100% found this document useful (1 vote)
111 views113 pages

PMS 1,2

1) The document discusses the fundamentals of mathematical modeling for chemical engineering processes. It covers the uses of mathematical models, principles of formulation including basis, assumptions, and verification. 2) The key steps in developing a mathematical model are determining the appropriate basis in physical and chemical laws, making reasonable simplifying assumptions, ensuring mathematical consistency, and verifying the model describes the real system. 3) Two examples are provided to demonstrate applying the total continuity equation to model different systems - a perfectly mixed tank and fluid flow through a pipe. Spatial and temporal variations are considered to account for non-uniform properties.

Uploaded by

KEY RUN
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
100% found this document useful (1 vote)
111 views113 pages

PMS 1,2

1) The document discusses the fundamentals of mathematical modeling for chemical engineering processes. It covers the uses of mathematical models, principles of formulation including basis, assumptions, and verification. 2) The key steps in developing a mathematical model are determining the appropriate basis in physical and chemical laws, making reasonable simplifying assumptions, ensuring mathematical consistency, and verifying the model describes the real system. 3) Two examples are provided to demonstrate applying the total continuity equation to model different systems - a perfectly mixed tank and fluid flow through a pipe. Spatial and temporal variations are considered to account for non-uniform properties.

Uploaded by

KEY RUN
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/ 113

PROCESS MODELING

AND SIMULATION
UNIT -I
INTRODUCTION AND FUNDAMENTALS
By
Mr. Koteswara Rao Maradana
Assistant Professor
Dept. of Chemical Engineering
ANIS
 Uses of Mathematical Models: Without doubt, the most important
result of developing a mathematical model of a chemical
engineering system is the understanding that is gained of what
really makes the process.
 Thisinsight enables you to strip away from the problem the many
extraneous “confusion factors” and to get to the core of the
system. You can see more clearly the cause-and-effect relationships
between the variables.
 Mathematical models can be useful in all phases of chemical
engineering, from research and development to plant operations,
and even in business and economic studies.
 Research and development:
Determining chemical kinetic mechanisms and parameters from
laboratory or pilot-plant reaction data.
Exploring the effects of different operating conditions for
optimization and control studies; aiding in scale-up calculations.
 Design:
Exploring the sizing and arrangement of processing equipment
for dynamic performance;
Studying the interactions of various parts of the process,
particularly when material recycle or heat integration is used;
Evaluating alternative process and control structures and
strategies; simulating start-up, shutdown, and emergency
situations and procedures.
 Plant operation:
Troubleshooting control and processing problems.
Aiding in start-up and operator training.
Studying the effects of and the requirements for expansion
(bottleneck-removal) projects, Optimizing plant operation.
It is usually much cheaper, safer, and faster to conduct the kinds of
studies listed above on a mathematical model than experimentally
on an operating unit.
This is not to say that plant tests are not needed. They are a vital
part of confirming the validity of the model and of verifying
important ideas and recommendations that evolve from the
model studies.
Principles of Formulation:
➢ Basis
➢ Assumptions
➢ Mathematical consistency of model
➢ Solution of the model equations
➢ Verification
 A. BASIS.
 The basis for mathematical models are the fundamental physical
and chemical laws, such as the laws of conservation of mass,
energy, and momentum.
 To study dynamics we will use them in their general form with
time derivatives included.
 B.
ASSUMPTIONS.
 Probably the most vital role that the engineer plays in modelling
is in exercising his engineering judgment as to what assumptions
can be validly made.
 Obviously an extremely rigorous model that includes every
phenomenon down to microscopic detail would be so complex
that it would take a long time to develop and might be
impractical to solve, even on the latest supercomputers.
An engineering compromise between a rigorous description
and getting an answer that is good enough is always required.
This has been called “optimum sloppiness.” It involves
making as many simplifying assumptions as are reasonable
without “throwing out the baby with the bath water.”
 In practice, this optimum usually corresponds to a model which is
as complex as the available computing facilities will permit. More
and more this is a personal computer.
 The development of a model that incorporates the basic
phenomena occurring in the process requires a lot of skill,
ingenuity, and practice.
 Itis an area where the creativity and innovativeness of the
engineer is a key element in the success of the process.
 The assumptions that are made should be carefully considered
and listed. They impose limitations on the model that should
always be kept in mind when evaluating its predicted results.
 C. MATHEMATICAL CONSISTENCY OF MODEL.
 Once all the equations of the mathematical model have been
written, it is usually a good idea, particularly with big, complex
systems of equations, to make sure that the number of variables
equals the number of equations.
 The so-called “degrees of freedom” of the system must be zero in
order to obtain a solution. If this is not true, the system is
underspecified or over specified and something is wrong with the
formulation of the problem.
 This kind of consistency check may seem trivial, but I can testify
from sad experience that it can save many hours of frustration,
confusion, and wasted computer time.
 Checking to see that the units of all terms in all equations are
consistent is perhaps another trivial and obvious step, but one that is
often forgotten.
 Itis essential to be particularly careful of the time units of parameters
in dynamic models.
 Any units can be used (seconds, minutes, hours, etc.), but they
cannot be mixed. We will use “minutes” in most of our examples,
but it should be remembered that many parameters are commonly
on other time bases and need to be converted appropriately, e.g.,
overall heat transfer coefficients in Btu/h °F ft2 or velocity in m/s.
 Dynamic simulation results are frequently in error because the
engineer has forgotten a factor of “60” somewhere in the equations.
D. SOLUTION OF THE MODEL EQUATIONS.
 However, the available solution techniques and tools must be kept
in mind as a mathematical model is developed.
 An equation without any way to solve it is not worth much.
E. VERIFICATION.
 An important but often neglected part of developing a
mathematical model is proving that the model describes the real-
world situation.
 At the design stage this sometimes cannot be done because the
plant has not yet been built.
 However, even in this situation there are usually either similar
existing plants or a pilot plant from which some experimental
dynamic data can be obtained.
➢ FUNDAMENTAL LAWS
❖A. TOTAL CONTINUITY EQUATION (MASS BALANCE)
𝑀𝑎𝑠𝑠 𝑓𝑙𝑜𝑤 𝑀𝑎𝑠𝑠 𝑓𝑙𝑜𝑤 𝑇𝑖𝑚𝑒 𝑟𝑎𝑡𝑒 𝑜𝑓 𝑐ℎ𝑎𝑛𝑔𝑒
− =
𝑖𝑛𝑡𝑜 𝑠𝑦𝑠𝑡𝑒𝑚 𝑜𝑢𝑡 𝑜𝑓 𝑡ℎ𝑒 𝑠𝑦𝑠𝑡𝑒𝑚 𝑜𝑓 𝑚𝑎𝑠𝑠 𝑖𝑛𝑠𝑖𝑑𝑒 𝑠𝑦𝑠𝑡𝑒𝑚
✓ The units of this equation are mass per time.
✓ Only one total continuity equation can be written for one
system.
✓ The normal steady state design equation that we are accustomed
to using says that “what goes in, comes out.”
✓ The right-hand side of Eq. will be either a partial derivative 𝜕/𝜕𝑡
or an ordinary derivative d/dt of the mass inside the system with
respect to the independent variable t.
Problem 1: Consider the tank of perfectly mixed
𝐹0 (𝑡)
liquid as shown in Fig. Into which flows a liquid
𝜌0 (𝑡)
stream at a volumetric rate of F0 (ft3/min or
m3/min) and with a density of 𝜌0 (lbm/ ft3 or kg/
𝐹(𝑡)
m3). The volumetric holdup of liquid in the tank is
V (ft3 or m3), and its density is 𝜌. The volumetric 𝑉 𝑡 𝜌(𝑡)
flow rate from the tank is F, and the density of the 𝜌(𝑡)
outflowing stream is the same as that of the tank’s
✓ Since the liquid is perfectly
contents. mixed, the density is the
Solution: The mass balance is around the whole same everywhere in the
tank; it does not vary with
tank radial or axial position; i.e.,
𝑀𝑎𝑠𝑠 𝑓𝑙𝑜𝑤 𝑀𝑎𝑠𝑠 𝑓𝑙𝑜𝑤 there are no spatial gradients
− in density in the tank.
𝑖𝑛𝑡𝑜 𝑠𝑦𝑠𝑡𝑒𝑚 𝑜𝑢𝑡 𝑜𝑓 𝑡ℎ𝑒 𝑠𝑦𝑠𝑡𝑒𝑚 ✓ This is why we can use a
𝑇𝑖𝑚𝑒 𝑟𝑎𝑡𝑒 𝑜𝑓 𝑐ℎ𝑎𝑛𝑔𝑒 macroscopic system. It also
= means that there is only one
𝑜𝑓 𝑚𝑎𝑠𝑠 𝑖𝑛𝑠𝑖𝑑𝑒 𝑠𝑦𝑠𝑡𝑒𝑚 independent variable, t.
𝐹0 𝜌0 − 𝐹𝜌 = 𝑡𝑖𝑚𝑒 𝑟𝑎𝑡𝑒 𝑜𝑓 𝑐ℎ𝑎𝑛𝑔𝑒 𝑜𝑓 𝜌𝑉
𝐹0 (𝑡)
The units of this equation are lbm/min or 𝜌0 (𝑡)
kg/min.
3 𝑙𝑏𝑚
3
𝑓𝑡 𝑙𝑏𝑚 3
𝑓𝑡 𝑙𝑏𝑚 𝑓𝑡 𝐹(𝑡)
𝑓𝑡 3
. 3 − . 3 = 𝑉 𝑡 𝜌(𝑡)
𝑚𝑖𝑛 𝑓𝑡 𝑚𝑖𝑛 𝑓𝑡 𝑚𝑖𝑛 𝜌(𝑡)
Since ρ and V are functions only of t, an
ordinary derivative is used
𝑑 𝜌𝑉
𝐹0 𝜌 − 𝐹𝜌 =
𝑑𝑡
Problem 2: Fluid is flowing through a constant-diameter cylindrical
pipe sketched in Fig. The flow is turbulent and therefore we can
assume plug-flow conditions, i.e., each “slice” of liquid flows down
the pipe as a unit. There are no radial gradients in velocity or any
other properties. However, axial gradients can exist.
dZ

𝒱(𝑡,𝑧)
𝜌(𝑡,𝑧)

Z=0 Z Z+dZ Z=L


Solution: Density and velocity can change as the fluid flows along
the axial or z direction. There are now two independent variables:
time t and position z. Density and velocity are functions of both t
and z: 𝜌 𝑡,𝑧 𝑎𝑛𝑑 𝑣 𝑡,𝑧 .
Apply the total continuity equation to a system that consists of a
small slice. The system is now a “microscopic” one.
The differential element is located at an arbitrary spot z down the
pipe. It is dz thick and has an area equal to the cross-sectional area
of the pipe A (𝑓𝑡 2 𝑜𝑟 𝑚2 ). dZ
ʋ(𝑡,𝑧)
𝜌(𝑡,𝑧)
Z=0 Z Z+dZ Z=L
The mass balance is around the small slice
𝑀𝑎𝑠𝑠 𝑓𝑙𝑜𝑤 𝑀𝑎𝑠𝑠 𝑓𝑙𝑜𝑤 𝑇𝑖𝑚𝑒 𝑟𝑎𝑡𝑒 𝑜𝑓 𝑐ℎ𝑎𝑛𝑔𝑒
− =
𝑖𝑛𝑡𝑜 𝑠𝑦𝑠𝑡𝑒𝑚 𝑜𝑢𝑡 𝑜𝑓 𝑡ℎ𝑒 𝑠𝑦𝑠𝑡𝑒𝑚 𝑜𝑓 𝑚𝑎𝑠𝑠 𝑖𝑛𝑠𝑖𝑑𝑒 𝑠𝑦𝑠𝑡𝑒𝑚
𝜕 𝐴𝜌𝑑𝑧
Time rate of change of mass inside system:
𝜕𝑡
Where Adz is the volume of the system; ρ is the density. The units
of this equation are lbm/min or kg/mm.
Mass flowing into system through boundary at z: 𝜌𝒱𝐴
Notice that the units are still lbm/min = (lbm/ft3) (ft/min)(ft2).
Mass flowing out of the system through boundary at z + dz: 𝜌𝒱𝐴 +
𝜕 𝜌𝒱𝐴
𝑑𝑧
𝜕𝑍
The above expression for the flow at z + dz may be thought of as a Taylor
series expansion of a function 𝑓 𝑧 around z. The value of the function at a
spot dz away from z is dZ

𝜕𝑓 𝜕2 𝑓 𝑑𝑧 2 ʋ(𝑡,𝑧) 𝜌(𝑡,𝑧)
𝑓 𝑧+𝑑𝑧 = 𝑓𝑧 + 𝑑𝑧 + +⋯ Z=0 Z Z+dZ Z=L
𝜕𝑧 𝑧 𝜕𝑧 2 𝑧 2!
If the dz is small, the series can be truncated after the first derivative term.
Letting 𝑓 𝑧 = 𝑣𝐴𝜌
𝜕 𝜌𝒱𝐴 𝜕 𝐴𝜌𝑑𝑧
𝜌𝒱𝐴 − 𝜌𝒱𝐴 + 𝑑𝑧 =
𝜕𝑍 𝜕𝑡
Cancelling out the dz terms and assuming A is constant yield
𝜕𝜌 𝜕 𝜌𝒱
+ =0
𝜕𝑡 𝜕𝑍
B. COMPONENT CONTINUITY EQUATIONS
(COMPONENT BALANCES).
𝐹𝑙𝑜𝑤 𝑜𝑓 𝑚𝑜𝑙𝑒𝑠 𝑜𝑓 𝑗𝑡ℎ 𝐹𝑙𝑜𝑤 𝑜𝑓 𝑚𝑜𝑙𝑒𝑠 𝑜𝑓 𝑗𝑡ℎ

𝑐𝑜𝑚𝑝𝑜𝑛𝑒𝑛𝑡 𝑖𝑛𝑡𝑜 𝑠𝑦𝑠𝑡𝑒𝑚 𝑐𝑜𝑚𝑝𝑜𝑛𝑒𝑛𝑡 𝑜𝑢𝑡 𝑜𝑓 𝑡ℎ𝑒 𝑠𝑦𝑠𝑡𝑒𝑚
𝑟𝑎𝑡𝑒 𝑜𝑓 𝑓𝑜𝑟𝑚𝑎𝑡𝑖𝑜𝑛 𝑜𝑓 𝑚𝑜𝑙𝑒𝑠 𝑜𝑓 𝑗𝑡ℎ
+
𝑐𝑜𝑚𝑝𝑜𝑛𝑒𝑛𝑡 𝑓𝑟𝑜𝑚 𝑐ℎ𝑒𝑚𝑖𝑐𝑎𝑙 𝑟𝑒𝑎𝑐𝑡𝑖𝑜𝑛𝑠
𝑇𝑖𝑚𝑒 𝑟𝑎𝑡𝑒 𝑜𝑓 𝑐ℎ𝑎𝑛𝑔𝑒 𝑜𝑓 𝑚𝑜𝑙𝑒𝑠 𝑜𝑓 𝑗𝑡ℎ
=
𝑐𝑜𝑚𝑝𝑜𝑛𝑒𝑛𝑡 𝑖𝑛𝑠𝑖𝑑𝑒 𝑠𝑦𝑠𝑡𝑒𝑚
✓ The units of this equation are moles of component j per unit time.
✓ The flows in and out can be both convective (due to bulk flow) and
molecular (due to diffusion).
✓ We can write one component continuity equation for each
component in the system.
✓ If there are NC components, there are NC component continuity
equations for any one system.
✓ However, the one total mass balance and these NC component
balances are not all independent, since the sum of all the moles
times their respective molecular weights equals the total mass.
✓ Therefore a given system has only NC independent continuity
equations.
✓ We usually use the total mass balance and NC - 1 component
balances.
✓ For example, in a binary (two-component) system, there would be
one total mass balance and one component balance.
Problem 3. Consider the perfectly mixed tank, a 𝐹0
chemical reaction takes place in the liquid in the 𝜌0
tank. The system is now a CSTR (continuous stirred- 𝐶𝐴0
tank reactor) as shown in Fig. Component A reacts 𝐶𝐵0
irreversibly and at a specific reaction rate k to form
product, component B.
𝑘 𝑉 𝑡 𝐹
𝐴՜𝐵 𝜌 𝑡 𝜌
𝐶𝐴 𝐶𝐴
Let the concentration of component A in the 𝐶𝐵 𝐶𝐵
inflowing feed stream be CA0 (moles of A per unit
volume) and in the reactor CA. Assuming a simple
first-order reaction, the rate of consumption of
reactant A per unit volume will be directly
proportional to the instantaneous concentration of
A in the tank.
❖ Component ‘A’ balance 𝐹0
𝐹𝑙𝑜𝑤 𝑜𝑓 𝑚𝑜𝑙𝑒𝑠 𝑜𝑓 𝜌0
𝑐𝑜𝑚𝑝𝑜𝑛𝑒𝑛𝑡 𝐴 𝑖𝑛𝑡𝑜 𝑠𝑦𝑠𝑡𝑒𝑚 𝐶𝐴0
𝐹𝑙𝑜𝑤 𝑜𝑓 𝑚𝑜𝑙𝑒𝑠 𝑜𝑓 𝐶𝐵0

𝑐𝑜𝑚𝑝𝑜𝑛𝑒𝑛𝑡 𝐴 𝑜𝑢𝑡 𝑜𝑓 𝑡ℎ𝑒 𝑠𝑦𝑠𝑡𝑒𝑚
𝑟𝑎𝑡𝑒 𝑜𝑓 𝑓𝑜𝑟𝑚𝑎𝑡𝑖𝑜𝑛 𝑜𝑓 𝑚𝑜𝑙𝑒𝑠 𝑜𝑓
+ 𝑉 𝑡
𝑐𝑜𝑚𝑝𝑜𝑛𝑒𝑛𝑡 𝐴 𝑓𝑟𝑜𝑚 𝑐ℎ𝑒𝑚𝑖𝑐𝑎𝑙 𝑟𝑒𝑎𝑐𝑡𝑖𝑜𝑛𝑠 𝐹
𝑇𝑖𝑚𝑒 𝑟𝑎𝑡𝑒 𝑜𝑓 𝑐ℎ𝑎𝑛𝑔𝑒 𝑜𝑓 𝑚𝑜𝑙𝑒𝑠 𝑜𝑓 𝜌 𝑡 𝜌
= 𝐶𝐴 𝐶𝐴
𝑐𝑜𝑚𝑝𝑜𝑛𝑒𝑛𝑡 𝐴 𝑖𝑛𝑠𝑖𝑑𝑒 𝑠𝑦𝑠𝑡𝑒𝑚 𝐶𝐵 𝐶𝐵
✓ 𝐹𝑙𝑜𝑤 𝑜𝑓 𝐴 𝑖𝑛𝑡𝑜 𝑠𝑦𝑠𝑡𝑒𝑚 = 𝐹0 𝐶𝐴0
✓ 𝐹𝑙𝑜𝑤 𝑜𝑓 𝐴 𝑜𝑢𝑡 𝑜𝑓 𝑠𝑦𝑠𝑡𝑒𝑚 = 𝐹𝐶𝐴
✓ 𝑅𝑎𝑡𝑒 𝑜𝑓 𝑓𝑜𝑟𝑚𝑎𝑡𝑖𝑜𝑛 𝑜𝑓 𝐴 𝑓𝑟𝑜𝑚 𝑟𝑒𝑎𝑐𝑡𝑖𝑜𝑛 = − 𝑉𝑘𝐶𝐴
✓ The minus sign comes from the fact that A is being consumed.
𝑑 𝑉𝐶𝐴
✓ 𝑇𝑖𝑚𝑒 𝑟𝑎𝑡𝑒 𝑜𝑓 𝑐ℎ𝑎𝑛𝑔𝑒 𝑜𝑓 𝐴 𝑖𝑛𝑠𝑖𝑑𝑒 𝑡𝑎𝑛𝑘 =
𝑑𝑡
𝑑 𝑉𝐶𝐴
✓ 𝐹0 𝐶𝐴0 − 𝐹𝐶𝐴 − 𝑉𝑘𝐶𝐴 =
𝑑𝑡
❖ Component ‘B’ balance
𝐹𝑙𝑜𝑤 𝑜𝑓 𝑚𝑜𝑙𝑒𝑠 𝑜𝑓 𝐹𝑙𝑜𝑤 𝑜𝑓 𝑚𝑜𝑙𝑒𝑠 𝑜𝑓

𝑐𝑜𝑚𝑝𝑜𝑛𝑒𝑛𝑡 𝐵 𝑖𝑛𝑡𝑜 𝑠𝑦𝑠𝑡𝑒𝑚 𝑐𝑜𝑚𝑝𝑜𝑛𝑒𝑛𝑡 𝐵 𝑜𝑢𝑡 𝑜𝑓 𝑡ℎ𝑒 𝑠𝑦𝑠𝑡𝑒𝑚
𝑟𝑎𝑡𝑒 𝑜𝑓 𝑓𝑜𝑟𝑚𝑎𝑡𝑖𝑜𝑛 𝑜𝑓 𝑚𝑜𝑙𝑒𝑠 𝑜𝑓 𝐹0
+
𝑐𝑜𝑚𝑝𝑜𝑛𝑒𝑛𝑡 𝐵 𝑓𝑟𝑜𝑚 𝑐ℎ𝑒𝑚𝑖𝑐𝑎𝑙 𝑟𝑒𝑎𝑐𝑡𝑖𝑜𝑛𝑠 𝜌0
𝑇𝑖𝑚𝑒 𝑟𝑎𝑡𝑒 𝑜𝑓 𝑐ℎ𝑎𝑛𝑔𝑒 𝑜𝑓 𝑚𝑜𝑙𝑒𝑠 𝑜𝑓 𝐶𝐴0
= 𝐶𝐵0
𝑐𝑜𝑚𝑝𝑜𝑛𝑒𝑛𝑡 𝐵 𝑖𝑛𝑠𝑖𝑑𝑒 𝑠𝑦𝑠𝑡𝑒𝑚
𝑑 𝑉𝐶𝐵
✓ 𝐹0 𝐶𝐵0 − 𝐹𝐶𝐵 + 𝑉𝑘𝐶𝐴 = 𝑉 𝑡 𝐹
𝑑𝑡
❖ Total continuity equation: 𝑀𝐴 𝐶𝐴 + 𝑀𝐵 𝐶𝐵 = 𝜌 𝜌 𝑡 𝜌
𝐶𝐴
where MA and MB, are the molecular weights of 𝐶𝐵
𝐶𝐴
components A and B, respectively. 𝐶𝐵
Problem 4: Suppose we have the same macroscopic system as above
except that now consecutive reactions occur. Reactant A goes to B at a
specific reaction rate k1 but B can react at a specific reaction rate k2, to
form a third component C. write the component balance equation for
each component and total continuity equation.
𝑘1 𝑘2
𝐴՜ 𝐵՜ 𝐶
❖ Continuity equations for Component ‘A, B and C’ are:
𝐹𝑙𝑜𝑤 𝑜𝑓 𝑚𝑜𝑙𝑒𝑠 𝑜𝑓 𝑗𝑡ℎ 𝐹𝑙𝑜𝑤 𝑜𝑓 𝑚𝑜𝑙𝑒𝑠 𝑜𝑓 𝑗𝑡ℎ
❖ − +
𝑐𝑜𝑚𝑝𝑜𝑛𝑒𝑛𝑡 𝑖𝑛𝑡𝑜 𝑠𝑦𝑠𝑡𝑒𝑚 𝑐𝑜𝑚𝑝𝑜𝑛𝑒𝑛𝑡 𝑜𝑢𝑡 𝑜𝑓 𝑡ℎ𝑒 𝑠𝑦𝑠𝑡𝑒𝑚
𝑟𝑎𝑡𝑒 𝑜𝑓 𝑓𝑜𝑟𝑚𝑎𝑡𝑖𝑜𝑛 𝑜𝑓 𝑚𝑜𝑙𝑒𝑠 𝑜𝑓 𝑗𝑡ℎ
=
𝑐𝑜𝑚𝑝𝑜𝑛𝑒𝑛𝑡 𝑓𝑟𝑜𝑚 𝑐ℎ𝑒𝑚𝑖𝑐𝑎𝑙 𝑟𝑒𝑎𝑐𝑡𝑖𝑜𝑛𝑠
𝑇𝑖𝑚𝑒 𝑟𝑎𝑡𝑒 𝑜𝑓 𝑐ℎ𝑎𝑛𝑔𝑒 𝑜𝑓 𝑚𝑜𝑙𝑒𝑠 𝑜𝑓 𝑗𝑡ℎ
𝑐𝑜𝑚𝑝𝑜𝑛𝑒𝑛𝑡 𝑖𝑛𝑠𝑖𝑑𝑒 𝑠𝑦𝑠𝑡𝑒𝑚
Component ‘A’ balance
𝑑 𝑉𝐶𝐴
✓ 𝐹0 𝐶𝐴0 − 𝐹𝐶𝐴 − 𝑉𝑘1 𝐶𝐴 =
𝑑𝑡
Component ‘B’ balance
𝑑 𝑉𝐶𝐵
✓ 𝐹0 𝐶𝐵0 − 𝐹𝐶𝐵 + 𝑉𝑘1 𝐶𝐴 − 𝑉𝑘2 𝐶𝐵 =
𝑑𝑡
Component ‘C’ balance
𝑑 𝑉𝐶𝐶
✓ 𝐹0 𝐶𝐶0 − 𝐹𝐶𝐶 + 𝑉𝑘2 𝐶𝐵 =
𝑑𝑡
Total continuity equation:
❖ 𝑀𝐴 𝐶𝐴 + 𝑀𝐵 𝐶𝐵 + 𝑀𝐶 𝐶𝐶 = 𝜌 = σ𝐶𝑗=𝐴 𝑀𝑗 𝐶𝑗
Problem 5. Fluid flowing down a tubular reactor in which the reaction
𝑘
𝐴 ՜ 𝐵 takes place. As a slice of material moves down the length of the
reactor the concentration of reactant CA decreases as A is consumed.
Density ρ, velocity 𝜐, and concentration CA can all vary with time and
axial position z. We still assume plug-flow conditions so that there are
no radial gradients in velocity, density, or concentration.
dZ
𝐶𝐴0 (𝑡) ʋ(𝑡,𝑧) 𝐶𝐴𝐿 (𝑡)
𝜌 𝑡,𝑧
𝐶𝐴(𝑡,𝑧)
Z=0 Z Z+dZ Z=L

Solution: The concentration of A fed to the inlet of the reactor at z = 0


is defined as
𝐶𝐴 𝑡,0 = 𝐶𝐴𝑂 𝑡
The concentration of A in the reactor effluent at z = L is defined as
𝐶𝐴 𝑡,𝐿 = 𝐶𝐴𝐿 𝑡
Apply the component continuity equation for reactant A to a small
differential slice of width dz, as shown in Fig.
The inflow terms can be split into two types: bulk flow and diffusion.
Diffusion can occur because of the concentration gradient in the axial
direction.
It is usually much less important than bulk flow in most practical
systems, but we include it here to see what it contributes to the model.
We will say that the diffusive flux of A, NA (moles of A per unit
time per unit area), is given by a Fick’s law type of relationship
dZ
𝜕𝐶𝐴 𝐶𝐴𝐿 (𝑡)
𝑁𝐴 = −𝒟𝐴 𝐶𝐴0 (𝑡) ʋ(𝑡,𝑧) 𝜌 𝑡,𝑧
𝜕𝑧 𝐶𝐴 (𝑡,𝑧)
Z=0 Z Z+dZ Z=L
Where 𝒟𝐴 , is a diffusion coefficient due to both diffusion and
turbulence in the fluid flow (so-called “eddy diffusivity”).
dZ
𝒟𝐴 has units
of length2 per unit time. 𝐶𝐴0 (𝑡)
ʋ(𝑡,𝑧)
𝜌 𝑡,𝑧 𝐶𝐴𝐿 (𝑡)
𝐶𝐴
Component continuity equation: Z=0 Z Z+dZ
(𝑡,𝑧)
Z=L
𝐹𝑙𝑜𝑤 𝑜𝑓 𝑚𝑜𝑙𝑒𝑠 𝑜𝑓 𝑗𝑡ℎ 𝐹𝑙𝑜𝑤 𝑜𝑓 𝑚𝑜𝑙𝑒𝑠 𝑜𝑓 𝑗𝑡ℎ

𝑐𝑜𝑚𝑝𝑜𝑛𝑒𝑛𝑡 𝑖𝑛𝑡𝑜 𝑠𝑦𝑠𝑡𝑒𝑚 𝑐𝑜𝑚𝑝𝑜𝑛𝑒𝑛𝑡 𝑜𝑢𝑡 𝑜𝑓 𝑡ℎ𝑒 𝑠𝑦𝑠𝑡𝑒𝑚
𝑟𝑎𝑡𝑒 𝑜𝑓 𝑓𝑜𝑟𝑚𝑎𝑡𝑖𝑜𝑛 𝑜𝑓 𝑚𝑜𝑙𝑒𝑠 𝑜𝑓 𝑗𝑡ℎ 𝑇𝑖𝑚𝑒 𝑟𝑎𝑡𝑒 𝑜𝑓 𝑐ℎ𝑎𝑛𝑔𝑒 𝑜𝑓 𝑚𝑜𝑙𝑒𝑠 𝑜𝑓 𝑗𝑡ℎ
+ =
𝑐𝑜𝑚𝑝𝑜𝑛𝑒𝑛𝑡 𝑓𝑟𝑜𝑚 𝑐ℎ𝑒𝑚𝑖𝑐𝑎𝑙 𝑟𝑒𝑎𝑐𝑡𝑖𝑜𝑛𝑠 𝑐𝑜𝑚𝑝𝑜𝑛𝑒𝑛𝑡 𝑖𝑛𝑠𝑖𝑑𝑒 𝑠𝑦𝑠𝑡𝑒𝑚
𝑀𝑜𝑙𝑎𝑟 𝑓𝑙𝑜𝑤 𝑜𝑓 𝐴 𝑖𝑛𝑡𝑜 𝑏𝑜𝑢𝑛𝑑𝑎𝑟𝑦 𝑎𝑡 𝑧 𝑏𝑢𝑙𝑘 𝑓𝑙𝑜𝑤 𝑎𝑛𝑑 𝑑𝑖𝑓𝑓𝑢𝑠𝑖𝑜𝑛
𝑚𝑜𝑙𝑒𝑠 𝑜𝑓 𝐴
= 𝑣𝐴𝐶𝐴 + 𝐴𝑁𝐴
𝑠
Molar flow of A leaving system at boundary z + dz = ( 𝑣𝐴𝐶𝐴 +
𝜕 𝑣𝐴𝐶𝐴 + 𝐴𝑁𝐴
𝐴𝑁𝐴 ) + 𝑑𝑧
𝜕𝑧
Rate of formation of A inside system = −𝑘𝐶𝐴 𝐴𝑑𝑧
𝜕 𝐶𝐴 𝐴𝑑𝑧
Time rate of change of A inside system =
𝜕𝑡
Substituting into Component continuity equation
dZ
𝐶𝐴0 (𝑡) ʋ(𝑡,𝑧) 𝐶𝐴𝐿 (𝑡)
𝜌 𝑡,𝑧
𝜕 𝐶𝐴 𝐴𝑑𝑧 𝐶𝐴(𝑡,𝑧)
𝜕𝑡 Z=0 Z Z+dZ Z=L

𝜕 𝑣𝐴𝐶𝐴 + 𝐴𝑁𝐴
= 𝑣𝐴𝐶𝐴 + 𝐴𝑁𝐴 − 𝑣𝐴𝐶𝐴 + 𝐴𝑁𝐴 + 𝑑𝑧
𝜕𝑧
− 𝑘𝐶𝐴 𝐴𝑑𝑧
Divide both sides with 𝐴𝑑𝑧
𝜕 𝐶𝐴 𝜕 𝑣𝐶𝐴 + 𝑁𝐴
+ + 𝑘𝐶𝐴 = 0
𝜕𝑡 𝜕𝑧
𝜕 𝐶𝐴 𝜕 𝑣𝐶𝐴 𝜕 𝑁𝐴
+ + + 𝑘𝐶𝐴 = 0
𝜕𝑡 𝜕𝑧 𝜕𝑡
𝜕𝐶𝐴
Substitute 𝑁𝐴 = −𝒟𝐴
𝜕𝑧
𝜕𝐶𝐴
𝜕 𝐶𝐴 𝜕 𝑣𝐶𝐴 𝜕 𝒟𝐴
𝜕𝑧
+ + 𝑘𝐶𝐴 =
𝜕𝑡 𝜕𝑧 𝜕𝑧
The units of the equation are moles A per volume per time.
Similarly
Component ‘B’ balance
𝜕𝐶𝐵
𝜕 𝐶𝐵 𝜕 𝑣𝐶𝐵 𝜕 𝒟𝐵
𝜕𝑧
+ − 𝑘𝐶𝐴 =
𝜕𝑡 𝜕𝑧 𝜕𝑧
2. Energy Balance Equation
The first law of thermodynamics puts forward the principle of
conservation of energy. Written for a general “open” system (where
flow of material in and out of the system can occur) it is
𝐹𝑙𝑜𝑤 𝑜𝑓 𝑖𝑛𝑡𝑒𝑟𝑛𝑎𝑙, 𝑘𝑖𝑛𝑒𝑡𝑖𝑐, 𝑎𝑛𝑑 𝐹𝑙𝑜𝑤 𝑜𝑓 𝑖𝑛𝑡𝑒𝑟𝑛𝑎𝑙, 𝑘𝑖𝑛𝑒𝑡𝑖𝑐, 𝑎𝑛𝑑
𝑝𝑜𝑡𝑒𝑛𝑡𝑖𝑎𝑙 𝑒𝑛𝑒𝑟𝑔𝑦 𝑖𝑛𝑡𝑜 𝑠𝑦𝑠𝑡𝑒𝑚 − 𝑝𝑜𝑡𝑒𝑛𝑡𝑖𝑎𝑙 𝑒𝑛𝑒𝑟𝑔𝑦 𝑜𝑢𝑡 𝑜𝑓 𝑠𝑦𝑠𝑡𝑒𝑚
𝑏𝑦 𝑐𝑜𝑛𝑣𝑒𝑐𝑡𝑖𝑜𝑛 𝑜𝑟 𝑑𝑖𝑓𝑓𝑢𝑠𝑖𝑜𝑛 𝑏𝑦 𝑐𝑜𝑛𝑣𝑒𝑐𝑡𝑖𝑜𝑛 𝑜𝑟 𝑑𝑖𝑓𝑓𝑢𝑠𝑖𝑜𝑛
𝐻𝑒𝑎𝑡 𝑎𝑑𝑑𝑒𝑑 𝑡𝑜 𝑠𝑦𝑠𝑡𝑒𝑚 𝑏𝑦
+
𝑐𝑜𝑛𝑑𝑢𝑐𝑡𝑖𝑜𝑛, 𝑟𝑎𝑑𝑖𝑎𝑡𝑖𝑜𝑛 𝑎𝑛𝑑 𝑟𝑒𝑎𝑐𝑡𝑖𝑜𝑛
𝑊𝑜𝑟𝑘 𝑑𝑜𝑛𝑒 𝑏𝑦 𝑠𝑦𝑠𝑡𝑒𝑚 𝑜𝑛 𝑠𝑢𝑟𝑟𝑜𝑢𝑛𝑑𝑖𝑛𝑔𝑠

𝑠ℎ𝑎𝑓𝑡 𝑤𝑜𝑟𝑘 𝑎𝑛𝑑 𝑃𝑉 𝑤𝑜𝑟𝑘
𝑇𝑖𝑚𝑒 𝑟𝑎𝑡𝑒 𝑜𝑓 𝑐ℎ𝑎𝑛𝑔𝑒 𝑜𝑓 𝑖𝑛𝑡𝑒𝑟𝑛𝑎𝑙,
=
𝑘𝑖𝑛𝑒𝑡𝑖𝑐, 𝑝𝑜𝑡𝑒𝑛𝑡𝑖𝑎𝑙 𝑒𝑛𝑒𝑟𝑔𝑦 𝑖𝑛𝑠𝑖𝑑𝑒 𝑠𝑦𝑠𝑡𝑒𝑚
Problem 6. Consider the perfectly mixed tank, a chemical reaction takes
𝑘
place in the liquid in the tank 𝐴՜𝐵 . The CSTR system with a cooling
coil inside the tank that can remove the exothermic heat of reaction 𝜆
(Btu/lb . mol of A reacted or Cal/g. mol of A reacted). We use the
normal convention that λ is negative for an exothermic reaction and
positive for an endothermic reaction. The rate of heat generation
(energy per time) due to reaction is the rate of consumption of A times
𝜆.
𝑄𝐺 = −𝜆𝑉𝐶𝐴 𝑘
The rate of heat removal from the reaction mass to the cooling coil is -
Q (energy per time). The temperature of the feed stream is To and the
temperature in the reactor is T (°R or K).
Solution: Energy Balance Equation: 𝐹0
𝜌0
𝐹𝑙𝑜𝑤 𝑜𝑓 𝑖𝑛𝑡𝑒𝑟𝑛𝑎𝑙, 𝑘𝑖𝑛𝑒𝑡𝑖𝑐, 𝑎𝑛𝑑 𝐹𝑙𝑜𝑤 𝑜𝑓 𝑖𝑛𝑡𝑒𝑟𝑛𝑎𝑙, 𝑘𝑖𝑛𝑒𝑡𝑖𝑐, 𝑎𝑛𝑑 𝐶𝐴0
𝑝𝑜𝑡𝑒𝑛𝑡𝑖𝑎𝑙 𝑒𝑛𝑒𝑟𝑔𝑦 𝑖𝑛𝑡𝑜 𝑠𝑦𝑠𝑡𝑒𝑚 − 𝑝𝑜𝑡𝑒𝑛𝑡𝑖𝑎𝑙 𝑒𝑛𝑒𝑟𝑔𝑦 𝑜𝑢𝑡 𝑜𝑓 𝑠𝑦𝑠𝑡𝑒𝑚 𝑇0
𝑏𝑦 𝑐𝑜𝑛𝑣𝑒𝑐𝑡𝑖𝑜𝑛 𝑜𝑟 𝑑𝑖𝑓𝑓𝑢𝑠𝑖𝑜𝑛 𝑏𝑦 𝑐𝑜𝑛𝑣𝑒𝑐𝑡𝑖𝑜𝑛 𝑜𝑟 𝑑𝑖𝑓𝑓𝑢𝑠𝑖𝑜𝑛
ℎ𝑒𝑎𝑡 𝑎𝑑𝑑𝑒𝑑 𝑡𝑜 𝑠𝑦𝑠𝑡𝑒𝑚 𝑏𝑦 𝑤𝑜𝑟𝑘 𝑑𝑜𝑛𝑒 𝑏𝑦 𝑠𝑦𝑠𝑡𝑒𝑚 𝑜𝑛 𝑠𝑢𝑟𝑟𝑜𝑢𝑛𝑑𝑖𝑛𝑔𝑠
+ − 𝐹
𝑐𝑜𝑛𝑑𝑢𝑐𝑡𝑖𝑜𝑛, 𝑟𝑎𝑑𝑖𝑎𝑡𝑖𝑜𝑛 𝑎𝑛𝑑 𝑟𝑒𝑎𝑐𝑡𝑖𝑜𝑛 𝑠ℎ𝑎𝑓𝑡 𝑤𝑜𝑟𝑘 𝑎𝑛𝑑 𝑃𝑉 𝑤𝑜𝑟𝑘
𝑇𝑖𝑚𝑒 𝑟𝑎𝑡𝑒 𝑜𝑓 𝑐ℎ𝑎𝑛𝑔𝑒 𝑜𝑓 𝑖𝑛𝑡𝑒𝑟𝑛𝑎𝑙, 𝜌
= 𝑉 𝐶𝐴
𝑘𝑖𝑛𝑒𝑡𝑖𝑐, 𝑝𝑜𝑡𝑒𝑛𝑡𝑖𝑎𝑙 𝑒𝑛𝑒𝑟𝑔𝑦 𝑖𝑛𝑠𝑖𝑑𝑒 𝑠𝑦𝑠𝑡𝑒𝑚
𝜌 𝑇
-Q
𝐶𝐴
𝑇

where U = internal energy (energy per unit mass)


K = kinetic energy (energy per unit mass)
𝜙 = potential energy (energy per unit mass) Note that all the terms in above Eq.
W = shaft work done by system (energy per time) must have the same units (energy
P = pressure of system per time) so the FP terms must use
P0 = pressure of feed stream the appropriate conversion factor
(778 ft. lbf/Btu in English
engineering units)
W=0, K, 𝜙 are negligible
𝑑 𝑉𝜌𝑈 𝑃 𝑃0
= 𝐹0 𝜌0 𝑈0 − 𝐹𝜌𝑈 + 𝑄𝐺 + 𝑄 − 𝐹𝜌 + 𝐹0 𝜌0
𝑑𝑡 𝜌 𝜌0
= 𝐹0 𝜌0 (𝑈0 + 𝑃0 𝑉0 ) − 𝐹𝜌 𝑈 + 𝑃𝑉ത + 𝑄𝐺 + 𝑄
1
where ഥ𝑉 = is the specific volume (ft3/lbm or m3/kg), the reciprocal of the
𝜌0
𝐹0
density. 𝜌0
𝐶𝐴0
Enthalpy, H or h, is defined: 𝑇0
𝐹
𝐻 𝑜𝑟 ℎ = 𝑈 + 𝑃𝑉ത 𝑉
𝜌
𝜌
-Q 𝐶𝐴
𝐶𝐴
𝑑 𝜌𝑉𝑈 𝑇 𝑇
= 𝐹0 𝜌0 ℎ0 − 𝐹𝜌ℎ + 𝑄 − 𝜆𝑉𝑘𝐶𝐴
𝑑𝑡
For liquids the 𝑃𝑉ത term is negligible compared to the U term, and we use the
time rate of change of the enthalpy of the system instead of the internal
energy of the system.
𝑑 𝜌𝑉ℎ
= 𝐹0 𝜌0 ℎ0 − 𝐹𝜌ℎ + 𝑄 − 𝜆𝑉𝑘𝐶𝐴
𝑑𝑡
The enthalpies are functions of composition, temperature, and pressure, but
primarily temperature. From thermodynamics, the heat capacities at constant
pressure, Cp and at constant volume, CV are
𝜕𝐻 𝜕𝑈
𝐶𝑃 = 𝐶𝑉 =
𝜕𝑇 𝑃 𝜕𝑇 𝑉
To illustrate that the energy is primarily influenced by temperature, let us
simplify the problem by assuming that the liquid enthalpy can be expressed
as a product of absolute temperature and an average heat capacity CP
(Btu/lbm°R or Cal/g K) that is constant. ℎ = 𝐶𝑃 𝑇
We will also assume that the densities of all the liquid streams are constant.
With these simplifications energy balance Equation becomes
𝑑 𝑉𝑇
𝜌𝑐𝑃 = 𝜌𝑐𝑃 𝐹0 𝑇0 − 𝐹𝑇 + 𝑄 − 𝜆𝑉𝑘𝐶𝐴
𝑑𝑡
Problem 7. To show what form the energy equation takes for a two-
phase system, consider the CSTR process shown in Fig.. Both a liquid
product stream F and a vapor product stream Fv(volumetric flow) are
withdrawn from the vessel. The pressure in the reactor is P. Vapor and
liquid volumes are Vv and V. The density and temperature of the
vapor phase are ρv and Tv . The mole fraction of A in the vapor is y. If
the phases are in thermal equilibrium, the vapor and liquid
temperatures are equal ( 𝑇 = 𝑇𝜈 ). If the phases are in phase
equilibrium, the liquid and vapor compositions are related by Raoult’s
law, a relative volatility relationship or some other vapor-liquid
equilibrium relationship. The enthalpy of the vapor phase H (Btu/lb,
or Cal/g) is a function of composition y, temperature 𝑇𝜈 and pressure
P. Neglecting kinetic-energy and potential-energy terms and the work
term, and replacing internal energies with enthalpies.
𝐹𝜈
𝑇𝜈
𝑦
𝑉𝜈 𝑃 𝜌𝜈 𝑇𝜈 𝑦 𝜌𝜈
𝐹0
𝐶𝐴0 𝑉𝐿 𝐶𝐴 𝜌 𝑇
𝐹
𝜌0 𝐶𝐴
𝑇0 𝜌

Solution: Energy Balance Equation: -Q

𝐹𝑙𝑜𝑤 𝑜𝑓 𝑖𝑛𝑡𝑒𝑟𝑛𝑎𝑙, 𝑘𝑖𝑛𝑒𝑡𝑖𝑐, 𝑎𝑛𝑑 𝐹𝑙𝑜𝑤 𝑜𝑓 𝑖𝑛𝑡𝑒𝑟𝑛𝑎𝑙, 𝑘𝑖𝑛𝑒𝑡𝑖𝑐, 𝑎𝑛𝑑


𝑝𝑜𝑡𝑒𝑛𝑡𝑖𝑎𝑙 𝑒𝑛𝑒𝑟𝑔𝑦 𝑖𝑛𝑡𝑜 𝑠𝑦𝑠𝑡𝑒𝑚 − 𝑝𝑜𝑡𝑒𝑛𝑡𝑖𝑎𝑙 𝑒𝑛𝑒𝑟𝑔𝑦 𝑜𝑢𝑡 𝑜𝑓 𝑠𝑦𝑠𝑡𝑒𝑚
𝑏𝑦 𝑐𝑜𝑛𝑣𝑒𝑐𝑡𝑖𝑜𝑛 𝑜𝑟 𝑑𝑖𝑓𝑓𝑢𝑠𝑖𝑜𝑛 𝑏𝑦 𝑐𝑜𝑛𝑣𝑒𝑐𝑡𝑖𝑜𝑛 𝑜𝑟 𝑑𝑖𝑓𝑓𝑢𝑠𝑖𝑜𝑛
ℎ𝑒𝑎𝑡 𝑎𝑑𝑑𝑒𝑑 𝑡𝑜 𝑠𝑦𝑠𝑡𝑒𝑚 𝑏𝑦 𝑤𝑜𝑟𝑘 𝑑𝑜𝑛𝑒 𝑏𝑦 𝑠𝑦𝑠𝑡𝑒𝑚 𝑜𝑛 𝑠𝑢𝑟𝑟𝑜𝑢𝑛𝑑𝑖𝑛𝑔𝑠
+ −
𝑐𝑜𝑛𝑑𝑢𝑐𝑡𝑖𝑜𝑛, 𝑟𝑎𝑑𝑖𝑎𝑡𝑖𝑜𝑛 𝑎𝑛𝑑 𝑟𝑒𝑎𝑐𝑡𝑖𝑜𝑛 𝑠ℎ𝑎𝑓𝑡 𝑤𝑜𝑟𝑘 𝑎𝑛𝑑 𝑃𝑉 𝑤𝑜𝑟𝑘
𝑇𝑖𝑚𝑒 𝑟𝑎𝑡𝑒 𝑜𝑓 𝑐ℎ𝑎𝑛𝑔𝑒 𝑜𝑓 𝑖𝑛𝑡𝑒𝑟𝑛𝑎𝑙,
=
𝑘𝑖𝑛𝑒𝑡𝑖𝑐, 𝑝𝑜𝑡𝑒𝑛𝑡𝑖𝑎𝑙 𝑒𝑛𝑒𝑟𝑔𝑦 𝑖𝑛𝑠𝑖𝑑𝑒 𝑠𝑦𝑠𝑡𝑒𝑚
𝑑 𝜌𝜈 𝑉𝜈 𝐻 + 𝜌𝑉𝐿 ℎ
𝐹0 𝜌0 ℎ0 − 𝐹𝜌ℎ − 𝐹𝜈 𝜌𝜈 𝐻 + 𝑄 − 𝜆𝑉𝑘𝐶𝐴 =
𝑑𝑡
In order to express this equation explicitly in terms of temperature,
let us again use a very simple form for h (ℎ = 𝐶𝑃 𝑇) and an
equally simple form for H.
𝐻 = 𝐶𝑃 𝑇 + 𝜆𝜈
Where 𝜆𝜈 is an average heat of vaporization of the mixture. In a
more rigorous model 𝜆𝜈 , could be a function of temperature 𝑇𝜈 ,
composition y, and pressure P.
𝐹0 𝜌0 𝐶𝑃 𝑇0 − 𝐹𝜌𝐶𝑃 𝑇 − 𝐹𝜈 𝜌𝜈 𝐶𝑃 𝑇 + 𝜆𝜈 + 𝑄 − 𝜆𝑉𝑘𝐶𝐴

𝑑 𝜌𝜈 𝑉𝜈 𝐶𝑃 𝑇 + 𝜆𝜈 + 𝜌𝑉𝐿 𝐶𝑃 𝑇
=
𝑑𝑡
Problem 8. To illustrate the application of the energy equation to a
microscopic system, let us return to the plug-flow tubular reactor and now
keep track of temperature changes as the fluid flows down the pipe. Assume
no radial gradients in velocity, concentration, or temperature (a very poor
assumption in some strongly exothermic systems if the pipe diameter is not
kept small). Suppose that the reactor has a cooling jacket around it as shown
in Fig.. Heat can be transferred from the process fluid reactants and products
at temperature T to the metal wall of the reactor at temperature TM. The
heat is subsequently transferred to the cooling water. For a complete
description of the system we would need energy equations for the process
fluid, the metal wall, and the cooling water. Here we will concern ourselves
only with the process energy equation. 𝑇
𝑀(𝑡,𝑧 ) Water

𝐶𝐴0(𝑡)
𝜈 𝑡,𝑧 𝐶𝐴𝐿(𝑡)
𝑇0 𝑡
𝑇𝐿 𝑡
𝑇 𝑡,𝑧
൞𝐶𝐴 𝑡,𝑧
𝜌 𝑡,𝑧
Solution: Energy Balance Equation:
𝐹𝑙𝑜𝑤 𝑜𝑓 𝑖𝑛𝑡𝑒𝑟𝑛𝑎𝑙, 𝑘𝑖𝑛𝑒𝑡𝑖𝑐, 𝑎𝑛𝑑 𝐹𝑙𝑜𝑤 𝑜𝑓 𝑖𝑛𝑡𝑒𝑟𝑛𝑎𝑙, 𝑘𝑖𝑛𝑒𝑡𝑖𝑐, 𝑎𝑛𝑑
𝑝𝑜𝑡𝑒𝑛𝑡𝑖𝑎𝑙 𝑒𝑛𝑒𝑟𝑔𝑦 𝑖𝑛𝑡𝑜 𝑠𝑦𝑠𝑡𝑒𝑚 − 𝑝𝑜𝑡𝑒𝑛𝑡𝑖𝑎𝑙 𝑒𝑛𝑒𝑟𝑔𝑦 𝑜𝑢𝑡 𝑜𝑓 𝑠𝑦𝑠𝑡𝑒𝑚
𝑏𝑦 𝑐𝑜𝑛𝑣𝑒𝑐𝑡𝑖𝑜𝑛 𝑜𝑟 𝑑𝑖𝑓𝑓𝑢𝑠𝑖𝑜𝑛 𝑏𝑦 𝑐𝑜𝑛𝑣𝑒𝑐𝑡𝑖𝑜𝑛 𝑜𝑟 𝑑𝑖𝑓𝑓𝑢𝑠𝑖𝑜𝑛
ℎ𝑒𝑎𝑡 𝑎𝑑𝑑𝑒𝑑 𝑡𝑜 𝑠𝑦𝑠𝑡𝑒𝑚 𝑏𝑦 𝑤𝑜𝑟𝑘 𝑑𝑜𝑛𝑒 𝑏𝑦 𝑠𝑦𝑠𝑡𝑒𝑚 𝑜𝑛 𝑠𝑢𝑟𝑟𝑜𝑢𝑛𝑑𝑖𝑛𝑔𝑠
+ −
𝑐𝑜𝑛𝑑𝑢𝑐𝑡𝑖𝑜𝑛, 𝑟𝑎𝑑𝑖𝑎𝑡𝑖𝑜𝑛 𝑎𝑛𝑑 𝑟𝑒𝑎𝑐𝑡𝑖𝑜𝑛 𝑠ℎ𝑎𝑓𝑡 𝑤𝑜𝑟𝑘 𝑎𝑛𝑑 𝑃𝑉 𝑤𝑜𝑟𝑘
𝑇𝑖𝑚𝑒 𝑟𝑎𝑡𝑒 𝑜𝑓 𝑐ℎ𝑎𝑛𝑔𝑒 𝑜𝑓 𝑖𝑛𝑡𝑒𝑟𝑛𝑎𝑙,
=
𝑘𝑖𝑛𝑒𝑡𝑖𝑐, 𝑝𝑜𝑡𝑒𝑛𝑡𝑖𝑎𝑙 𝑒𝑛𝑒𝑟𝑔𝑦 𝑖𝑛𝑠𝑖𝑑𝑒 𝑠𝑦𝑠𝑡𝑒𝑚
Flow of energy (enthalpy) into boundary at z due to bulk flow : 𝜈𝐴𝜌𝐶𝑃 𝑇
𝑓𝑡 𝑙𝑏𝑚 𝐵𝑡𝑢 𝐵𝑡𝑢
with English engineering units of 𝑓𝑡 2
3 °𝑅 =
𝑚𝑖𝑛 𝑓𝑡 𝑙𝑏𝑚 °𝑅 𝑚𝑖𝑛
Flow of energy (enthalpy) out of boundary at z + dz: 𝑣𝐴𝜌𝐶
𝑇
𝑃 𝑇 + Wat
𝑀(𝑡,𝑧 )
𝜕 𝑣𝐴𝜌𝐶𝑝𝑇 er
𝑑𝑧 𝐶𝐴0(𝑡)
𝜕𝑧 𝑇0 𝑡
𝜈 𝑡,𝑧 𝐶𝐴𝐿(𝑡)

Heat generated by chemical reaction = −𝐴 𝑑𝑧 𝑘𝐶𝐴 𝜆 𝑇 𝑡,𝑧


𝑇𝐿 𝑡

Heat transferred to metal wall = −ℎ 𝑇 𝜋𝐷 𝑑𝑧 𝑇 – 𝑇𝑀


൞𝐶𝐴 𝑡,𝑧
𝜌 𝑡,𝑧

𝐵𝑡𝑢
𝑤ℎ𝑒𝑟𝑒 ℎ 𝑇 = ℎ𝑒𝑎𝑡 𝑡𝑟𝑎𝑛𝑠𝑓𝑒𝑟 𝑓𝑖𝑙𝑚 𝑐𝑜𝑒𝑓𝑓𝑖𝑐𝑖𝑒𝑛𝑡, 2
Heat conduction into boundary at 𝑧 = 𝑞𝑧 𝐴
where qZ is a heat flux in the z direction due to conduction.
We will use Fourier’s law to express qZ in terms of a temperature driving
force:
𝜕𝑇
𝑞𝑧 = −𝑘 𝑇
𝜕𝑧
where kT is an effective thermal conductivity with English engineering units
𝐵𝑡𝑢
of
min 𝑓𝑡 °𝑅
𝜕 𝑞𝑧 𝐴
Heat conduction out of boundary at 𝑧 + 𝑑𝑧 = 𝑞𝑧 𝐴 + 𝑑𝑧
𝜕𝑧
𝜕 𝜌𝐴𝑑𝑧𝐶𝑃 𝑇
Rate of change of internal energy (enthalpy) of the system =
𝜕𝑡
Combining all the above gives
𝜕 𝑣𝐴𝜌𝐶𝑝𝑇 𝜕 𝑞𝑧 𝐴
𝜈𝐴𝜌𝐶𝑃 𝑇 − 𝑣𝐴𝜌𝐶𝑃 𝑇 + 𝑑𝑧 + 𝑞𝑧 𝐴 − 𝑞𝑧 𝐴 + 𝑑𝑧
𝜕𝑧 𝜕𝑧
𝜕 𝜌𝐴𝑑𝑧𝐶𝑃 𝑇
+ −𝐴 𝑑𝑧 𝑘𝐶𝐴 𝜆 + −ℎ 𝑇 𝜋𝐷 𝑑𝑧 𝑇 – 𝑇𝑀 =
𝜕𝑡
𝜋𝐷2
Divide both sides with 𝐴𝑑𝑧, where 𝐴 𝑖𝑠 𝑐𝑟𝑜𝑠𝑠 𝑠𝑒𝑐𝑡𝑖𝑜𝑛𝑎𝑙 𝑎𝑟𝑒𝑎 =
4
𝜕 𝑣𝜌𝐶𝑝𝑇 𝜕 𝑞𝑧 ℎ𝑇 𝜕 𝜌𝐶𝑃 𝑇
− − − 𝑘𝐶𝐴 𝜆 − 2 (𝜋𝐷) 𝑇 – 𝑇𝑀 =
𝜕𝑧 𝜕𝑧 𝜋𝐷 𝜕𝑡
4

𝜕 𝑣𝜌𝐶𝑝𝑇 𝜕 𝑞𝑧 4ℎ 𝑇 𝜕 𝜌𝐶𝑃 𝑇
− − − 𝑘𝐶𝐴 𝜆 − 𝑇 – 𝑇𝑀 =
𝜕𝑧 𝜕𝑧 𝐷 𝜕𝑡
Substitute 𝑞𝑧 from Fourier’s law then rearrange
𝜕𝑇
𝜕 𝜌𝐶𝑃 𝑇 𝜕 𝑣𝜌𝐶𝑝𝑇 4ℎ𝑇 𝜕 𝑘𝑇
𝜕𝑧
+ + 𝑘𝐶𝐴 𝜆 + 𝑇 − 𝑇𝑀 =
𝜕𝑡 𝜕𝑧 𝐷 𝜕𝑧
Equations of Motion:
Newton’s second law of motion says that force is equal to mass times
acceleration for a system with constant mass M.
𝑀𝑎
𝐹=
𝑔𝑐
where F = force, lbf, M = mass, lbm, a = acceleration, ft/s2
gc = conversion constant needed when English engineering units are
used to keep units consistent = 32.2 lbm ft/lbf s2
This is the basic relationship that is used in writing the equations of
motion for a system.
In a slightly more general form, where mass can vary with time,
1 𝑑 𝑀𝜈𝑖
= σ𝑁𝑗=1 𝐹𝑗𝑖
𝑔𝑐 𝑑𝑡
where 𝜈𝑖 = velocity in the i direction, ft/s 𝐹𝑗𝑖 = jth force acting in
the i direction
Equation says that the time rate of change of momentum in the i
direction (mass times velocity in the i direction) is equal to the net sum
of the forces pushing in the i direction.
It can be thought of as a dynamic force balance. Or more eloquently it
is called the conservation of momentum.
In the real world there are three directions: x, y and z. Thus, three
force balances can be written for any system. Therefore, each system
has three equations of motion (plus one total mass balance, one energy
equation, and NC – 1 component balances).
Problem 9. Figure shows a tank into which an incompressible (constant
density) liquid is pumped at a variable rate F0 (ft3/s). This inflow rate
can vary with time because of changes in operations upstream. The
height of liquid in the vertical cylindrical tank is h (ft). The flow rate
out of the tank is F (ft3/s).
Now F0, h, and F will all vary with time and are therefore functions of
time t. Liquid leaves the base of the tank via a long horizontal pipe and
discharges into the top of another tank. 𝐹0
Both tanks are open to the atmosphere. Let the
length of the exit line be L(ft) and its cross-sectional
area be AP(ft2). The vertical, cylindrical tank has a
h
cross-sectional area of AT (ft ).
2
F
L F
Solution: Assume plug-flow conditions and incompressible liquid, and
therefore all the liquid is moving at the same velocity, more or less like
a solid rod. If the flow is turbulent, this is not a bad assumption.
Mass equal to the volume of the pipe (APL) times the density of the
liquid (ρ). 𝑀 = 𝐴𝑃 𝐿𝜌
Velocity 𝜈 (ft/s) equal to the volumetric flow divided by the cross-
𝐹
sectional area of the pipe. 𝑣 =
𝐴𝑃
𝐹0
The amount of liquid in the pipe will not change
with time, but if we want to change the rate of
outflow, the velocity of the liquid must be changed. h
And to change the velocity or the momentum of the F
F
liquid we must exert a force on the liquid. L
The direction of interest in this problem is the horizontal, since the pipe
is assumed to be horizontal. The force pushing on the liquid at the left
end of the pipe is the hydraulic pressure force of the liquid in the tank.
𝑔
𝐻𝑦𝑑𝑟𝑎𝑢𝑙𝑖𝑐 𝑓𝑜𝑟𝑐𝑒 = 𝐴𝑃 𝜌ℎ 𝐹 0
𝑔𝑐
The units of this force are
𝑓𝑡
32.2 2 h
2 𝑙𝑏𝑚
(in English engineering units):𝑓𝑡 3 𝑓𝑡 𝑠
𝑙𝑏 𝑓𝑡 = 𝑙𝑏𝑓 F
𝑓𝑡 32.2 𝑚 F
𝑙𝑏𝑓 𝑠2 L

The only force pushing in the opposite direction from right to left and
opposing the flow is the frictional force due to the viscosity of the
liquid.
If the flow is turbulent, the frictional force will be proportional to the
square of the velocity and the length of the pipe.
𝐹𝑟𝑖𝑐𝑡𝑖𝑜𝑛𝑎𝑙 𝑓𝑜𝑟𝑐𝑒 = 𝐾𝐹 𝐿𝜈 2
Substituting these forces into equation of motion, we get
1 𝑑 𝐴𝑃 𝐿𝜌𝜈 𝑔
= 𝐴𝑃 𝜌ℎ − 𝐾𝐹 𝐿𝜈 2
𝑔𝑐 𝑑𝑡 𝑔𝑐
𝐴𝑃 𝐿𝜌
AP , L, and ρ are constants, so divide both sides with
𝑔𝑐
𝑑𝜈 𝑔ℎ 𝐾𝐹 𝑔𝑐 𝜈2
= −
𝑑𝑡 𝐿 𝜌𝐴𝑃
The sign of the frictional force is negative because it acts in the direction
opposite the flow. We have defined left to right as the positive
direction.
Problem 10. Petroleum pipelines are sometimes used for transferring
several products from one location to another on a batch basis, i.e.,
one product at a time. To reduce product contamination at the end of
a batch transfer, a leather ball or “pig” that just fits the pipe is inserted
in one end of the line. Inert gas is introduced behind the pig to push it
through the line, thus purging the line of whatever liquid is in it.
Solution: To write a force balance on Pig
Liquid
the liquid still in the pipe as it is Inert
gas
𝑃0 𝝂

pushed out, we must take into 𝑧 Pipeline


𝐿
account the changing mass of
material.
Assume the pig is weightless and
frictionless compared with the liquid
in the line.
Let z be the axial position of the pig at any time. The liquid is
incompressible (density p) and flows in plug flow.
It exerts a frictional force proportional to the square of its velocity
and to the length of pipe still containing liquid.
𝐹𝑟𝑖𝑐𝑡𝑖𝑜𝑛𝑎𝑙 𝑓𝑜𝑟𝑐𝑒 = 𝐾𝐹 𝐿 − 𝑧 𝜈 2
The cross-sectional area of the pipe is 𝐴𝑃 .
The mass of fluid in the pipe is 𝐿 − 𝑧 𝐴𝑃 𝜌
The pressure P0 (lbr/ft2 gauge) of inert gas behind the pig is
essentially constant all the way down the pipeline.
The tank into which the liquid dumps is at atmospheric pressure.
The pipeline is horizontal. Pig
Liquid

Inert
𝑃0 𝝂
gas
𝑧 Pipeline

𝐿
A force balance in the horizontal z direction yields
1 𝑑 𝜌𝐴𝑃 𝜈 𝐿 − 𝑧
= 𝑃0 𝐴𝑃 − 𝐾𝐹 𝐿 − 𝑧 𝑣 2
𝑔𝑐 𝑑𝑡 Pig
Liquid
𝐴𝑃 𝜌
Divide both sides with and Inert
gas
𝑃0 𝝂
𝑔𝑐 𝑧 Pipeline

𝐿
𝑑𝑧
Substituting that 𝑣 = we get
𝑑𝑡
2
𝑑 𝑑𝑧 𝑃0 𝑔𝑐 𝑔𝑐 𝐾𝐹 𝑑𝑧
𝐿−𝑧 = − 𝐿−𝑧
𝑑𝑡 𝑑𝑡 𝜌 𝜌𝐴𝑃 𝑑𝑡
Transport Equations:
We have already used most of the laws governing the transfer of
energy, mass, and momentum.
These transport laws all have the form of a flux (rate of transfer per
unit area) being proportional to a driving force (a gradient in
temperature, concentration, or velocity).
The proportionality constant is a physical property of the system (like
thermal conductivity, diffusivity, or viscosity).
For transport on a molecular level, the laws bear the familiar names of
Fourier, Fick, and Newton.
Transfer relationships of a more macroscopic overall form are also used;
for example, film coefficients and overall coefficients in heat transfer.
Table summarizes some to the various relationships used in developing
models.
Quantity Heat Mass Momentum
Flux q 𝑁𝐴 𝜏𝑟𝑧
𝜕𝑇 𝜕𝐶𝐴 𝜕𝜈𝑧
Driving force
𝜕𝑧 𝜕𝑧 𝜕𝑟
Molecular transport
Law Fourier’s Fick’s Newton’s
Thermal Newton’s
Property Diffusivity (𝒟𝐴 )
conductivity 𝑘 𝑇 Viscosity(𝜇)
Overall transport
Driving force Δ𝑇 Δ𝐶𝐴 Δ𝑃
𝑔𝐷Δ𝑃
Relationship 𝑞 = ℎ 𝑇 Δ𝑇 𝑁𝐴 = 𝑘𝐿 ΔC 𝐿
𝑓=
2𝜌𝜈 2
Equations of State
To write mathematical models we need equations that tell us how the
physical properties, primarily density and enthalpy, change with
temperature, pressure, and composition.
𝐿𝑖𝑞𝑢𝑖𝑑 𝑑𝑒𝑛𝑠𝑖𝑡𝑦 = 𝜌𝐿 = 𝑓 𝑃,𝑇,𝑥𝑖
𝑉𝑎𝑝𝑜𝑟 𝑑𝑒𝑛𝑠𝑖𝑡𝑦 = 𝜌𝑣 = 𝑓 𝑃,𝑇,𝑦𝑖
𝐿𝑖𝑞𝑢𝑖𝑑 𝑒𝑛𝑡ℎ𝑎𝑙𝑝𝑦 = ℎ = 𝑓 𝑃,𝑇,𝑥𝑖
𝑉𝑎𝑝𝑜𝑟 𝑒𝑛𝑡ℎ𝑎𝑙𝑝𝑦 = 𝐻 = 𝑓 𝑃,𝑇,𝑦𝑖
Occasionally these relationships have to be fairly complex to describe
the system accurately.
But in many cases simplification can be made without sacrificing much
overall accuracy.
We have already used some simple enthalpy equations in the examples
of energy balances.
ℎ = 𝐶𝑃 𝑇
𝐻 = 𝐶𝑃 𝑇 + 𝜆𝜈
The next level of complexity would be to make the 𝐶𝑃 ‘𝑠 functions of
𝑇
temperature: h = ‫𝑇𝑑 𝑇 𝑃𝐶 𝑇׬‬
0
A polynomial in T is often used for 𝐶𝑃 : 𝐶𝑃 𝑇 = 𝐴1 + 𝐴2 𝑇
Then Eq. becomes
𝑇
𝑇2 𝐴2
ℎ = 𝐴1 𝑇 + 𝐴2 = 𝐴1 𝑇 − 𝑇0 + 𝑇 2 − 𝑇02
2 𝑇0 2
Of course, with mixtures of components the total enthalpy is needed.
If heat of mixing effects are negligible, the pure-component enthalpies
can be averaged:
σ𝑁𝐶
𝑗=1 𝑥𝑗 ℎ𝑗 𝑀𝑗
ℎ= σ𝑁𝐶
𝑗=1 𝑥𝑗 𝑀𝑗
where xj= mole fraction of jth component
Mj = molecular weight of jth component
hj = pure-component enthalpy of jth component, energy per unit mass
Liquid densities can be assumed constant in many systems unless large
changes in composition and temperature occur.
Vapor densities usually cannot be considered invariant and some sort of PVT
relationship is almost always required. The simplest and most often used is
the perfect-gas law : 𝑃𝑉 = 𝑛𝑅𝑇
where P = absolute pressure (lbf/ft2 or kilopascals) n = number of moles (lb.
mol or kg. mol)
R = constant = 1545 lbf ft/lb. mol °R or 8.314 kPa m3/kg .mol K
T = absolute temperature (°R or K) V = volume (ft3 or m3)
Rearranging to get an equation for density ρv (lbm/ft3 or kg/m3) of a perfect
gas with a molecular weight M, we get
𝑛𝑀 𝑀𝑃
𝜌𝜈 = =
𝑉 𝑅𝑇
EQUILIBRIUM
The second law of thermodynamics is the basis for the equations that
tell us the conditions of a system when equilibrium conditions prevail.
A. CHEMICAL EQUILIBRIUM. Equilibrium occurs in a reacting system
when
σ𝑁𝐶
𝑗=1 𝜈𝑗 𝜇𝑗 = 0
where vj = stoichiometric coefficient of the jth component with
reactants having a negative sign and products a positive sign, 𝜇𝑗 =
chemical potential of jth component
The usual way to work with this equation is in terms of an equilibrium
constant for a reaction.
For example, consider a reversible gas-phase reaction of A to form B at
a specific rate k1 and B reacting back to A at a specific reaction rate k2.
The stoichiometry of the reaction is such that va moles of A react to form vb
𝑘1
moles of B. 𝜈𝑎 𝐴 ֞ 𝜈𝑏 𝐵
𝑘2
Equation says equilibrium will occur when 𝑣𝑏 𝜇𝐵 − 𝑣𝑎 𝜇𝐴 = 0
The chemical potentials for perfect gas mixture can be written
𝜇𝑗 = 𝜇𝑗0 + 𝑅𝑇 ln 𝑝𝑗
where 𝜇𝑗0 = standard chemical potential (or Gibbs free energy per mole) of
the jth component, which is a function of temperature only
pj = partial pressure of the jth component, R = perfect-gas law constant,
T = absolute temperature
Substituting into Eq. 𝑣𝑏 (𝜇𝐵0 + 𝑅𝑇 ln 𝑝𝐵 ) − 𝑣𝑎 (𝜇𝐴0 + 𝑅𝑇 ln 𝑝𝐴 ) = 0
𝑅𝑇 ln 𝑝𝐵 𝑣𝑏 − 𝑅𝑇 ln 𝑝𝐴 𝑣𝑎 = 𝑣𝑎 𝜇𝐴0 − 𝑣𝑏 𝜇𝐵0
𝜈𝑏
𝑝𝐵 (𝑣𝑎 𝜇𝐴0 − 𝑣𝑏 𝜇𝐵0 )
ln 𝜈𝑎 =
𝑝𝐴 𝑅𝑇
The right-hand side of this equation is a function of temperature only.
The term in parenthesis on the left-hand side is defined as the equilibrium
constant KP and it tells us the equilibrium ratios of products and reactants.
𝜈
𝑝𝐵𝑏
𝐾𝑃 ≡ 𝜈𝑎
𝑝𝐴
B.PHASE EQUILIBRIUM: Equilibrium between two phases occurs when
the chemical potential of each component is the same in the two phases:
𝜇𝑗𝐼 = 𝜇𝑗𝐼𝐼
where 𝜇𝑗𝐼 = chemical potential of the jth component in phase I
𝜇𝑗𝐼𝐼 = chemical potential of the jth component in phase II
Since the vast majority of chemical engineering systems involve liquid and
vapor phases, many vapor-liquid equilibrium relationships are used.
They range from the very simple to the very complex. Some of the most
commonly used relationships are listed below.
Basically we need a relationship that permits us to calculate the vapor
composition if we know the liquid composition, or vice versa.
The most common problem is a bubble point calculation: calculate the
temperature T and vapor composition yj, given the pressure P and the
liquid composition xi.
This usually involves a trial-and-error, iterative solution because the
equations can be solved explicitly only in the simplest cases.
Sometimes we have bubble point calculations that start from known
values of xi and T and want to find P and yj.
This is frequently easier than when pressure is known because the
bubble point calculation is usually noniterative.
Dewpoint calculations must be made when we know the composition
of the vapor yi and P (or) T and want to find the liquid composition x,
and T (or P).
Flash calculations must be made when we know neither xj nor yj and
must combine phase equilibrium relationships, component balance
equations, and an energy balance to solve for all the unknowns.
We will assume ideal vapor-phase behavior in our examples, i.e., the
partial pressure of the jth component in the vapor is equal to the total
pressure P times the mole fraction of the jth component in the vapor yj
(Dalton’s law): 𝑝𝑗 = 𝑃𝑦𝑗
Corrections may be required at high pressures.
In the liquid phase several approaches are widely used.
1. Raoult’s law.: Liquids that obey Raoult’s are called ideal.
𝑁𝐶 𝑆
σ
𝑝𝑗 = 𝑗=1 𝑥𝑗 𝑃𝑗
𝑥𝑗 𝑃𝑗𝑆
𝑦𝑗 =
𝑃
where 𝑃𝑗𝑆 is the vapor pressure of pure component j.
Vapor pressures are functions of temperature only.
This dependence is often described by
𝑠
𝐴𝑗
ln 𝑃𝑗 = + 𝐵𝑗
𝑇
2. Relative volatility. The relative volatility 𝛼𝑖𝑗 of component i to
component j is defined :
𝑦𝑖
ൗ𝑥𝑖
𝛼𝑖𝑗 = 𝑦𝑗
ൗ𝑥𝑗

Relative volatilities are fairly constant in a number of systems.


They are convenient so they are frequently used.
In a binary system the relative volatility α of the more volatile
component compared with the less volatile component is
𝑦ൗ
𝑥
𝛼=
1−𝑦 / 1−𝑥
Rearranging,
𝛼𝑥
𝑦=
1+ 𝛼−1 𝑥
3. K values. Equilibrium vaporization ratios or K values are widely
used, particularly in the petroleum industry.
𝑦𝑗
𝐾𝑗 =
𝑥𝑗

The K’s are functions of temperature and composition, and to a lesser


extent, pressure.
4. Activity coefficients. For nonideal liquids, Raoult’s law must be
modified to account for the nonideality in the liquid phase. The “fudge
factors” used are called activity coefficients.
𝑝𝑗 = σ𝑁𝐶 𝑠
𝑗=1 𝑗 𝑗 𝛾𝑗
𝑥 𝑃
where 𝛾𝑗 is the activity coefficient for the jth component.
The activity coefficient is equal to 1 if the component is ideal.
The 𝛾’s are functions of composition and temperature.
Chemical Kinetics
We will be modeling many chemical reactors, and we must be familiar
with the basic relationships and terminology used in describing the
kinetics (rate of reaction) of chemical reactions. For more details,
consult one of the several excellent texts in this field.
A. ARRHENIUS TEMPERATURE DEPENDENCE. The effect of
temperature on the specific reaction rate k is usually found to be
exponential :
𝐸

𝑘 = 𝑘0 𝑒
𝑅𝑇

where k = specific reaction rate


k0 = pre exponential factor
E = activation energy; shows the temperature dependence of the
reaction rate, i.e., the bigger E, the faster the increase in k with
increasing temperature (Btu/lb .mol or Cal/g .mol)
T = absolute temperature
R = perfect-gas constant = 1.99 Btu/lb. mol °R or 1.99 Cal/g. mol K
This exponential temperature dependence represents one of the most
severe nonlinearities in chemical engineering systems.
Keep in mind that the “apparent” temperature dependence of a
reaction may not be exponential if the reaction is mass-transfer limited,
not chemical-rate limited.
If both zones are encountered in the operation of the reactor, the
mathematical model must obviously include both reaction-rate and
mass-transfer effects.
B. LAW OF MASS ACTION. Using the conventional notation, we will
define an overall reaction rate ℛ as the rate of change of moles of any
component per volume due to chemical reaction divided by that
component’s stoichiometric coefficient.
1 𝑑𝑛𝑗
ℛ=
𝑣𝑗 𝑉 𝑑𝑡 𝑅
The stoichiometric coefficients vj are positive for products of the
reaction and negative for reactants.
Note that ℛ is an intensive property and can be applied to systems of
any size.
For example, assume we are dealing with an irreversible reaction in
which components A and B react to form components C and D.
𝑘
𝑣𝑎 𝐴 + 𝑣𝑏 𝐵՜𝑣𝑐 𝐶 + 𝑣𝑑 𝐷
Then
1 𝑑𝑛𝐴 1 𝑑𝑛𝐵 1 𝑑𝑛𝐶 1 𝑑𝑛𝐷
ℛ= = = =
−𝑣𝑎 𝑉 𝑑𝑡 𝑅 −𝑣𝑏 𝑉 𝑑𝑡 𝑅 𝑣𝑐 𝑉 𝑑𝑡 𝑅 𝑣𝑑 𝑉 𝑑𝑡 𝑅
The law of mass action says that the overall reaction rate ℛ will vary
with temperature (since k is temperature-dependent) and with the
concentration of reactants raised to some powers.
ℛ = 𝑘 𝑇 𝐶𝐴 𝑎 𝐶𝐵 𝑏
where CA = Concentration of component A
CB = Concentration of component B
The constants a and b are not, in general, equal to the stoichiometric
coefficients va and vb.
The reaction is said to be first-order in A if a = 1. It is second-order in A
if a = 2. The constants a and b can be fractional numbers.
As indicated earlier, the units of the specific reaction rate k depend on
the order of the reaction. This is because the overall reaction rate %
always has the same units (moles per unit time per unit volume).
For a first-order reaction of A reacting to form B, the overall reaction
rate ℛ, written for component A, would have units of moles of A/min
ft3.
ℛ = 𝑘𝐶𝐴
If CA has units of moles of A/ft3, k must have units of min-1.
If the overall reaction rate for the system above is second-order in A,
ℛ = 𝑘𝐶𝐴2
ℛ 𝑠till has units of moles of A/min ft3. Therefore, k must have units of
ft3/min mol A.
Consider the reaction 𝐴 + 𝐵 ՜ 𝐶. If the overall reaction rate is first-
order in both A and B,
ℛ = 𝑘𝐶𝐴 𝐶𝐵
ℛ still has units of moles of A/min ft3. Therefore k must have units of
ft3/min mol B
Dalton’s Law for vapor 𝑝𝑗 = 𝑃𝑦𝑗 Algorithm:
Raoult’s Law for Liquid 𝑝𝑗 = 𝑥𝑗 𝑃𝑗𝑆  Start

When vapor and liquid are in equilibrium:  Read temperature ‘T’.


𝑦𝑖 𝑃 = 𝑥𝑖 𝑝𝑖𝑠  Read Antoine constants 𝐴1 , 𝐵1 , 𝐶1 , 𝐴2 , 𝐵2 , 𝐶2 .

Antoine Equation: ln 𝑝𝑖𝑠 = 𝐴𝑖 −


𝐵𝑖
 Calculate vapour pressures of component 1 and 2
𝑇+𝐶𝑖
𝐵1
 𝑝1𝑠 = 𝑒𝑥𝑝 𝐴1 −
𝐶1 +𝑇
P-X-Y DATA FOR IDEAL MIXTURES USING  𝑝2𝑠 = 𝑒𝑥𝑝 𝐴2 −
𝐵2
𝐶2 +𝑇
BUBBLE PRESSURE (𝒇𝒐𝒓 𝒌𝒏𝒐𝒘𝒏 𝑻 𝒂𝒏𝒅 𝒙𝒊 )
 Calculate total pressure at different x1 values
For binary mixture 𝑦1 𝑃 = 𝑥1 𝑝1𝑠  for x1=0 to 1.0 insteps of 0.1 do
𝑦2 𝑃 = 𝑥2 𝑝2𝑠  𝑃 = 𝑥1 𝑝1𝑠 + (1 − 𝑥1 )𝑝2𝑠
Adding above two equation 𝑥1 𝑝1𝑠
 𝑦1 =
𝑃
(𝑦1 +𝑦2 )𝑃 = 𝑥1 𝑝1𝑠 + 𝑥2 𝑝2𝑠
𝑥2 = 1 − 𝑥1

𝑃 = 𝑥1 𝑝1𝑠 + 𝑥2 𝑝2𝑠  𝑦2 = 1 − 𝑦1
𝑥2 = 1 − 𝑥1  Display the P-x-y data at Bubble pressure.
𝑃 = 𝑥1 𝑝1𝑠 + (1 − 𝑥1 )𝑝2𝑠  end for
 End.
P-X-Y DATA FOR IDEAL MIXTURES USING DEW Algorithm:
PRUSSURE (𝒇𝒐𝒓 𝒌𝒏𝒐𝒘𝒏 𝑻 𝒂𝒏𝒅 𝒚𝒊 )  Start

From Roult’s law  Read temperature ‘T’.


 Read Antoine constants 𝐴1 , 𝐵1 , 𝐶1 , 𝐴2 , 𝐵2 , 𝐶2 .
𝑦1 𝑃
𝑦1 𝑃 = 𝑥1 𝑝1𝑠 −՜ 𝑥1 = 𝑠  Calculate vapour pressures of component 1 and 2
𝑝1 𝐵1
 𝑝1𝑠 = 𝑒𝑥𝑝 𝐴1 −
𝐶1 +𝑇
𝑦2 𝑃
𝑦2 𝑃 = 𝑥2 𝑝2𝑠
−՜ 𝑥2 = 𝑠  𝑝2𝑠 = 𝑒𝑥𝑝 𝐴2 −
𝐵2
𝑝2 𝐶2 +𝑇
Adding above two equations  Calculate total pressure at different y1 values
𝑦1 𝑃 𝑦2 𝑃  for y1=0 to 1.0 insteps of 0.1 do
𝑥1 + 𝑥2 = 𝑠 + 𝑠
𝑝1 𝑝2 1
 𝑃= 𝑦1 𝑃 𝑦2 𝑃
+ 𝑠
1 𝑝𝑠1 𝑝2
𝑃= 𝑦1 𝑃
𝑦1 𝑦2  𝑥1 =
𝑝1𝑠
𝑠+ 𝑠
𝑝1 𝑝2  𝑥2 = 1 − 𝑥1
𝑦2 = 1 − 𝑦1
 𝑦2 = 1 − 𝑦1
The Liquid composition is found using
𝑦1𝑃 𝑦2𝑃  Display the P-x-y data at Dew pressure
formula 𝑥1 = 𝑠 and 𝑥2 = 𝑠  end for
𝑝1 𝑝2
 End.
Algorithm:
T-X-Y DATA FOR IDEAL MIXTURES USING  Start
BUBBLE TEMPERATURE (𝒇𝒐𝒓 𝒌𝒏𝒐𝒘𝒏 𝑷 𝒂𝒏𝒅 𝒙𝒊 )  Read pressure ‘P’.
𝑏𝑖
𝑡𝑖 = − 𝑐𝑖  Read Antoine constants 𝐴1 , 𝐵1 , 𝐶1 , 𝐴2 , 𝐵2 , 𝐶2.
𝑎𝑖 − ln 𝑝𝑖𝑠
𝑠  Calculate boiling points of component 1 and 2
𝑠 𝑥1 𝑝1
𝑃= 𝑥1 𝑝1𝑠 + 𝑥2 𝑝2𝑠 = 𝑝2 + 𝑥2
𝑝2𝑠 𝐵1
𝑃  𝑡1 = − 𝐶1
𝑝2𝑠 = 𝑥1 𝑝𝑠1
𝐴1 −ln 𝑝1𝑠
+𝑥2 𝐵2
𝑝𝑠2
𝑝1𝑠
 𝑡2 = − 𝐶2
𝑃 𝐴2 −ln 𝑝2𝑠
𝑝2𝑠 = where ∝=
𝑥1 ∝+𝑥2 𝑝2𝑠
 for x1=0 to 1.0 insteps of 0.1 do
𝑏1
We know that ln 𝑝1𝑠 = 𝑎1 −
(𝑡+𝑐1 )  𝑡𝑔 = 𝑥1 ∗ 𝑡1 + 𝑥2 ∗ 𝑡2
𝑏2 𝐵1 𝐵2
ln 𝑝2𝑠 = 𝑎2 −  ∝ = 𝑒𝑥𝑝 𝐴 − − 𝐴2 +
(𝑡 + 𝑐2 ) (𝑡𝑔 +𝐶1 ) (𝑡𝑔 +𝐶2 )
𝑃
𝑏1 𝑏2  𝑝2𝑠 = 𝑥1 𝑝𝑠1
ln 𝑝1𝑠 − ln 𝑝2𝑠 = 𝑎1 − − 𝑎2 + +𝑥2
(𝑡 + 𝑐1 ) (𝑡 + 𝑐2 ) 𝑝𝑠2
𝐵2
𝑝1𝑠 𝑏1 𝑏2  𝑡𝑛 = − 𝐶2
𝑙𝑛 𝑠 = ln ∝ = 𝑎1 − − 𝑎2 + 𝐴2 −ln 𝑝2𝑠
𝑝2 (𝑡 + 𝑐1 ) (𝑡 + 𝑐2 )  𝑥2 = 1 − 𝑥1
𝑏1 𝑏2 𝑥2 ∗𝑝2𝑠
∝ = 𝑒𝑥𝑝 𝑎1 − − 𝑎2 +  𝑦2 =
(𝑡 + 𝑐1 ) (𝑡 + 𝑐2 ) 𝑃
𝑦1 = 1 − 𝑦2

𝑏2
𝑡𝑛 = − 𝑐2  Display the T-x-y data at Bubble temperature
𝑎2 − ln 𝑝2𝑠  end for
 End.
Algorithm:
T-X-Y DATA FOR IDEAL MIXTURES USING DEW  Start
TEMPERATURE (𝒇𝒐𝒓 𝒌𝒏𝒐𝒘𝒏 𝑷 𝒂𝒏𝒅 𝒚𝒊 )
 Read pressure ‘P’.
𝑏𝑖
 𝑡𝑖 = − 𝑐𝑖  Read Antoine constants 𝐴1 , 𝐵1 , 𝐶1 , 𝐴2 , 𝐵2 , 𝐶2.
𝑎𝑖 −ln 𝑝𝑖𝑠
𝑦1 𝑃
 𝑦1 𝑃 = 𝑥1 𝑝1𝑠 −՜ 𝑥1 =  Calculate boiling points of component 1 and 2
𝑝1𝑠
𝑦 𝑃 𝐵1
 𝑦2 𝑃 = 𝑥2 𝑝2𝑠 −՜ 𝑥2 = 2𝑠  𝑡1 = − 𝐶1
𝑝2 𝐴1 −ln 𝑝1𝑠
𝑦1 𝑃 𝑦2 𝑃
 Adding above two equations 𝑥1 + 𝑥2 = + 𝐵2
𝑝1𝑠 𝑝2𝑠
1  𝑡2 = 𝐴2 −ln 𝑝2𝑠
− 𝐶2
 𝑃= 𝑦1 𝑦2
+  for y1=0 to 1.0 insteps of 0.1 do
𝑝𝑠1 𝑝𝑠2
1
 𝑃= 1 𝑦 𝑝 𝑠  𝑡𝑔 = 𝑥1 ∗ 𝑡1 + 𝑥2 ∗ 𝑡2
𝑦1 + 2 𝑠 1
𝑝𝑠1 𝑝 2
𝑝1𝑠 𝐵1 𝐵2
 𝑝1𝑠 = 𝑃 𝑦1 + ∝ 𝑦2 where ∝=  ∝ = 𝑒𝑥𝑝 𝐴1 − (𝑡𝑔 +𝐶1 )
− 𝐴2 + (𝑡𝑔 +𝐶2 )
𝑝2𝑠
𝑏1
 We know that ln 𝑝1𝑠 = 𝑎1 −
(𝑡+𝑐1 )  𝑝1𝑠 = 𝑃 𝑦1 + ∝ 𝑦2
𝑏2
ln 𝑝2𝑠 = 𝑎2 − 𝐵1
(𝑡 + 𝑐2 )  𝑡𝑛 = − 𝐶1
𝑏1 𝑏2 𝐴1 −ln 𝑝1𝑠
 ln 𝑝1𝑠 − ln 𝑝2𝑠 = 𝑎1 − − 𝑎2 +
(𝑡+𝑐1 ) (𝑡+𝑐2 )
𝑝1𝑠 𝑏1 𝑏2
 𝑥2 = 1 − 𝑥1
 𝑙𝑛 = ln ∝ = 𝑎1 − − 𝑎2 + 𝑥1 ∗𝑝1𝑠
𝑝2𝑠 (𝑡+𝑐1 ) (𝑡+𝑐2 )
𝑏1 𝑏2
 𝑦1 = 𝑃
 ∝ = 𝑒𝑥𝑝 𝑎1 − − 𝑎2 +
(𝑡+𝑐1 ) (𝑡+𝑐2 ) 𝑦2 = 1 − 𝑦1

𝐵1  Display the T-x-y data at Dew temperature
 𝑡𝑛 = 𝐴1 −ln 𝑝1𝑠
− 𝐶1
 end for
 End.
UNIT-I
Objectives:

1. To use the fundamental laws in developing model equations.

2. To understand various chemical engineering systems.

Outcome: By the end of the unit the student will be able to apply the
fundamental laws to develop a mathematical model for simple flow
systems.
THANK YOU
PROCESS MODELING
AND SIMULATION
UNIT -II
MATHEMATICAL MODELING-I
By
Mr. Koteswara Rao Maradana
Assistant Professor
Dept. of Chemical Engineering
ANIS
UNIT-II

Objectives:
1. To understand various chemical engineering systems.
2. To develop mathematical models for solving process problems.
Gravity flow tank:
Figure shows a tank into which an incompressible (constant density)
liquid is pumped at a variable rate F0 (ft3/s). This inflow rate can vary
with time because of changes in operations upstream. The height of
liquid in the vertical cylindrical tank is h (ft). The flow rate out of the
tank is F (ft3/s).
Now F0, h, and F will all vary with time and are therefore functions of
time t. Liquid leaves the base of the tank via a long horizontal pipe and
discharges into the top of another tank. 𝐹0

Both tanks are open to the atmosphere. Let the


length of the exit line be L(ft) and its cross-sectional
area be AP(ft2). The vertical, cylindrical tank has a h
F
cross-sectional area of AT (ft2). F
L
Solution: Assume plug-flow conditions and incompressible liquid, and
therefore all the liquid is moving at the same velocity, more or less like
a solid rod. If the flow is turbulent, this is not a bad assumption.
Mass equal to the volume of the pipe (APL) times the density of the
liquid (ρ). 𝑀 = 𝐴𝑃 𝐿𝜌
Velocity 𝜈 (ft/s) equal to the volumetric flow divided by the cross-
𝐹
sectional area of the pipe. 𝑣 =
𝐴𝑃

The amount of liquid in the pipe will not change 𝐹0

with time, but if we want to change the rate of


outflow, the velocity of the liquid must be changed.
h
And to change the velocity or the momentum of the F
liquid we must exert a force on the liquid. L
F
The direction of interest in this problem is the horizontal, since the pipe
is assumed to be horizontal. The force pushing on the liquid at the left
end of the pipe is the hydraulic pressure force of the liquid in the tank.
𝑔
𝐻𝑦𝑑𝑟𝑎𝑢𝑙𝑖𝑐 𝑓𝑜𝑟𝑐𝑒 = 𝐴𝑃 𝜌ℎ 𝐹 0
𝑔𝑐
The units of this force are
𝑓𝑡
32.2 2 h
2 𝑙𝑏𝑚
(in English engineering units):𝑓𝑡 3 𝑓𝑡 𝑠
𝑙𝑏 𝑓𝑡 = 𝑙𝑏𝑓 F
𝑓𝑡 32.2 𝑚 F
𝑙𝑏𝑓 𝑠2 L

The only force pushing in the opposite direction from right to left and
opposing the flow is the frictional force due to the viscosity of the
liquid.
If the flow is turbulent, the frictional force will be proportional to the
square of the velocity and the length of the pipe.
𝐹𝑟𝑖𝑐𝑡𝑖𝑜𝑛𝑎𝑙 𝑓𝑜𝑟𝑐𝑒 = 𝐾𝐹 𝐿𝜈 2
Substituting these forces into equation of motion, we get
1 𝑑 𝐴𝑃 𝐿𝜌𝜈 𝑔
= 𝐴𝑃 𝜌ℎ − 𝐾𝐹 𝐿𝜈 2
𝑔𝑐 𝑑𝑡 𝑔𝑐
𝐴𝑃 𝐿𝜌
AP , L, and ρ are constants, so divide both sides with
𝑔𝑐
𝑑𝜈 𝐾𝐹 𝑔𝑐 𝜈 2 𝑔ℎ
+ =
𝑑𝑡 𝜌𝐴𝑃 𝐿
The sign of the frictional force is negative because it acts in the direction
opposite the flow. We have defined left to right as the positive
direction.
𝑑𝑉𝜌
𝐹0 𝜌 − 𝐹𝜌 =
𝑑𝑡
𝑑(𝐴𝑃 ℎ)
𝐹0 − 𝐹 =
𝑑𝑡
SERIES OF ISOTHERMAL CONSTANT-HOLDUP CSTRs
 The system is sketched in Fig. and is a simple extension of the CSTR
considered. Product B is produced and reactant A is consumed in each
of the three perfectly mixed reactors by a first-order reaction occurring
in the liquid. For the moment let us assume that the temperatures and
holdups (volumes) of the three tanks can be different, but both
temperatures and the liquid volumes are assumed to be constant
(isothermal and constant holdup). Density is assumed constant
throughout the system, which is a binary mixture of A and B.
𝑘
(𝐴 ՜ 𝐵)
𝐹0
𝐶𝐴0

𝑉1 𝐹1 𝐹2 𝑉3 𝐹3
𝑉2
𝑘1 𝐶𝐴1 𝑘2 𝐶𝐴2 𝑘3
𝐶𝐴3
 With these assumptions in mind, we are ready to formulate our
model. If the volume and density of each tank are constant, the total
mass in each tank is constant.
 The total mass balance(continuity) equation for the first reactor is
𝑀𝑎𝑠𝑠 𝑓𝑙𝑜𝑤 𝑀𝑎𝑠𝑠 𝑓𝑙𝑜𝑤 𝑇𝑖𝑚𝑒 𝑟𝑎𝑡𝑒 𝑜𝑓 𝑐ℎ𝑎𝑛𝑔𝑒
− =
𝑖𝑛𝑡𝑜 𝑠𝑦𝑠𝑡𝑒𝑚 𝑜𝑢𝑡 𝑜𝑓 𝑡ℎ𝑒 𝑠𝑦𝑠𝑡𝑒𝑚 𝑜𝑓 𝑚𝑎𝑠𝑠 𝑖𝑛𝑠𝑖𝑑𝑒 𝑠𝑦𝑠𝑡𝑒𝑚
𝐹0
𝑑 𝜌𝑉1
= 𝜌𝐹0 − 𝜌𝐹1 = 0 (𝑜𝑟) 𝐹0 = 𝐹1 𝐶𝐴0
𝑑𝑡
𝑑 𝜌𝑉2 𝑉1 𝐹1 𝐹2 𝐹3
= 𝜌𝐹1 − 𝜌𝐹2 = 0 (𝑜𝑟)
𝑉3
𝐹1 = 𝐹2 𝑘1
𝐶𝐴1
𝑉2
𝑘2 𝐶𝐴2
𝑘3
𝐶𝐴3
𝑑𝑡
𝑑 𝜌𝑉3
= 𝜌𝐹2 − 𝜌𝐹3 = 0 (𝑜𝑟) 𝐹2 = 𝐹3
𝑑𝑡
Therefore
𝐹0 = 𝐹1 = 𝐹2 = 𝐹3 = 𝐹
where F is defined as the throughput (m3/min).
 We want to keep track of the amounts of reactant A and product B
in each tank, so component continuity equations are needed.
 If we arbitrarily choose A, the equations describing the dynamic
changes in the amounts of reactant A in each tank are (with units of
kg - mol of A/min)
 Component A Balance:
𝐹𝑙𝑜𝑤 𝑜𝑓 𝑚𝑜𝑙𝑒𝑠 𝑜𝑓 𝐹𝑙𝑜𝑤 𝑜𝑓 𝑚𝑜𝑙𝑒𝑠 𝑜𝑓
 − +
𝑐𝑜𝑚𝑝𝑜𝑛𝑒𝑛𝑡 𝐴 𝑖𝑛𝑡𝑜 𝑠𝑦𝑠𝑡𝑒𝑚 𝑐𝑜𝑚𝑝𝑜𝑛𝑒𝑛𝑡 𝐴 𝑜𝑢𝑡 𝑜𝑓 𝑡ℎ𝑒 𝑠𝑦𝑠𝑡𝑒𝑚
𝑟𝑎𝑡𝑒 𝑜𝑓 𝑓𝑜𝑟𝑚𝑎𝑡𝑖𝑜𝑛 𝑜𝑓 𝑚𝑜𝑙𝑒𝑠 𝑜𝑓 𝑇𝑖𝑚𝑒 𝑟𝑎𝑡𝑒 𝑜𝑓 𝑐ℎ𝑎𝑛𝑔𝑒 𝑜𝑓 𝑚𝑜𝑙𝑒𝑠 𝑜𝑓
=
𝑐𝑜𝑚𝑝𝑜𝑛𝑒𝑛𝑡 𝐴 𝑓𝑟𝑜𝑚 𝑐ℎ𝑒𝑚𝑖𝑐𝑎𝑙 𝑟𝑒𝑎𝑐𝑡𝑖𝑜𝑛𝑠 𝑐𝑜𝑚𝑝𝑜𝑛𝑒𝑛𝑡 𝐴 𝑖𝑛𝑠𝑖𝑑𝑒 𝑠𝑦𝑠𝑡𝑒𝑚
𝑑𝐶𝐴1
𝐹 𝐶𝐴0 − 𝐶𝐴1 − 𝑉1 𝐾1 𝐶𝐴1 = 𝑉1
𝑑𝑡
𝑑𝐶𝐴2
𝐹 𝐶𝐴1 − 𝐶𝐴2 − 𝑉2 𝐾2 𝐶𝐴2 = 𝑉2
𝑑𝑡
𝑑𝐶𝐴3
𝐹 𝐶𝐴2 − 𝐶𝐴3 − 𝑉3 𝐾3 𝐶𝐴3 = 𝑉3
𝑑𝑡
 The specific reaction rates 𝑘𝑛 are given by the Arrhenius equation
𝐸

𝑘𝑛 = 𝑘0 𝑒 𝑅𝑇𝑛 n = 1, 2, 3
 If the temperatures in the reactors are different, the k’s are different.
The n refers to the reactor number.
 The three first-order nonlinear ordinary differential equations given
are the mathematical model of the system.
 The parameters that must be known are 𝑉1 , 𝑉2 , 𝑉3 , 𝑘1 , 𝑘2 , 𝑎𝑛𝑑 𝑘3 .
 The variables that must be specified before these equations can be
solved are F and CA0.
 “Specified”
does not mean that they must be constant. They can be
time-varying, but they must be known or given functions of time.
They are the forcing functions.
 The initial conditions of the three concentrations (their values at time
equal zero) must also be known.
 Let us now check the degrees of freedom of the system.
 There are three equations and, with the parameters and forcing
functions specified, there are only three unknowns or dependent
variables: CA1, CA2, and CA3.
 Simple system used for controller design and stability analysis, we will
use an even simpler version.
 Ifthe throughput F is constant and the holdups and temperatures are
the same in all three tanks
𝑉
 where 𝜏 = with units of minutes.
𝐹
 The above equations become
𝑑𝐶𝐴1 𝑑𝐶𝐴1 1 1
𝐹 𝐶𝐴0 − 𝐶𝐴1 − 𝑉1 𝐾1 𝐶𝐴1 = 𝑉1 ՜ + 𝑘+ 𝐶𝐴1 = 𝐶𝐴0
𝑑𝑡 𝑑𝑡 𝜏 𝜏
𝑑𝐶𝐴2
𝐹 𝐶𝐴1 − 𝐶𝐴2 − 𝑉2 𝐾2 𝐶𝐴2 = 𝑉2 ՜ 𝑑𝐶𝐴2 + 𝑘 + 1 𝐶𝐴2 = 1 𝐶𝐴1
𝑑𝑡 𝑑𝑡 𝜏 𝜏
𝑑𝐶𝐴3
𝐹 𝐶𝐴2 − 𝐶𝐴3 − 𝑉3 𝐾3 𝐶𝐴3 = 𝑉3 ՜ 𝑑𝐶𝐴3 1 1
𝑑𝑡 + 𝑘+ 𝐶𝐴3 = 𝐶𝐴2
𝑑𝑡 𝜏 𝜏
 There is only one forcing function or input variable, 𝐶𝐴0 .
 CSTRs WITH VARIABLE HOLDUPS
 Ifthe previous example is modified slightly to permit the volumes in
each reactor to vary with time, both total and component continuity
equations are required for each reactor.
 To show the effects of higher-order kinetics, assume the reaction is
now nth-order in reactant
 Total mass balance (continuity) equation:
𝑀𝑎𝑠𝑠 𝑓𝑙𝑜𝑤 𝑀𝑎𝑠𝑠 𝑓𝑙𝑜𝑤 𝑇𝑖𝑚𝑒 𝑟𝑎𝑡𝑒 𝑜𝑓 𝑐ℎ𝑎𝑛𝑔𝑒
 − =
𝑖𝑛𝑡𝑜 𝑠𝑦𝑠𝑡𝑒𝑚 𝑜𝑢𝑡 𝑜𝑓 𝑡ℎ𝑒 𝑠𝑦𝑠𝑡𝑒𝑚 𝑜𝑓 𝑚𝑎𝑠𝑠 𝑖𝑛𝑠𝑖𝑑𝑒 𝑠𝑦𝑠𝑡𝑒𝑚
 Component Balance:
𝐹𝑙𝑜𝑤 𝑜𝑓 𝑚𝑜𝑙𝑒𝑠 𝑜𝑓 𝑗𝑡ℎ 𝐹𝑙𝑜𝑤 𝑜𝑓 𝑚𝑜𝑙𝑒𝑠 𝑜𝑓 𝑗𝑡ℎ
 − +
𝑐𝑜𝑚𝑝𝑜𝑛𝑒𝑛𝑡 𝑖𝑛𝑡𝑜 𝑠𝑦𝑠𝑡𝑒𝑚 𝑐𝑜𝑚𝑝𝑜𝑛𝑒𝑛𝑡 𝑜𝑢𝑡 𝑜𝑓 𝑡ℎ𝑒 𝑠𝑦𝑠𝑡𝑒𝑚
𝑟𝑎𝑡𝑒 𝑜𝑓 𝑓𝑜𝑟𝑚𝑎𝑡𝑖𝑜𝑛 𝑜𝑓 𝑚𝑜𝑙𝑒𝑠 𝑜𝑓 𝑗𝑡ℎ 𝑇𝑖𝑚𝑒 𝑟𝑎𝑡𝑒 𝑜𝑓 𝑐ℎ𝑎𝑛𝑔𝑒 𝑜𝑓 𝑚𝑜𝑙𝑒𝑠 𝑜𝑓 𝑗𝑡ℎ
=
𝑐𝑜𝑚𝑝𝑜𝑛𝑒𝑛𝑡 𝑓𝑟𝑜𝑚 𝑐ℎ𝑒𝑚𝑖𝑐𝑎𝑙 𝑟𝑒𝑎𝑐𝑡𝑖𝑜𝑛𝑠 𝑐𝑜𝑚𝑝𝑜𝑛𝑒𝑛𝑡 𝑖𝑛𝑠𝑖𝑑𝑒 𝑠𝑦𝑠𝑡𝑒𝑚
 Reactor 1:
𝑑 𝜌𝑉1 𝑑𝑉1 𝐹0
= 𝜌𝐹0 − 𝜌𝐹1 ՜ = 𝐹0 − 𝐹1
𝑑𝑡 𝑑𝑡 𝐶𝐴0
𝑑 𝑛
𝑉1 𝐶𝐴1 = 𝐹0 𝐶𝐴0 − 𝐹1 𝐶𝐴1 − 𝑉1 𝑘1 𝐶𝐴1
𝑑𝑡 𝑉1 𝐹1 𝐹2 𝑉3 𝐹3
𝑉2
 Reactor 2: 𝑘1 𝐶𝐴1 𝑘2 𝐶𝐴2 𝑘3 𝐶𝐴3

𝑑 𝜌𝑉2 𝑑𝑉2
= 𝜌𝐹1 − 𝜌𝐹2 = 𝐹1 − 𝐹2
𝑑𝑡 𝑑𝑡
𝑑 𝑛
𝑉2 𝐶𝐴2 = 𝐹1 𝐶𝐴1 − 𝐹2 𝐶𝐴2 − 𝑉2 𝑘2 𝐶𝐴2
𝑑𝑡
 Reactor 3:
𝑑 𝜌𝑉3 𝑑𝑉3
= 𝜌𝐹2 − 𝜌𝐹3 ՜ = 𝐹2 − 𝐹3
𝑑𝑡 𝑑𝑡
𝑑 𝑛
𝑉3 𝐶𝐴3 = 𝐹2 𝐶𝐴2 − 𝐹3 𝐶𝐴3 − 𝑉3 𝑘3 𝐶𝐴3
𝑑𝑡
 Our mathematical model now contains six first-order nonlinear ordinary
differential equations.
 Parameters that must be known are 𝑘1 , 𝑘2 , 𝑘3 , and n.
 Initial conditions for all the dependent variables that are to be integrated
must be given: 𝐶𝐴1 , 𝐶𝐴2 , 𝐶𝐴3 , 𝑉1 , 𝑉2 , 𝑎𝑛𝑑 𝑉3 .
 The forcing functions CA0(t) and Fo(t) must also be given.
 Let us now check the degrees of freedom of this system.
 There are six equations. But there are nine unknowns:
𝐶𝐴1 , 𝐶𝐴2 , 𝐶𝐴3 , 𝑉1 , 𝑉2 , 𝑉3 , 𝐹1 , 𝐹2 , 𝑎𝑛𝑑 𝐹3 .
 Clearly this system is not sufficiently specified and a solution could not be
obtained.
 What have we missed in our modeling?
 A good plant operator could take one look at the system and see what
the problem is.
 We have not specified how the flows out of the tanks are to be set.
 Physically there would probably be control valves in the outlet lines to
regulate the flows.
 How are these control valves to be set?
 A common configuration is to have the level in the tank controlled by the
outflow, i.e., a level controller opens the control valve on the exit line to
increase the outflow if the level in the tank increases.
 Thus there must be a relationship between tank holdup and flow.
𝐹1 = 𝑓 𝑉1 𝐹2 = 𝑓 𝑉2 𝐹3 = 𝑓(𝑉3 )
 The f functions will describe the level controller and the control valve.
 These three equations reduce the degrees of freedom to zero.
 It might be worth noting that we could have considered the flow from
the third tank 𝐹3 as the forcing function.
 Then the level in tank 3 would probably be maintained by the flow into
the tank, 𝐹2 .
 The level in tank 2 would be controlled by 𝐹1 , and tank 1 level by 𝐹0 . We
would still have three equations.
 The reactors shown in Fig. would operate at atmospheric pressure if they
were open to the atmosphere as sketched.
 If the reactors are not vented and if no inert blanketing is assumed, they
would run at the bubble point pressure for the specified temperature and
varying composition.
 Therefore the pressures could be different in each reactor, and they would
vary with time, even though temperatures are assumed constant, as the
𝐶𝐴 ‘𝑠 change. 𝐹0
𝐶𝐴0

𝑉1 𝐹1 𝐹2 𝑉3 𝐹3
𝑉2
𝑘1 𝐶𝐴1 𝑘2 𝐶𝐴2 𝑘3 𝐶𝐴3
TWO HEATED TANKS
As our next fairly simple system let us consider a process in which two energy
balances are needed to model the system. The flow rate F of oil passing
through two perfectly mixed tanks in series is constant at 90 ft3/min. The
𝑙𝑏𝑚
density 𝜌 of the oil is constant at 40 ൗ𝑓𝑡 3, and its heat capacity 𝐶𝑃 is 0.6
Btu/lbm°F. The volume of the first tank 𝑉1 is constant at 450 ft3, and the
volume of the second tank 𝑉2 is constant at 90 ft3.
The temperature of the oil entering the first 𝐹0
tank is 𝑇0 and is 150°F at the initial steady 𝑇0
state. The temperatures in the two tanks
are 𝑇1 and 𝑇2 . They are both equal to 𝑉1 𝐹1
𝑉2
𝐹2

250°F at the initial steady state. A heating 𝑄1 𝑇1 𝑇1 𝑇2 𝑇2

coil in the first tank uses steam to heat the


oil. Let Q1 be the heat addition rate in the
first tank.
Solution: Energy Balance for single reactor: 𝐹0
𝜌0
𝐹𝑙𝑜𝑤 𝑜𝑓 𝑖𝑛𝑡𝑒𝑟𝑛𝑎𝑙, 𝑘𝑖𝑛𝑒𝑡𝑖𝑐, 𝑎𝑛𝑑 𝐹𝑙𝑜𝑤 𝑜𝑓 𝑖𝑛𝑡𝑒𝑟𝑛𝑎𝑙, 𝑘𝑖𝑛𝑒𝑡𝑖𝑐, 𝑎𝑛𝑑 𝐶𝐴0
𝑝𝑜𝑡𝑒𝑛𝑡𝑖𝑎𝑙 𝑒𝑛𝑒𝑟𝑔𝑦 𝑖𝑛𝑡𝑜 𝑠𝑦𝑠𝑡𝑒𝑚 − 𝑝𝑜𝑡𝑒𝑛𝑡𝑖𝑎𝑙 𝑒𝑛𝑒𝑟𝑔𝑦 𝑜𝑢𝑡 𝑜𝑓 𝑠𝑦𝑠𝑡𝑒𝑚 𝑇0
𝑏𝑦 𝑐𝑜𝑛𝑣𝑒𝑐𝑡𝑖𝑜𝑛 𝑜𝑟 𝑑𝑖𝑓𝑓𝑢𝑠𝑖𝑜𝑛 𝑏𝑦 𝑐𝑜𝑛𝑣𝑒𝑐𝑡𝑖𝑜𝑛 𝑜𝑟 𝑑𝑖𝑓𝑓𝑢𝑠𝑖𝑜𝑛
ℎ𝑒𝑎𝑡 𝑎𝑑𝑑𝑒𝑑 𝑡𝑜 𝑠𝑦𝑠𝑡𝑒𝑚 𝑏𝑦 𝑤𝑜𝑟𝑘 𝑑𝑜𝑛𝑒 𝑏𝑦 𝑠𝑦𝑠𝑡𝑒𝑚 𝑜𝑛 𝑠𝑢𝑟𝑟𝑜𝑢𝑛𝑑𝑖𝑛𝑔𝑠
+ − 𝐹
𝑐𝑜𝑛𝑑𝑢𝑐𝑡𝑖𝑜𝑛, 𝑟𝑎𝑑𝑖𝑎𝑡𝑖𝑜𝑛 𝑎𝑛𝑑 𝑟𝑒𝑎𝑐𝑡𝑖𝑜𝑛 𝑠ℎ𝑎𝑓𝑡 𝑤𝑜𝑟𝑘 𝑎𝑛𝑑 𝑃𝑉 𝑤𝑜𝑟𝑘
𝑇𝑖𝑚𝑒 𝑟𝑎𝑡𝑒 𝑜𝑓 𝑐ℎ𝑎𝑛𝑔𝑒 𝑜𝑓 𝑖𝑛𝑡𝑒𝑟𝑛𝑎𝑙, 𝜌
= 𝑉 𝐶𝐴
𝑘𝑖𝑛𝑒𝑡𝑖𝑐, 𝑝𝑜𝑡𝑒𝑛𝑡𝑖𝑎𝑙 𝑒𝑛𝑒𝑟𝑔𝑦 𝑖𝑛𝑠𝑖𝑑𝑒 𝑠𝑦𝑠𝑡𝑒𝑚
𝜌 𝑇
-Q
𝐹0 𝜌0 𝑈0 + 𝐾0 + 𝜙0 − 𝐹𝜌 𝑈 + 𝐾 + 𝜙 + 𝑄𝐺 + 𝑄 𝐶𝐴
𝑇
− 𝑊 + 𝐹𝑃 − 𝐹0 𝑃0
𝑑 𝑈 + 𝐾 + 𝜙 𝑉𝜌
=
𝑑𝑡 Note that all the terms in above Eq.
where U = internal energy (energy per unit mass)
K = kinetic energy (energy per unit mass)
must have the same units (energy
𝜙 = potential energy (energy per unit mass) per time) so the FP terms must use
W = shaft work done by system (energy per time) the appropriate conversion factor
P = pressure of system (778 ft. lbf/Btu in English
P0 = pressure of feed stream
engineering units)
W=0, K, 𝜙 are negligible
𝑑 𝑉𝜌𝑈 𝑃 𝑃0
= 𝐹0 𝜌0 𝑈0 − 𝐹𝜌𝑈 + 𝑄𝐺 + 𝑄 − 𝐹𝜌 + 𝐹0 𝜌0
𝑑𝑡 𝜌 𝜌0
= 𝐹0 𝜌0 (𝑈0 + 𝑃0 𝑉0 ) − 𝐹𝜌 𝑈 + 𝑃𝑉ത + 𝑄𝐺 + 𝑄
1
where ഥ𝑉 = is the specific volume (ft3/lbm or m3/kg), the reciprocal of the
𝜌0
𝐹0
density. 𝜌0
𝐶𝐴0
Enthalpy, H or h, is defined: 𝑇0
𝐹
𝐻 𝑜𝑟 ℎ = 𝑈 + 𝑃𝑉ത 𝑉
𝜌
𝜌
-Q 𝐶𝐴
𝐶𝐴
𝑑 𝜌𝑉𝑈 𝑇 𝑇
= 𝐹0 𝜌0 ℎ0 − 𝐹𝜌ℎ + 𝑄 − 𝜆𝑉𝑘𝐶𝐴
𝑑𝑡
For liquids the 𝑃𝑉ത term is negligible compared to the U term, and we use the
time rate of change of the enthalpy of the system instead of the internal
energy of the system.
𝑑 𝜌𝑉ℎ
= 𝐹0 𝜌0 ℎ0 − 𝐹𝜌ℎ + 𝑄 − 𝜆𝑉𝑘𝐶𝐴
𝑑𝑡
The enthalpies are functions of composition, temperature, and pressure, but
primarily temperature. From thermodynamics, the heat capacities at constant
pressure, Cp and at constant volume, CV are
𝜕𝐻 𝜕𝑈
𝐶𝑃 = 𝐶𝑉 =
𝜕𝑇 𝑃 𝜕𝑇 𝑉
To illustrate that the energy is primarily influenced by temperature, let us
simplify the problem by assuming that the liquid enthalpy can be expressed
as a product of absolute temperature and an average heat capacity CP
(Btu/lbm°R or Cal/g K) that is constant. ℎ = 𝐶𝑃 𝑇
We will also assume that the densities of all the liquid streams are constant.
With these simplifications energy balance Equation becomes
𝑑 𝑉𝑇
𝜌𝑐𝑃 = 𝜌𝑐𝑃 𝐹0 𝑇0 − 𝐹𝑇 + 𝑄 − 𝜆𝑉𝑘𝐶𝐴
𝑑𝑡
There is one energy balance for each tank, except there is no reaction
involved in this process.
𝐹𝑙𝑜𝑤 𝑜𝑓 𝑖𝑛𝑡𝑒𝑟𝑛𝑎𝑙, 𝑘𝑖𝑛𝑒𝑡𝑖𝑐, 𝑎𝑛𝑑 𝐹𝑙𝑜𝑤 𝑜𝑓 𝑖𝑛𝑡𝑒𝑟𝑛𝑎𝑙, 𝑘𝑖𝑛𝑒𝑡𝑖𝑐, 𝑎𝑛𝑑
𝑝𝑜𝑡𝑒𝑛𝑡𝑖𝑎𝑙 𝑒𝑛𝑒𝑟𝑔𝑦 𝑖𝑛𝑡𝑜 𝑠𝑦𝑠𝑡𝑒𝑚 − 𝑝𝑜𝑡𝑒𝑛𝑡𝑖𝑎𝑙 𝑒𝑛𝑒𝑟𝑔𝑦 𝑜𝑢𝑡 𝑜𝑓 𝑠𝑦𝑠𝑡𝑒𝑚
𝑏𝑦 𝑐𝑜𝑛𝑣𝑒𝑐𝑡𝑖𝑜𝑛 𝑜𝑟 𝑑𝑖𝑓𝑓𝑢𝑠𝑖𝑜𝑛 𝑏𝑦 𝑐𝑜𝑛𝑣𝑒𝑐𝑡𝑖𝑜𝑛 𝑜𝑟 𝑑𝑖𝑓𝑓𝑢𝑠𝑖𝑜𝑛
ℎ𝑒𝑎𝑡 𝑎𝑑𝑑𝑒𝑑 𝑡𝑜 𝑠𝑦𝑠𝑡𝑒𝑚 𝑏𝑦 𝑤𝑜𝑟𝑘 𝑑𝑜𝑛𝑒 𝑏𝑦 𝑠𝑦𝑠𝑡𝑒𝑚 𝑜𝑛 𝑠𝑢𝑟𝑟𝑜𝑢𝑛𝑑𝑖𝑛𝑔𝑠
+ −
𝑐𝑜𝑛𝑑𝑢𝑐𝑡𝑖𝑜𝑛, 𝑟𝑎𝑑𝑖𝑎𝑡𝑖𝑜𝑛 𝑎𝑛𝑑 𝑟𝑒𝑎𝑐𝑡𝑖𝑜𝑛 𝑠ℎ𝑎𝑓𝑡 𝑤𝑜𝑟𝑘 𝑎𝑛𝑑 𝑃𝑉 𝑤𝑜𝑟𝑘
𝑇𝑖𝑚𝑒 𝑟𝑎𝑡𝑒 𝑜𝑓 𝑐ℎ𝑎𝑛𝑔𝑒 𝑜𝑓 𝑖𝑛𝑡𝑒𝑟𝑛𝑎𝑙,
=
𝑘𝑖𝑛𝑒𝑡𝑖𝑐, 𝑝𝑜𝑡𝑒𝑛𝑡𝑖𝑎𝑙 𝑒𝑛𝑒𝑟𝑔𝑦 𝑖𝑛𝑠𝑖𝑑𝑒 𝑠𝑦𝑠𝑡𝑒𝑚 𝐹0
𝑇0
Energy balance for tank 1:
𝑑 𝜌𝐶𝑃 𝑉1 𝑇1 𝑉1 𝐹1 𝐹2
= 𝜌𝐶𝑃 𝐹0 𝑇0 − 𝐹1 𝑇1 + 𝑄1 𝑉2
𝑑𝑡 𝑄1 𝑇1 𝑇1 𝑇2 𝑇2
Energy balance for tank 2:
𝑑 𝜌𝐶𝑃 𝑉2 𝑇2
= 𝜌𝐶𝑃 𝐹1 𝑇1 − 𝐹2 𝑇2
𝑑𝑡
Since the throughput is constant 𝐹0 = 𝐹1 = 𝐹2 = 𝐹. Since volumes,
densities, and heat capacities are all constant, above Eqs. can be
simplified.
𝑑𝑇1
𝜌𝐶𝑃 𝑉1 = 𝜌𝐶𝑃 𝐹 𝑇0 − 𝑇1 + 𝑄1
𝑑𝑡
𝑑𝑇2
𝜌𝐶𝑃 𝑉2 = 𝜌𝐶𝑃 𝐹 𝑇1 − 𝑇2
𝑑𝑡
Let’s check the degrees of freedom of this system. The parameter
values that are known are 𝝆, 𝑪𝑷 , 𝑽𝟏 , 𝑽𝟐 , 𝒂𝒏𝒅 𝑭. The heat input to the
first tank Q1 would be set by the position of the control valve in the
steam line.
Thus we are left with two dependent variables, T1 and T2, and we
have two equations. So the system is correctly specified.
GAS-PHASE PRESSURIZED CSTR
Suppose a mixture of gases is fed into the reactor sketched in Fig. The
reactor is filled with reacting gases which are perfectly mixed. A reversible
reaction occurs: 𝑘1
2𝐴֞ 𝐵
𝑘2
The forward reaction is 1.5th-order in A, the reverse reaction is first-order
in B. Note that the stoichiometric coefficient for A and the order of the
reaction are not the same. The mole fraction of reactant A in the reactor is
y. The pressure inside the vessel is P (absolute). Both P and y can vary with
time. The volume of the reactor V is constant.
We will assume an isothermal system so
the temperature T is constant. Perfect 𝐹0
𝑉 𝑃 𝐹
𝜌0
gases are also assumed. The feed stream 𝑦0 𝑇 𝑦 𝜌 𝑃𝐷
𝑦
has a density 𝜌0 and a mole fraction y0 of
reactant A. its volumetric flow rate is F0.
The flow out of the reactor passes through a restriction (control valve) into
another vessel which is held at a constant pressure PD (absolute). The
outflow will vary with the pressure and the composition of the reactor.
Flows through control valves are
𝑃−𝑃𝐷
𝐹 = 𝐶𝑣 (1)
𝜌
Cv is the valve-sizing coefficient.
Density varies with pressure and composition.
𝑃𝑀 𝑃
𝜌= = 𝑦𝑀𝐴 + 1 − 𝑦 𝑀𝐵 (2)
𝑅𝑇 𝑅𝑇
where M = average molecular weight 𝐹
0
MA = molecular weight of reactant A 𝜌0 𝑉
𝑇
𝑃
𝑦
𝐹
𝜌 𝑃𝐷
𝑦0
MB = molecular weight of product B 𝑦

The concentration of reactant in the reactor is


𝑃𝑦
𝐶𝐴 = (3) with units of moles of A per unit volume.
𝑅𝑇
The overall reaction rate for the forward reaction is
1.5 1 𝑑𝑛𝐴 1 𝑑𝑛𝐵
ℛ𝐹 = 𝑘1 𝐶𝐴 = − =
2𝑉 𝑑𝑡 𝑅 𝑉 𝑑𝑡 𝑅
The overall reaction rate for the reverse reaction is
1 𝑑𝑛𝐴 1 𝑑𝑛𝐵
ℛ𝑅 = 𝑘2 𝐶𝐵 = =−
2𝑉 𝑑𝑡 𝑅 𝑉 𝑑𝑡 𝑅
With these fundamental relationships pinned down, we are ready to write
the total and component continuity equations.
Total continuity :
𝑑𝜌 𝐹0
𝐹
𝑉 = 𝜌0 𝐹0 − 𝜌𝐹 (4) 𝜌0 𝑉 𝑃
𝑃𝐷
𝑑𝑡 𝑦0 𝑇 𝑦 𝜌
Component A continuity: 𝑦

𝑑𝐶𝐴
𝑉 = 𝐹0 𝐶𝐴0 − 𝐹𝐶𝐴 − 2𝑉𝑘1 𝐶𝐴1.5 + 2𝑉𝑘2 𝐶𝐵 (5)
𝑑𝑡
The 2 in the reaction terms comes from the stoichiometric coefficient of A.
There are five equations that make up the mathematical model of this
system.
The parameters that must be known are 𝑽, 𝑪𝒗 , 𝒌𝟏 , 𝒌𝟐 , 𝑹, 𝑴𝑨 , 𝒂𝒏𝒅 𝑴𝑩 .
The forcing functions (or inputs) could be 𝑷𝑫 , 𝝆𝟎 , 𝑭𝟎 , 𝒂𝒏𝒅 𝑪𝑨𝟎 .
This leaves five unknowns (dependent variables): 𝑪𝑨 , 𝝆, 𝑷, 𝑭, 𝒂𝒏𝒅 𝒚.

𝐹0
𝜌0 𝑉 𝑃 𝐹
𝑦0 𝑇 𝑦 𝜌 𝑃𝐷
𝑦
NON ISOTHERMAL REACTOR
NONISOTHERMAL CSTR
In the reactors studied so far, we have shown the effects of variable
holdups, variable densities, and higher-order kinetics on the total and
component continuity equations. Energy equations were not needed
because we assumed isothermal operations.
Let us now consider a system in which temperature can change with time.
𝑘
An irreversible, exothermic reaction 𝐴 ՜ 𝐵 is carried out in a single
perfectly mixed CSTR as shown in Fig.
The reaction is nth-order in reactant A and
𝐹0
has a heat of reaction 𝜆 (Btu/lb. mol of A 𝐶𝐴0
reacted). Negligible heat losses and constant 𝑇0 𝑉
𝐹𝑗
𝑇𝑗
densities are assumed. To remove the heat of 𝑇𝑗
𝑇
reaction, a cooling jacket surrounds the 𝐹𝑗 𝑉𝑗 𝐶𝐴
𝐹
reactor. Cooling water is added to the jacket 𝑇 𝑗0
𝐶𝐴0
at a volumetric flow rate Fj and with an inlet 𝑇

temperature of 𝑇𝑗0 .
The volume of water in the jacket Vj is constant. The mass of the metal
walls is assumed negligible so the “thermal inertia” of the metal need not
be considered. This is often a fairly good assumption because the heat
capacity of steel is only about 0.1 Btu/lb,°F, which is an order of
magnitude less than that of water.
A. PERFECTLY MIXED COOLING ACKET. We assume that the temperature
everywhere in the jacket is Tj. The heat transfer between the process at
temperature T and the cooling water at temperature Tj is described by an
overall heat transfer coefficient. 𝐹0
𝐶𝐴0
𝐹𝑗
𝑇0
𝑄 = 𝑈𝐴𝐻 𝑇 − 𝑇𝑗 𝑇𝑗
𝑉 𝑇𝑗
𝑇
where Q = heat transfer rate 𝐹𝑗 𝑉𝑗 𝐶𝐴
𝐹
𝑇 𝑗0
𝐶𝐴
U = overall heat transfer coefficient 𝑇

AH = heat transfer area


In general the heat transfer area could vary with the holdup in the reactor
if some area was not completely covered with reaction mass liquid at all
times. The equations describing the system are:
Reactor total continuity: I/P –O/P =ACC
𝐹0
𝑑(𝜌𝑉) 𝑑𝑉 𝐶𝐴0
𝐹𝑗
= 𝜌𝐹0 − 𝜌𝐹 ՜ = 𝐹0 − 𝐹 𝑇0 𝑉
𝑑𝑡 𝑑𝑡 𝑇𝑗 𝑇𝑗
𝑇
Reactor component A continuity: 𝐹𝑗 𝑉𝑗 𝐶𝐴
𝑇 𝑗0 𝐹
𝑑 𝑉𝐶𝐴 𝐶𝐴
= 𝐹0 𝐶𝐴0 − 𝐹𝐶𝐴 − 𝑉𝑘𝐶𝐴𝑛 𝑇
𝑑𝑡
Reactor energy equation :
𝑑 𝑉ℎ
𝜌 = 𝜌(𝐹0 ℎ0 − 𝐹ℎ) − 𝜆𝑉𝑘𝐶𝐴𝑛 − 𝑈𝐴𝐻 (𝑇 − 𝑇𝑗 )
𝑑𝑡
Jacket energy equation : here 𝜌𝑗 = density of cooling water
𝑑ℎ𝑗
𝜌𝑗 𝑉𝑗 = 𝐹𝑗 𝜌𝑗 ℎ𝑗0 − ℎ𝑗 + 𝑈𝐴𝐻 𝑇 − 𝑇𝑗 h = enthalpy of process liquid
𝑑𝑡
ℎ𝑗 = enthalpy of cooling water
The assumption of constant densities makes 𝐶𝑃 = 𝐶𝑉 and permits us to
use enthalpies in the time derivatives to replace internal energies.
A hydraulic relationship between reactor holdup and the flow out of the
reactor is also needed.
A level controller is assumed to change the outflow as the volume in the
tank rises or falls: the higher the volume, the larger the outflow. The
outflow is shut off completely when the volume drops to a minimum
value Vmin . 𝐹 = 𝐾𝑉 𝑉 − 𝑉𝑚𝑖𝑛
The level controller is a proportional-only feedback controller.
Finally, we need enthalpy data to relate the h’s to compositions and
𝐹0
temperatures. Let us assume the simple forms 𝐶𝐴0
𝐹𝑗
𝑇0
ℎ = 𝐶𝑃 𝑇 𝑎𝑛𝑑 ℎ𝑗 = 𝐶𝑗 𝑇𝑗 𝑇𝑗
𝑉 𝑇𝑗
𝑇
where 𝐶𝑃 = heat capacity of the process liquid 𝐹𝑗 𝑉𝑗 𝐶𝐴
𝐹
𝑇 𝑗0
𝐶𝑗 = heat capacity of the cooling water 𝐶𝐴
𝑇
Using enthalpy Eqs. and the Arrhenius relationship for k, the five
𝐹0
equations that describe the process are 𝐶𝐴0
𝐹𝑗
𝑑𝑉 𝑇0 𝑉
= 𝐹0 − 𝐹 𝑇𝑗 𝑇𝑗
𝑑𝑡 𝑇
𝐸 𝐹𝑗 𝑉𝑗 𝐶𝐴
𝑑 𝑉𝐶𝐴 −𝑅𝑇 𝐹
= 𝐹0 𝐶𝐴0 − 𝐹𝐶𝐴 − 𝑉𝑘0 𝐶𝐴𝑛 𝑒 𝑇 𝑗0
𝐶𝐴
𝑑𝑡
𝐸 𝑇
𝑑 𝑉𝑇
𝜌𝐶𝑃 = 𝜌𝐶𝑃 (𝐹0 𝑇0 − 𝐹𝑇) − 𝜆𝑉𝑘0 𝐶𝐴𝑛 𝑒 −𝑅𝑇 − 𝑈𝐴𝐻 (𝑇 − 𝑇𝑗 )
𝑑𝑡
𝑑𝑇𝑗
𝜌𝑗 𝑉𝑗 𝐶𝑗 = 𝐹𝑗 𝜌𝑗 𝐶𝑗 𝑇𝑗0 − 𝑇𝑗 + 𝑈𝐴𝐻 𝑇 − 𝑇𝑗
𝑑𝑡
𝐹 = 𝐾𝑉 𝑉 − 𝑉𝑚𝑖𝑛
Checking the degrees of freedom, we see that there are five equations and
five unknowns: V, F, CA, T, and TJ. We must have initial conditions for
these five dependent variables. The forcing functions are T0, F0 , CA0, and
Fj.
The parameters that must be known are 𝑛, 𝑘0 , 𝐸, 𝑅, 𝜌, 𝐶𝑃 , 𝑈, 𝐴𝐻 , 𝜌𝑗 , 𝑉𝑗 , 𝐶𝑗 , 𝑇𝑗0 ,
𝐾𝑉 , 𝑎𝑛𝑑 𝑉𝑚𝑖𝑛 .
If the heat transfer area varies with the reactor holdup it would be included
as another variable, but we would also have another equation; the
relationship between area and holdup.
If the reactor is a flat-bottomed vertical cylinder with diameter D and if the
jacket is only around the outside, not around the bottom
𝜋 2 𝑉
𝐴𝐻 = 𝜋𝐷𝐻 ∵𝑉= 𝐷 𝐻՜𝐻=𝜋 2
4 𝐷
4
𝜋𝐷𝑉 4
𝐴𝐻 = 𝜋 = 𝑉
𝐷2 𝐷
4
We have assumed the overall heat 𝐹0
transfer coefficient U is constant. It may 𝐶𝐴0
𝑇0
𝐹𝑗
be a function of the coolant flow rate F, 𝑇𝑗
𝑉
𝑇
𝑇𝑗

or the composition of the reaction mass, 𝐹𝑗 𝑉𝑗 𝐶𝐴


𝐹
giving one more variable but also one 𝑇 𝑗0
𝐶𝐴
more equation. 𝑇
B. PLUG FLOW COOLING JACKET.
In the model derived above, the cooling water inside the jacket was
assumed to be perfectly mixed. In many jacketed vessels this is not a
particularly good assumption.
If the water flow rate is high enough so that the water temperature
does not change much as it goes through the jacket, the mixing pattern
makes little difference.
However, if the water temperature rise is significant and if the flow is
more like plug flow than a perfect mix (this would certainly be the
case if a cooling coil is used inside the reactor instead of a jacket), then
an average jacket temperature TjA may be used.
𝑇𝑗0 +𝑇𝑗𝑒𝑥𝑖𝑡 𝐹0 𝐹𝑗
𝑇𝑗𝐴 = 𝐶𝐴0 𝑇 𝑗𝑒𝑥𝑖𝑡
2 𝑇0 𝑇𝑗
𝑉 𝑇 𝐶𝐴
where 𝑇𝑗𝑒𝑥𝑖𝑡 is the outlet cooling-water temp. 𝐹𝑗 𝑉𝑗 𝐹
𝑇 𝑗0 𝐶𝐴0
𝑇
The average temperature is used in the heat transfer equation and to
represent the enthalpy of jacket material.
Reactor Energy equation:
𝐸
𝑑 𝑉𝑇 𝑛 −𝑅𝑇
𝜌𝐶𝑃 = 𝜌𝐶𝑃 (𝐹0 𝑇0 − 𝐹𝑇) − 𝜆𝑉𝑘0 𝐶𝐴 𝑒 − 𝑈𝐴𝐻 (𝑇 − 𝑇𝑗𝐴 )
𝑑𝑡
Jacket Energy Equation
𝑑𝑇𝑗𝐴
𝜌𝑗 𝑉𝑗 𝐶𝑗 = 𝐹𝑗 𝜌𝑗 𝐶𝑗 𝑇𝑗0 − 𝑇𝑗𝑒𝑥𝑖𝑡 + 𝑈𝐴𝐻 𝑇 − 𝑇𝑗𝐴
𝑑𝑡
Equation is integrated to obtain TjA at each instant in time, and 𝑇𝑗𝐴 is used
to calculate Tjexit, also as a function of time. 𝐹0
𝐶𝐴0
𝐹𝑗
𝑇0 𝑉
𝑇𝑗 𝑇𝑗
𝑇
𝐹𝑗 𝑉𝑗 𝐶𝐴
𝑇 𝑗0 𝐹
𝐶𝐴0
𝑇
C. LUMPED JACKET MODEL. Another alternative is to break up the jacket
volume into a number of perfectly mixed “lumps” as shown in Fig..
An energy equation is needed for & lump. Assuming four lumps of equal
volume and heat transfer area, we get four energy equations for the
jacket:
1 𝑑𝑇𝑗1 1 𝑄 𝑇
𝜌𝑗 𝑉𝑗 𝐶𝑗 = 𝐹𝑗 𝜌𝑗 𝐶𝑗 𝑇𝑗0 − 𝑇𝑗1 + 𝑈𝐴𝐻 𝑇 − 𝑇𝑗1 4 𝑗4

4 𝑑𝑡 4
1 𝑑𝑇𝑗2 1 𝑄 3 𝑇
𝑗3

𝜌𝑗 𝑉𝑗 𝐶𝑗 = 𝐹𝑗 𝜌𝑗 𝐶𝑗 𝑇𝑗1 − 𝑇𝑗2 + 𝑈𝐴𝐻 𝑇 − 𝑇𝑗2


4 𝑑𝑡 4 𝑄 2 𝑇
𝑗2
1 𝑑𝑇𝑗3 1 𝐹
𝜌𝑗 𝑉𝑗 𝐶𝑗 = 𝐹𝑗 𝜌𝑗 𝐶𝑗 𝑇𝑗2 − 𝑇𝑗3 + 𝑈𝐴𝐻 𝑇 − 𝑇𝑗3 𝑗1

4 𝑑𝑡 4 𝑄 1 𝑇
𝑗1
𝑇
𝑗0

1 𝑑𝑇𝑗4 1
𝜌 𝑉𝐶 = 𝐹𝑗 𝜌𝑗 𝐶𝑗 𝑇𝑗3 − 𝑇𝑗4 + 𝑈𝐴𝐻 𝑇 − 𝑇𝑗4
4 𝑗 𝑗 𝑗 𝑑𝑡 4
D. SIGNIFICANT METAL WALL CAPCITANCE: In some reactors,
particularly high pressure vessels or smaller-scale equipment, the mass
of the metal walls and its effects on the thermal dynamics must be
considered.
To be rigorous, the energy equation for the wall should be a partial
differential equation in time and radial position.
A less rigorous but frequently used approximation is to “lump” the
mass of the metal and assume the metal is all at one temperature TM.
This assumption is a fairly good one when the wall is not too thick and
Metal
the thermal conductivity of the metal is large. Process
Jacket

Then effective inside and outside film coefficients


hi and ho, are used as shown in 𝑇
𝑇𝑚 𝑇𝑗

The three energy equations for the process are: 𝑄𝑖 𝑄𝑜

𝐹𝑗

𝑇𝑗0
𝐹𝑙𝑜𝑤 𝑜𝑓 𝑖𝑛𝑡𝑒𝑟𝑛𝑎𝑙, 𝑘𝑖𝑛𝑒𝑡𝑖𝑐, 𝑎𝑛𝑑 𝐹𝑙𝑜𝑤 𝑜𝑓 𝑖𝑛𝑡𝑒𝑟𝑛𝑎𝑙, 𝑘𝑖𝑛𝑒𝑡𝑖𝑐, 𝑎𝑛𝑑
𝑝𝑜𝑡𝑒𝑛𝑡𝑖𝑎𝑙 𝑒𝑛𝑒𝑟𝑔𝑦 𝑖𝑛𝑡𝑜 𝑠𝑦𝑠𝑡𝑒𝑚 − 𝑝𝑜𝑡𝑒𝑛𝑡𝑖𝑎𝑙 𝑒𝑛𝑒𝑟𝑔𝑦 𝑜𝑢𝑡 𝑜𝑓 𝑠𝑦𝑠𝑡𝑒𝑚
𝑏𝑦 𝑐𝑜𝑛𝑣𝑒𝑐𝑡𝑖𝑜𝑛 𝑜𝑟 𝑑𝑖𝑓𝑓𝑢𝑠𝑖𝑜𝑛 𝑏𝑦 𝑐𝑜𝑛𝑣𝑒𝑐𝑡𝑖𝑜𝑛 𝑜𝑟 𝑑𝑖𝑓𝑓𝑢𝑠𝑖𝑜𝑛
ℎ𝑒𝑎𝑡 𝑎𝑑𝑑𝑒𝑑 𝑡𝑜 𝑠𝑦𝑠𝑡𝑒𝑚 𝑏𝑦 𝑤𝑜𝑟𝑘 𝑑𝑜𝑛𝑒 𝑏𝑦 𝑠𝑦𝑠𝑡𝑒𝑚 𝑜𝑛 𝑠𝑢𝑟𝑟𝑜𝑢𝑛𝑑𝑖𝑛𝑔𝑠
+ −
𝑐𝑜𝑛𝑑𝑢𝑐𝑡𝑖𝑜𝑛, 𝑟𝑎𝑑𝑖𝑎𝑡𝑖𝑜𝑛 𝑎𝑛𝑑 𝑟𝑒𝑎𝑐𝑡𝑖𝑜𝑛 𝑠ℎ𝑎𝑓𝑡 𝑤𝑜𝑟𝑘 𝑎𝑛𝑑 𝑃𝑉 𝑤𝑜𝑟𝑘
𝑇𝑖𝑚𝑒 𝑟𝑎𝑡𝑒 𝑜𝑓 𝑐ℎ𝑎𝑛𝑔𝑒 𝑜𝑓 𝑖𝑛𝑡𝑒𝑟𝑛𝑎𝑙,
= Process
Metal
𝑘𝑖𝑛𝑒𝑡𝑖𝑐, 𝑝𝑜𝑡𝑒𝑛𝑡𝑖𝑎𝑙 𝑒𝑛𝑒𝑟𝑔𝑦 𝑖𝑛𝑠𝑖𝑑𝑒 𝑠𝑦𝑠𝑡𝑒𝑚 Jacket

𝑑 𝑉𝑇 −
𝐸
𝜌𝐶𝑃 = 𝜌𝐶𝑃 𝐹0 𝑇0 − 𝐹𝑇 − 𝜆𝑉𝑘0 𝐶𝐴𝑛 𝑒 𝑅𝑇 − ℎ𝑖 𝐴𝑖 𝑇 − 𝑇𝑀
𝑑𝑡 𝑇
𝑑𝑇𝑀 𝑇
𝑚 𝑗 𝑇

𝜌𝑀 𝑉𝑀 𝐶𝑀 = ℎ𝑖 𝐴𝑖 𝑇 − 𝑇𝑀 − ℎ𝑜 𝐴𝑜 𝑇𝑀 − 𝑇𝑗
𝑑𝑡 𝑖 𝑄 𝑄
𝑜
𝑑𝑇𝑗
𝜌𝑗 𝑉𝑗 𝐶𝑗 = 𝐹𝑗 𝜌𝑗 𝐶𝑗 𝑇𝑗0 − 𝑇𝑗 + ℎ𝑜 𝐴𝑜 𝑇𝑀 − 𝑇𝑗 𝐹𝑗
𝑑𝑡
𝑇𝑗0
where hi = inside heat transfer film coefficient
ho = outside heat transfer film coefficient, 𝜌𝑀 = density of metal wall
CM = heat capacity of metal wall, VM = volume of metal wall
Ai = inside heat transfer area
Ao = outside heat transfer area

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