coal in green hydrogen
processes.
 We focus so much on “identifying a set of ... new chemical engineering areas,”
but what we miss is the ability to discover truly new areas or solutions to grand
challenges. The investments the U.S. is making in science and engineering lag
far behind especially Asia (China, Singapore, Korea) and the investments made
by industry are even worse. How do we reverse this trend? How do we recommit
U.S. enterprises to innovation in the chemical and biological sciences? Perhaps
the richest area today for chemical engineers in terms of new science and
technology is pharmaceutical manufacturing, especially biologics, and “new”
technologies (or finally recognized technologies) like mRNA vaccines.
 I’d like to see AIChE spend more time teaching/mentoring younger students.
 At present, the chemical engineering community is losing its potential chemical
engineers of the future to other branches of science, and engineering, for the
simple fact that these students who are in the high school age group are not
exposed to the wonders of chemical engineering, and they do not have role
models to look up to. For instance, news such as self-driving cars, or
autonomous drones, really gets students excited; however, the same is not made
visible to these students by the chemical engineering field, and industry experts.
Right now, quite a number of students identify the field of chemical engineering
with global warming challenges and CO2 emissions, which is not a fair
assessment.
 Chemical engineering, especially in the leading academic institutions, has
increasingly turned to the microscopic. Process technology has been a
technologically mature field, with incremental advances, and mostly an industry
with a sedate pace of new construction and retrofits. But handling large volumes
of materials cannot remain at the micro scale outside the lab. It requires the
competent design, construction, and operation of process operations.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Appendix D
Acknowledgments

The committee would like to acknowledge the intellectual contributions of the


following individuals:

Noubar Afeyan, Flagship Pioneering Lee Ellen Drechsler, The Procter &
and Moderna Therapeutics Gamble Company
Alina Alexeenko, Purdue University Allessandro Faldi, ExxonMobil
Kristi Anseth, University of Colorado Research and Engineering Company
Boulder Glenn Frederickson, University of
Frances Arnold, California Institute of California, Santa Barbara
Technology Benny Freeman, The University of
Alán Aspuru-Guzik, University of Texas at Austin
Toronto Shishir Gadam, Bristol Myers Squibb
Norman Augustine, Lockheed Martin Salvador Garcia Muñoz, Carnegie
(retired) Mellon University
David Awschalom, University of Dario Gil, IBM
Chicago Rajamani Gounder, Purdue University
William Banholzer, University of Michael Graetzel, École Polytechnique
Wisconsin–Madison Fédérale de Lausanne
Zhenan Bao, Stanford University Ignacio Grossman, Carnegie Mellon
Carlos Barroso, CJB and Associates University
Saad Bhamla, Georgia Institute of Supratik Guha, Argonne National
Technology Laboratory
Donna Blackmond, Scripps Research Frank Gupton, Virginia Commonwealth
Santanu Chaudhuri, Argonne National University
Laboratory Eric Hagemeister, The Procter &
Shannon Ciston, Lawrence Berkeley Gamble Company
National Laboratory, Molecular Nick Halla, Impossible Foods
Foundry William Hammack, University of
Ismaila Dabo, The Pennsylvania State Illinois Urbana-Champaign
University Evelynn Hammonds, Harvard
Cathy Davidson, City University of University
New York Phillip Hustad, The 3M Company
Pablo Debenedetti, Princeton Ah-Hyung (Alissa) Park, Columbia
University University
Joseph DeSimone, Stanford University Robert Johnson, ExxonMobil Research
Francis (Frank) Doyle, Harvard and Engineering Company
University Christopher Jones, Georgia Institute of
Technology

339

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

340 New Directions for Chemical Engineering

Cherie Kagan, University of Jim Pfaendtner, University of


Pennsylvania Washington
Lynn Katz, The University of Texas at Bryan Pivovar, National Renewable
Austin Energy Laboratory
Jay Keasling, University of California, Katie Randolph, U.S. Department of
Berkeley Energy, Office of Energy Efficiency
Ermias Kebreab, University of and Renewable Energy
California, Davis Jeffrey Reimer, University of
Konstantin Konstantinov, Codiak California, Berkeley
Biosciences Gintaris (Rex) Reklaitis, Purdue
Christine Lambert, Ford Research & University
Advanced Engineering William Ristenpart, University of
Dan Lambert, Savannah River National California, Davis
Laboratory James Rogers, Apeel Sciences
Michael Lawson, National Renewable Don Roe, The Procter & Gamble
Energy Laboratory Company
John Layman, The Procter & Gamble Kirsten Sinclair Rosselot, Process
Company Profiles
Kelvin Lee, The National Institute for Tony Ryan, University of Sheffield
Innovation in Manufacturing Aaron Sarafinas, Sarafinas Process &
Biopharmaceuticals Mixing Consulting, LLC
Thomas Lograsso, Ames Laboratory, David Sedlak, University of California,
Critical Materials Institute Berkeley
Lee Lynd, Dartmouth College Jeffrey Selingo, Arizona State
Hang Lu, Georgia Institute of University
Technology Avi Shultz, U.S. Department of Energy,
Julius Lucks, Northwestern University Office of Energy Efficiency and
Gargi Maheshwari, Bristol Myers Renewable Energy
Squibb Justin G. Sink, ExxonMobil Research
Benjamin Maurer, National Renewable and Engineering Company
Energy Laboratory Mark Sivik, The Procter & Gamble
Paul McKenzie, CSL Behring Company
Faye McNeill, Columbia University Henry Snaith, University of Oxford
Mark Meili, The Procter & Gamble Scott Stanley, The Procter & Gamble
Company Company
Carl Mesters, Shell (retired) George Stephanopoulos, Arizona State
Eric Miller, U.S. Department of University and Massachusetts
Energy, Office of Energy Efficiency Institute of Technology
and Renewable Energy Vijay Swarup, ExxonMobil Research
Ahmad Moini, BASF Corporation and Engineering Company
Raul Miranda, U.S. Department of Kazuhiro Takanabe, University of
Energy, Office of Science Tokyo
Lynn Orr, Stanford University Robert Thresher, National Renewable
Energy Laboratory

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Appendix D 341

Jean Tom, Bristol Myers Squibb


Annabelle Watts, The 3M Company
Phillip Westmoreland, North Carolina
State University
Dane Wittrup, Massachusetts Institute
of Technology
Omar Yaghi, University of California,
Berkeley
Yushan Yan, University of Delaware
Aleksey Yezerets, Cummins Inc.
Joe Zasadzinski, University of
Minnesota
Stacey Zones, Chevron Energy
Technology Company

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

342 New Directions for Chemical Engineering

The committee would like to acknowledge the financial contributions of the fol-
lowing organizations:

The American Chemical Society University of Houston


The American Institute of Chemical University of Maryland, Baltimore
Engineers County
Colorado School of Mines University of Michigan
Georgia Institute of Technology University of Minnesota
The Johns Hopkins University University of Notre Dame
Louisiana State University The University of Texas at Austin
Massachusetts Institute of Technology University of Virginia
North Carolina State University University of Wisconsin
Northwestern University West Virginia University
The Pennsylvania State University The American Chemistry Council
Princeton University Arkema
Purdue University Bristol Myers Squibb Company
Rice University The Dow Chemical Company
Texas A&M University DuPont de Nemours, Inc.
University at Buffalo Eastman Chemical Company
University of Arkansas Evonik Industries
University of California, Berkeley ExxonMobil Corporation
University of California, Davis Honeywell International, Inc.
University of California, Los Angeles PPG Industries, Inc.
University of California, Merced The Procter & Gamble Company
University of Delaware Shell Global
University of Florida

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Appendix E
Committee Member and Staff Biographical Sketches

Eric W. Kaler (Chair), NAE, is president of Case Western Reserve University.


Previously, he was president emeritus and professor of chemical engineering and
materials science at the University of Minnesota and served on the faculty of the
Department of Chemical Engineering at the University of Washington. Dr. Kaler also
served as associate professor and dean of the College of Engineering at the University of
Delaware. Dr. Kaler’s research interests are in complex fluids containing surfactants,
polymers, proteins, or colloidal particles, either separately or in mixtures. Additionally,
he studies statistical mechanics and thermodynamics. Dr. Kaler received one of the first
Presidential Young Investigator Awards from the National Science Foundation in 1984.
He has received numerous other awards for his research and is a fellow of several
scientific societies. Dr. Kaler has authored or coauthored more than 200 peer-reviewed
papers and holds 10 U.S. patents. He was elected to the National Academy of Engineering
in 2010 and named a fellow of the American Academy of Arts and Sciences in 2014. Dr.
Kaler earned a PhD in chemical engineering from the University of Minnesota in 1982.

Monty M. Alger, NAE, is professor of chemical engineering at The Pennsylvania State


University. His experience in the chemical and energy industries includes positions as
vice president and chief technology officer at Air Products and Chemicals Inc., and as
senior vice president of research at Myriant. Dr. Alger spent 23 years at General Electric
(GE), where he led technology development at the Global Research Center of GE Plastics
and was general manager of technology for the advanced materials business. Prior to GE,
he was director of the Massachusetts Institute of Technology (MIT) Chemical
Engineering Practice School Station at GE Plastics. Dr. Alger has served on advisory
boards for several universities and organizations, including the Shenhua National Institute
of Clean and Low-Carbon Energy and PTT Global Chemical (Thailand). He is a fellow
and past president of the American Institute of Chemical Engineers. Dr. Alger has an SB
and SM in chemical engineering from MIT and a PhD in chemical engineering from the
University of Illinois at Urbana-Champaign.

Gilda A. Barabino, NAE, NAM, is president of Olin College of Engineering and


professor of biomedical and chemical engineering. She previously served as Daniel and
Frances Berg professor and dean at the City College of New York (CCNY) Grove School
of Engineering. Prior to joining CCNY, Dr. Barabino was associate chair for graduate
studies and professor in the Wallace H. Coulter Department of Biomedical Engineering
at the Georgia Institute of Technology and Emory University. At Georgia Tech, she also
served as inaugural vice provost for academic diversity. Dr. Barabino is a noted
investigator in the areas of sickle cell disease; cellular and tissue engineering; and the role
of race, ethnicity, and gender in science and engineering. She is president-elect of the

343

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

344 New Directions for Chemical Engineering

American Association for the Advancement of Science, the world’s largest


interdisciplinary scientific society. Dr. Barabino is also an active member of the National
Academy of Engineering and the National Academy of Medicine and serves on numerous
committees of the National Academies, including the Roundtable on Black Men and
Black Women in Science, Engineering, and Medicine; the Health and Medicine Division
Committee; and the Committee on Women in Science Engineering and Medicine, which
she chairs. She consults nationally and internationally on STEM (science, technology,
engineering, and mathematics) education and research, diversity in higher education,
policy, and faculty and workforce development. Dr. Barabino received a PhD in chemical
engineering from Rice University.

Gregg T. Beckham is a senior research fellow and group leader at the National
Renewable Energy Laboratory (NREL), where he leads an interdisciplinary team of
biologists, chemists, engineers, and material scientists developing green processes and
products from lignocellulosic biomass and waste plastics. He has published more than
200 peer-reviewed articles. Dr. Beckham was awarded the American Chemical Society
(ACS) OpenEye Outstanding Junior Faculty Award, the American Institute of Chemical
Engineers (AIChE) Computational Science and Engineering Forum Young Investigator
Award, an inaugural ACS Sustainable Chemistry and Engineering Lectureship, the
Society for Industrial Microbiology and Biotechnology (SIMB) Young Investigator
Award, the Royal Society of Chemistry Beilby Medal and Prize, the SIMB Charles D.
Scott Award, the BioEnvironmental Polymer Society Outstanding Young Scientist
Award, and the NREL Innovator of the Year Award. He is also founding cochair of both
the Lignin Gordon Research Conference (2018) and the Plastics Upcycling and Recycling
Gordon Research Conference (2022). He testified before the U.S. House of
Representatives Committee on Science, Space, and Technology’s Subcommittee on
Research and Technology on “Closing the Loop: Emerging Technologies in Plastics
Recycling” and coorganized the National Academies of Sciences, Engineering, and
Medicine Symposium on “Closing the Loop on the Plastics Dilemma,” both in 2019. At
NREL, Dr. Beckham cofounded and leads the U.S. Department of Energy (DOE)–funded
BOTTLE Consortium, which develops advanced plastics recycling and redesign
strategies. He also was a founding member of the Agile BioFoundry and several other
DOE-funded consortia. Before NREL, Dr. Beckham served as director of a Massachusetts
Institute of Technology (MIT) Practice School Station in Singapore. He received an
MSCEP in 2004 and PhD in chemical engineering in 2007 at MIT.

Dimitris I. Collias is research fellow in the Corporate Research and Development


Department of the Procter & Gamble (P&G) Company, leading the development of
technologies in the bio- and circular economy space, such as bioacrylic acid,
biosurfactants alcohols, recycled polyolefins, and recycled superabsorbent polymers,
which are currently in various stages of commercialization. His experience in industrial
materials processing, properties, and delivery systems spans 29 years. Dr. Collias has
experience in technology development and management; has worked extensively with
outside industrial partners, start up companies, and universities; and has seen many of his

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Appendix E 345

technical developments commercialized in various P&G products. He coauthored the


book Polymer Processing—Principles and Design, is coauthoring and coediting a book
on Circular Economy of Polymers—Topics in Recycling Technologies, has published
more than 40 articles in journals and conference proceedings, and holds more than 100
granted U.S. patents and numerous patent applications. Dr. Collias has earned many
awards, such as the 2020 American Chemical Society’s (ACS’s) Affordable Green
Chemistry Award, P&G’s CTO Pathfinder Awards in 2019 and 2010, P&G’s IP Strategy
Award in 2018, Los Alamos National Laboratory’s Recognition and Commendation in
2009, and P&G’s Cost Innovation Award in 2007. He is a member of ACS, the American
Institute of Chemical Engineers, the Society of Rheology, the Society of Plastics
Engineers, and the American Oil Chemists’ Society. Dr. Collias earned a diploma in
chemical engineering from the National Technical University of Athens, Greece, and a
PhD in chemical engineering from Princeton University.

Juan J. de Pablo, NAE, is Liew Family professor and executive vice president for
national laboratories, science strategy, innovation, and global initiatives at the University
of Chicago. He is also a senior scientist at Argonne National Laboratory. Dr. de Pablo is
a leader in developing models and simulations of molecular and large-scale phenomena,
including advanced molecular simulation methods and artificial-intelligence–based
algorithms; he also conducts supercomputer simulations to design and find applications
for new materials, including protein optimization and aggregation, DNA folding and
hybridization, glassy materials, block copolymers, liquid crystals, and active molecular
systems. He holds more than 20 patents and has authored or coauthored approximately
650 publications. Dr. de Pablo received the DuPont Medal for Excellence in Nutrition and
Health Sciences, the Intel Patterning Science Award, the Charles Stine Award from the
American Institute of Chemical Engineers (AIChE), and the Polymer Physics Prize from
the American Physical Society (APS). He is also a member of AIChE and has served as
chair of the awards selection subcommittee. Dr. de Pablo is founding editor of Molecular
Systems Designing and Engineering and deputy editor of Science Advances. He has
served as chair of the Mathematical and Physical Sciences Advisory Committee of the
National Science Foundation and the Committee on Condensed Matter and Materials
Research at the National Academies of Sciences, Engineering, and Medicine, and he is a
fellow of the American Academy of Arts and Sciences and of the APS. Dr. de Pablo was
elected as foreign correspondent member of the Mexican Academy of Sciences in 2014,
and he was elected into the U.S. National Academy of Engineering in 2016. He earned a
PhD in chemical engineering at the University of California, Berkeley.

Sharon C. Glotzer, NAS, NAE, is Anthony C. Lembke Department chair of chemical


engineering and John W. Cahn distinguished university professor at the University of
Michigan. Her research on computational assembly science and engineering aims toward
predictive materials design of colloidal and soft matter. Dr. Glotzer is a fellow of the
American Physical Society (APS), the American Association for the Advancement of
Science, the American Institute of Chemical Engineers (AIChE), and the Materials
Research Society (MRS). She has received numerous awards, including the Nanoscale

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

346 New Directions for Chemical Engineering

Science & Engineering Forum Award, the Alpha Chi Sigma Award, and the Charles M.A.
Stine Award from AIChE; the Kavli Lectureship from the MRS and the MRS Medal; the
Aneesur Rahman Prize in Computational Physics from the APS; and the Presidential
Early Career Award for Scientists and Engineers. Dr. Glotzer has participated in many
activities of the National Academies of Sciences, Engineering, and Medicine, and she
serves currently on the Division on Engineering and Physical Sciences Committee and
served previously on the Board on Chemical Sciences and Technology (2015–2020). She
is also a member of the American Academy of Arts and Sciences. Dr. Glotzer earned a
PhD in physics from Boston University.

Paula Hammond, NAS, NAE, NAM, is instititute professor and department head of the
Chemical Engineering Department at the Massachusetts Institute of Technology (MIT);
she is also a member of MIT’s Koch Institute for Integrative Cancer Research and the
MIT Energy Initiative, and she is a founding member of the MIT Institute for Soldier
Nanotechnologies. She previously served as executive officer (associate chair) of the
Chemical Engineering Department and is associate editor for the journal ACS Nano. Dr.
Hammond’s research is focused on the self-assembly of polymeric nanomaterials,
particularly the use of electrostatics and other complementary interactions to generate
functional materials with highly controlled architectures, including the development of
new biomaterials and electrochemical energy devices. She served on the Board on
Chemical Sciences and Technology of the National Academies of Sciences, Engineering,
and Medicine from 2006 to 2009, and was chair of the American Institute of Chemical
Engineers (AIChE) Materials Science and Engineering Division. Dr. Hammond has been
involved for several years in developing Polymer programming, and in the past has served
as faculty advisor for the AIChE MIT student chapter. She received the William Grimes
Award and the Distinguished Scientist Award from Harvard University, and she is a
fellow of the American Physical Society, the Polymer Division of the American Chemical
Society, and the American Institute of Biological and Medical Engineers. Dr. Hammond
earned a PhD in chemical engineering from MIT.

Enrique Iglesia, NAE, is Theodore Vermeulen chair in chemical engineering at the


University of California, Berkeley, and laboratory fellow at the Pacific Northwest
National Laboratory. His research addresses the synthesis and structural and mechanistic
characterization of porous inorganic catalysts useful in energy conversion, chemical
synthesis, and environmental control. He is a fellow of the American Chemical Society
(ACS) and the American Institute of Chemical Engineers (AIChE), and an honorary
fellow of the Chinese Chemical Society. Dr. Iglesia received a Senior Scientist Award
from the Alexander von Humboldt Foundation and doctors honoris causa degrees from
the Universidad Politecnica de Valencia and the Technical University of Munich. He is a
member of the American Academy of Arts and Sciences, the National Institute of
Inventors, and the Real Academia de Ciencias (Spain). Dr. Iglesia received the Olah,
Somorjai, and Murphree awards from the ACS; the Wilhelm, Alpha Chi Sigma, and
Walker awards from the AIChE; the Emmett, Burwell, Gault, Boudart, and Distinguished
Service awards from the European and North American Catalysis Societies; and the Cross

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Appendix E 347

Canada Lectureship from the Chemical Institute of Canada. He is former editor-in-chief


of The Journal of Catalysis and has served as president of the North American Catalysis
Society and the International Association of Catalysis Societies. His dedication to
teaching has been recognized with several campus awards, most notably the Noyce Prize,
the most prestigious teaching award in the physical sciences at Berkeley. Dr. Iglesia
received a B.S. from Princeton University and a PhD from Stanford University, both in
chemical engineering.

Sangtae Kim, NAE, is distinguished professor of chemical engineering at Purdue


University, where he was also inaugural Donald W. Feddersen distinguished professor of
mechanical engineering. He served previously as executive director, Morgridge Institute
for Research; inaugural division director, National Science Foundation
Cyberinfrastructure Division; and vice president of research and development in
information technology at Eli Lilly and Warner Lambert. Dr. Kim started his career as a
faculty member in chemical engineering at the University of Wisconsin, where he
developed mathematical and computational methods for microhydrodynamics and
coauthored a book on this topic, published in 1991. He is a fellow of the American
Institute of Medical and Biological Engineers and the American Institute of Chemical
Engineers (AIChE), and a trustee of the AIChE Foundation. Dr. Kim received the 2013
Ho-Am Prize in Engineering, AICHE’s George Lappin and Colburn awards, and the 1992
Award for Initiatives in Research from the National Academies of Sciences, Engineering,
and Medicine. He received a PhD in chemical engineering from Princeton University.

Samir Mitragotri, NAE, NAM, is Hiller professor of bioengineering and Hansjorg Wyss
professor of biologically inspired engineering at Harvard University, John A. Paulson
School of Engineering and Applied Sciences. His research is focused on transdermal, oral,
and targeted drug delivery systems. Dr. Mitragotri has made groundbreaking
contributions to the field of biological barriers and drug delivery, advancing fundamental
understanding of biological barriers and enabling the development of new materials and
technologies for diagnosis and treatment of various ailments, including diabetes and
cardiovascular, skin, and infectious diseases. He is an elected fellow of the National
Academy of Inventors, the American Association for the Advancement of Science, the
Biomedical Engineering Society, the American Institute for Medical and Biological
Engineering, and the American Association of Pharmaceutical Scientists. Dr. Mitragotri
is an author of more than 350 publications; an inventor on more than 200 patents or patent
applications; and a recipient of the American Institute of Chemical Engineers’s Colburn
and Acrivos Professional Progress awards, as well as the Society for Biomaterials’s
Clemson award. He received a PhD in chemical engineering from the Massachusetts
Institute of Technology.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

348 New Directions for Chemical Engineering

Babatunde A. Ogunnaike, NAE, was the William L. Friend Chaired professor of


chemical engineering at the University of Delaware, where he began serving after a 13-
year research career with DuPont. His research focused on process control, modeling and
simulation, systems biology, and applied statistics. Dr. Ogunnaike made notable
contributions to the modeling and control of industrial polymer reactors and to
understanding biological control systems. He was the author or coauthor of four books,
including the widely used textbook Process Dynamics, Modeling and Control and
Random Phenomena: Fundamentals of Probability and Statistics for Engineers. Dr.
Ogunnaike received the American Institute of Chemical Engineers’ (AIChE’s) CAST
Computing Practice Award, the University of Delaware’s College of Engineering
Excellence in Teaching Award, the International Society of Automation’s Eckman
Award, and the American Automation and Control Council’s Control Engineering
Practice Award. He was a fellow of the AIChE, the American Association for the
Advancement of Science, the International Federation of Automatic Control, and the
Nigerian Academy of Engineering. Dr. Ogunnaike received a PhD in chemical
engineering from the University of Wisconsin–Madison.

Anne Robinson is trustee professor and department head of chemical engineering at


Carnegie Mellon University. She holds several patents and has authored more than 100
publications in the areas of protein (re)folding and aggregation, protein biophysics, and
protein expression of therapeutically relevant protein molecules. From 2015 to 2017, Dr.
Robinson served on the board of directors of the American Institute of Chemical
Engineers (AIChE); she is on the advisory board of Biotechnology and Bioengineering
and the editorial board of Biotechnology Journal, and has been an ad hoc reviewer for
many National Institutes of Health and National Science Foundation study sections. She
is also a member of the Advisory Committee for Pharmaceutical Sciences of the Food and
Drug Administration. Dr. Robinson received a DuPont Young Professor Award and a
National Science Foundation Presidential Early Career Award for Science and
Engineering, and she is a fellow of the American Institute for Medical and Biological
Engineering and of the AIChE. Dr. Robinson received a PhD in chemical engineering
from the University of Illinois Urbana-Champaign.

José G. Santiesteban, NAE, is recently retired from ExxonMobil, where he served for
more than 30 years in a number of technical leadership and management roles, including,
mostly recently, strategy manager for ExxonMobil Research and Engineering Company.
In this role, he led a team for developing strategic technology direction, providing research
guidance, and ensuring robustness of the research and development portfolio. His
expertise is in heterogeneous catalysis, including design, synthesis, physical chemical
characterization of novel catalytic materials, and reaction mechanisms and kinetics. Dr.
Santiesteban is inventor or coinventor on more than 85 U.S. patents, editor of two special
catalysis journals, and coauthor of more than 20 referenced publications. He has led and
made significant technical contributions to the discovery, development, and
commercialization of more than 20 novel catalyst technologies for the production of high-
performing lubricants, clean fuels, and petrochemicals. Dr. Santiesteban was elected a

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Appendix E 349

member of The Academy of Medicine, Engineering and Science of Texas in 2018. He


received the Society of Hispanic Professional Engineers 2018 Innovator Award and the
“Key to the City” of Parral from Chihuahua in 2016; he also received multiple technical
and leadership awards within ExxonMobil Research and Engineering Company and
Mobil Research and Development Company. Dr. Sanitesteban is a board member of the
Board on Energy and Environmental Systems of the National Academies of Sciences,
Engineering, and Medicine and a senior member of the American Institute of Chemical
Engineers and the North American Catalysis Society. He has served on the advisory board
of various academic and research institutions around the world. Dr. Santiesteban received
a PhD in physical chemistry from Lehigh University.

Rachel A. Segalman, NAE, is Schlinger department chair of chemical engineering and


Edward Noble Kramer professor of materials at the University of California, Santa
Barbara. Previously, she served as acting director at Lawrence Berkeley Laboratories in
the Materials Science Division and as professor of chemical engineering at the University
of California, Berkeley. Dr. Segalman’s research is focused on molecular structure control
over soft matter on molecular through nanoscopic length scales to optimize properties for
applications ranging from energy to biomaterials. She is particularly interested in
materials for energy applications, such as batteries, photovoltaics, fuel cells, and
thermoelectrics. Dr. Segalman received the Dillon Medal from the American Physical
Society (APS), the Presidential Early Career Award in Science and Engineering, and the
Innovation Award from the Journal of Polymer Science. She has been elected to the
American Academy of Arts and Sciences and as a fellow of APS, and is an elected senior
member of the American Institute of Chemical Engineers. Dr. Segalman received a PhD
in chemical engineering from the University of California, Santa Barbara.

David Sholl is director of the Transformational Decarbonization Initiative at the Oak


Ridge National Laboratory. From 2013 to 2021, he was school chair of chemical and
biomolecular engineering at the Georgia Institute of Technology. His research uses
computational materials modeling to accelerate development of new materials for energy-
related applications, including generation and storage of gaseous and liquid fuels and CO2
mitigation. Before his appointment at Georgia Tech, Dr. Sholl served on the faculty of
Carnegie Mellon University for 10 years. He has published more than 360 papers, which
have been cited more than 21,000 times. He has also written a textbook on density
functional theory, a quantum-chemistry method applied widely through the physical
sciences and engineering. Dr. Sholl served as a member on the National Academies of
Sciences, Engineering, and Medicine Committee on a Research Agenda for a New Era in
Separations Science. He was senior editor of the American Chemical Society journal
Langmuir for 10 years, and he was instrumental in the development of the Rapid
Advancement in Process Intensification Deployment (RAPID) Institute, a $70 million,
U.S. Department of Energy–funded manufacturing institute focused on process
intensification run by the American Institute of Chemical Engineers. Dr. Sholl received a
PhD in applied mathematics from the University of Colorado Boulder.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

350 New Directions for Chemical Engineering

Kathleen J. Stebe, NAE, is Goodwin professor in the School of Engineering and Applied
Sciences at the University of Pennsylvania. Her research is focused on directed assembly
in soft matter and at fluid interfaces. After training at the Levich Institute under the
guidance of Charles Maldarelli, she spent a postdoctoral year in Compiegne, France,
under the guidance of Dominique Barthes-Biesel. Thereafter, Dr. Stebe joined the faculty
of the Department of Chemical Engineering at The Johns Hopkins University, and later
joined the Department of Chemical and Biomolecular Engineering at the University of
Pennsylvania. She has been recognized by the American Academy of Arts and Sciences,
by the Johns Hopkins Society of Scholars, and as a fellow of the American Physical
Society and of the Radcliffe Institute. Dr. Stebe received a PhD in chemical engineering
at the City College of New York.

STAFF

Maggie L. Walser is associate executive director of the Division on Earth and Life
Studies and has been with the National Academies of Sciences, Engineering, and
Medicine since 2010. She previously served as senior program officer with the Board on
Chemical Sciences and Technology and as director of education and capacity building for
the Gulf Research Program, where she contributed to strategic planning for the program
and oversaw education and training activities and fellowship programs that support early-
career scientists. From 2010 to 2014, Dr. Walser was a program officer with the Board on
Atmospheric Sciences and Climate and worked on such topics as climate science, weather
research and policy, climate change and water security, and Arctic research priorities.
Before joining the staff of the National Academies, she was congressional science fellow
with the American Geophysical Union (AGU)/American Association for the
Advancement of Science, and worked on water and energy policy and legislation with the
U.S. Senate Committee on Energy and Natural Resources. Dr. Walser is past president of
the AGU Science and Society Section. She holds bachelor’s degrees in chemistry and
chemical engineering and a PhD in chemistry from the University of California, Irvine.

Brittany P. Bishop was a Christine Mirzayan science and technology policy fellow in
2020 with the Board on Chemical Sciences and Technology of the National Academies
of Sciences, Engineering, and Medicine. Her research interests include clean energy and
renewable technologies. Dr. Bishop earned a B.S.E. in chemical engineering from Case
Western Reserve University, where she researched continuous flow nanocrystal synthesis
techniques, and a PhD in chemical engineering and nanotechnology and molecular
engineering from the University of Washington.

Kesiah Clement was research associate with the Board on Chemical Sciences and
Technology of the National Academies of Sciences, Engineering, and Medicine, which
she joined as a program assistant in July 2019. She graduated from Georgetown
University’s School of Foreign Service with a B.S. in science, technology, and
international affairs focusing on global environmental health. In 2021, Ms. Clement left
the National Academies to pursue a JD at the University of Colorado Boulder Law School.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Appendix E 351

Liana Vaccari is program officer with the Board on Chemical Sciences and Technology
of the National Academies of Sciences, Engineering, and Medicine. Prior to this role, she
acted as the Resiliency Working Group lead at the New Jersey Department of
Transportation as a science fellow of the Rutgers University Eagleton Institute of Politics,
where she coordinated internal efforts to incorporate projected risks from climate change
into project prioritization, asset management, and other agency processes. Previously, Dr.
Vaccari worked at the Consortium for Ocean Leadership, managing community
engagement for the Ocean Observatories Initiative, a federally funded resource for ocean
scientists. She was a Mirzayan science and technology policy graduate fellow in 2018
with the Ocean Studies Board at the National Academies, where she contributed to studies
examining the use of dispersants in oil spills and coral reef resilience. Dr. Vaccari earned
a B.E. in chemical engineering from the Stevens Institute of Technology, a B.S. in
chemistry from New York University, and a PhD in chemical engineering from the
University of Pennsylvania.

Jessica Wolfman is research assistant for the Board on Chemical Sciences and
Technology (BCST). She joined the National Academies of Sciences, Engineering, and
Medicine in 2017 as a senior program assistant with both BCST and the Board on
Environmental Studies and Toxicology. Ms. Wolfman has worked on a variety of
activities at the National Academies, including studies examining the states of chemical
separations, the future of chemical engineering, and the chemical economy. She currently
leads a webinar series for the Chemical Sciences Roundtable, a standing body within the
BCST. Ms. Wolfman graduated Phi Beta Kappa from Dickinson College with a B.S. in
earth sciences and a minor in mathematics.

Elise Zaidi is communications and media associate for the Division on Earth and Life
Studies of the National Academies of Sciences, Engineering, and Medicine. Her primary
responsibilities include promoting report releases, creating derivative products for
National Academies projects and publications, and formulating targeted outreach cam-
paigns for committee and study activities. Prior to starting her work with the National
Academies in July 2019, Ms. Zaidi held positions with the Council on Foreign Relations
and the Pan American Health Organization. She graduated from The George Washington
University with a B.A. in international affairs with a concentration in global public health
and a minor in journalism and mass communication.

Copyright National Academy of Sciences. All rights reserved.


New Directions for Chemical Engineering

Copyright National Academy of Sciences. All rights reserved.

You might also like

pFad - Phonifier reborn

Pfad - The Proxy pFad of © 2024 Garber Painting. All rights reserved.

Note: This service is not intended for secure transactions such as banking, social media, email, or purchasing. Use at your own risk. We assume no liability whatsoever for broken pages.


Alternative Proxies:

Alternative Proxy

pFad Proxy

pFad v3 Proxy

pFad v4 Proxy