Alginates and Their Biomedical Applications 2018
Alginates and Their Biomedical Applications 2018
Engineering 11
Alginates and
Their Biomedical
Applications
Springer Series in Biomaterials Science
and Engineering
Volume 11
Series editor
Prof. Min Wang
Department of Mechanical Engineering
The University of Hong Kong
Pokfulam Road, Hong Kong
e-mail: memwang@hku.hk
Aims and scope
The Springer Series in Biomaterials Science and Engineering addresses the
manufacture, structure and properties, and applications of materials that are in
contact with biological systems, temporarily or permanently. It deals with many
aspects of modern biomaterials, from basic science to clinical applications, as well as
host responses. It covers the whole spectrum of biomaterials – polymers, metals,
glasses and ceramics, and composites/hybrids – and includes both biological
materials (collagen, polysaccharides, biological apatites, etc.) and synthetic materials.
The materials can be in different forms: single crystals, polycrystalline materials,
particles, fibers/wires, coatings, non-porous materials, porous scaffolds, etc. New
and developing areas of biomaterials, such as nano-biomaterials and diagnostic and
therapeutic nanodevices, are also focuses in this series. Advanced analytical
techniques that are applicable in R & D and theoretical methods and analyses for
biomaterials are also important topics. Frontiers in nanomedicine, regenerative
medicine and other rapidly advancing areas calling for great explorations are highly
relevant.
The Springer Series in Biomaterials Science and Engineering aims to provide
critical reviews of important subjects in the field, publish new discoveries and sig-
nificant progresses that have been made in both biomaterials development and the
advancement of principles, theories and designs, and report cutting-edge research
and relevant technologies. The individual volumes in the series are thematic. The
goal of each volume is to give readers a comprehensive overview of an area where
new knowledge has been gained and insights made. Significant topics in the area are
dealt with in good depth and future directions are predicted on the basis of current
developments. As a collection, the series provides authoritative works to a wide
audience in academia, the research community, and industry.
Alginates are polysaccharides, which are naturally produced by seaweeds and bac-
teria, and they have been known by humans for more than a century. Mannuronic
acid and guluronic acid are the basic constituents of these polymers. Alginates pos-
sess unique properties, which are harnessed for various applications such as in
foods, cosmetics, and fabric products as well as pharmaceutical/biomedical and
other industrial purposes. From being harvested as nutritional source from seaweeds
in the oceans to being applied in the biomedical field, alginates have been exten-
sively researched for almost a century. Owing to the revolution of scientific meth-
ods, technological advancements, and interdisciplinary approaches, the application
of alginates has been revolutionized in the recent two decades because their proper-
ties and modification capabilities have been better understood for adjusting them to
our needs. Nowadays, it is well established that alginates are suitable for pharma-
ceutical and biomedical engineering approaches. Historically, supplying alginates
via harvesting seaweeds imposed ecological concerns, while increasing oceanic
impurities and limited chemical modification methods could not extend alginate
applications beyond traditional usage. In the last decade, scientific efforts have
described in more detail the physicochemical properties and molecular interaction
of alginates with other polymeric and non-polymeric substances. Furthermore,
technological and engineering advancements in the modification and fabrication of
biopolymers have extended alginate applications into advanced biomaterial engi-
neering for the production of high-value pharmaceutical and biomedical products.
Current research outputs demonstrate that alginates and associated derivatives are
invaluable components in therapeutic developments such as tissue engineering, cell
therapy, cancer treatment, drug delivery, and treatment of cardiovascular diseases
and metabolic disorders. However, the current biomedical applicability of alginates
has been based on algal alginates, while bacterial alginates which display different
physicochemical properties have remained unharnessed. Indeed, alginate-producing
bacteria are cell factories with the potential of supplying such biopolymers for bio-
medical purposes. More importantly, in contrast to alginates originated from sea-
weeds, our understanding of alginate biosynthesis pathways has been based on
alginate-producing bacteria. Hence, extensive molecular studies on alginate
v
vi Preface
biosynthesis in bacteria have paved the path toward biotechnological alginate pro-
duction and showed that this is a promising path for tailoring alginates to exhibit
novel and reproducible compositions and properties for high-value purposes.
Therefore, we believe that the application of alginates can go further and beyond
current biomedical applications by harnessing various algal and bacterial alginates
in combination with advanced technological and biotechnological techniques.
While the number of scientific studies and published data on alginates and their
biomedical and pharmaceutical applications are enormous, they have not been
reviewed and covered in a single book before.
This book consists of 11 chapters and presents recent advances on the character-
ization and production of alginates from seaweeds and bacteria, as well as it outlines
applications of alginates for advanced biomedical and pharmaceutical purposes.
Chapter 1 highlights our understanding of alginate biosynthesis in bacteria and
algae toward biotechnological production. Chapter 2 focuses on algal alginates and
their farming which is accepted as an alternative to utilizing natural resources.
Chapter 3 distils recent advances by which alginates are considered as an important
biopolymer in cell microencapsulation technology and as a platform for controlled
drug and therapeutic factor delivery through cell encapsulation. This chapter pro-
vides the state-of-the-art technologies and current research strategies by which algi-
nates have been employed in formulations for treating prevalent human diseases and
disorders. Chapter 4 focuses on the processing techniques mainly used for manufac-
turing and processing products of alginates and their potential applications in bio-
medical science, tissue engineering, and drug delivery. This chapter also suggests
future perspectives for their novel applications in the biomedical field. Chapter 5
provides a comprehensive overview of the applications of alginate-based hydrogels
to design various forms of constructs and scaffolds for tissue engineering, for exam-
ple, via bioprinting, an area of high-tech research that has created a substantial
foundation for treating and curing many prevalent diseases. Chapter 6 reviews
application of alginates based on 3D in vitro models in the field of tumor and cancer
research. Chapter 7 describes the versatile biomedical applications of alginates in
creating solutions for treatment of heart and cardiovascular diseases. Chapter 8 dis-
cusses the potentials of alginates for designing dressings for the treatment and man-
agement of wounds. Chapter 9 highlights different strategies by which alginates
have been introduced in formulations for treating metabolic syndromes such as gas-
trointestinal tract disorders, obesity, type 2 diabetes, hypertension, nonalcoholic
fatty liver disease, and dyslipidemia. Chapter 10 outlines the research performed to
date for introducing alginate oligomers in designing effective formulations with
multiple therapeutic applications such as the management of chronic lung diseases,
biofilm infections, and antibiotic use. Chapter 11 highlights the application of man-
nuronic acid as one of the safest drugs with potent anti-inflammatory and immuno-
suppressive properties.
vii
viii Contents
ix
x Contributors
cystic fibrosis patient, they found that this bacteria form an unusually large mucoid
colonies on plates and constituting polysaccharides was reported resembling alginic
acid [17]. Later in 1966, they reported that alginate from this Pseudomonas isolate
was acetylated contrary to algal alginates while acetyl groups were lost during alka-
line extraction [18]. In the same year, Gorin and Spencer reported alginic acid pro-
duction by Azotobacter vinelandii [19]. In 1981, Govan and coworkers introduced
other alginate-producing bacteria including Pseudomonas fluorescens, P. putida,
and P. mendocina [20].
Regarding earliest applications of alginate, although some UK companies such
as British Algin Company Ltd (1885), Blandola Ltd (1908), and Liverpool Borax
Ltd (1909) were initially established to harness alginates, the first successful com-
pany in producing large and commercial scale pure alginates was Kelco Company
established by F.C. Thornley in San Diego, USA, in 1929 [21].
In 1934, Cefoil Ltd was established in the UK to exploit alginates extracted from
seaweeds for producing sodium alginate fiber applicable in camouflage netting and
other military items [21]. During that period of time, C. W. Bonniksen from the
Chemistry Department of University College in London attempted to investigate
alginic acid for production of cellophane-like material. For this purpose, Bonniksen
and his colleagues established a small company and named it The Kintyre factory.
In 1939, they relatively achieved the production of cellophane-like material from
alginic acid [3]. But, their operation was coincident with triggering World War II,
and the fate of alginates application was directed by the war. Accordingly, requested
by the Government, Bonniksen and his colleagues were responsible for establishing
more factories in that region in order to extract alginic acid for production of cam-
ouflage textile. Also, in parallel, other research groups were pursuing the same pur-
poses in several other laboratories in the UK [3]. The Kintyre factory was closed in
1942, but its production was transferred to newly established factories. On the other
hand, after the war (1945) Cefoil Ltd was changed to Alginate Industries Ltd. Later,
the two largest alginate producers, i.e., Kelco Company (USA) and Alginate
Industries Ltd (UK), were acquired by Merck and Co. Inc., USA, respectively, in
1972 and 1979, becoming the largest alginate producer worldwide [21]. In the
1980s and 1990s, more companies in other countries such as Norway, Germany,
France, Japan, and China were established in particular close to abundant natural
seaweed resources.
and its C5 epimer α-L-guluronic acid (G) linking via 1,4-glycosidic bonds (Fig. 1.1).
In nature, alginates are usually found with heteropolymeric structure, i.e., combina-
tion of both M and G residues, while production of monopolymeric structure
(polyM) has been reported at initial stage of alginate polymerization in bacteria and
genetically manipulated P. aeruginosa via inactivating catalytic domain of alginate
epimerase (e.g., PDO300∆algG (pBBR1MCS-5: algG (D324A)) [22, 23]. The
occurrence of M and G residues in polymeric structure varies significantly among
alginates as variable numbers and lengths of M blocks, G blocks, and MG blocks.
Composition of alginates and molecular mass may differ significantly depending on
the source of production and growth condition of the producer. However, while
algal alginates usually show a high content of G blocks, alginate produced by P.
aeruginosa does not possess G blocks. Another significant structural modification is
natural acetylation of alginates at O-2 and/or O-3 positions which have been so far
reported only in bacterial alginates (Fig. 1.1), while acetylating algal alginates via
chemical treatments has also been reported [24, 25].
Composition of polymers determines their physicochemical properties. The
intrinsic viscoelasticity of alginates depends on the frequency of constituting blocks
as flexibility decreases in the order MG block > MM block > GG block. The most
important features of alginates are related to its ability to efficiently and selectively
bind divalent cations leading eventually to hydrogel formation and crosslinked
polymeric scaffolds [26] (Fig. 1.1). The affinity of alginates toward different diva-
lent ions was found to increase in the order Mg2+ << Mn2+ <Ca2+ <Sr2+ <Ba2+ <Cu2+
<Pb2+ [27, 28]. The strength, dimension, stability, and mechanical property of
resulting hydrogels differ based on the type of interacting cation, the G content, and
variability of G blocks in the polymer [26, 29–31].
Fig. 1.1 Chemical structure of alginates produced by various organisms including seaweeds and
bacteria belonging to the genera Pseudomonas and Azotobacter. G blocks occurring only in algal
alginates and Azotobacter spp. bind selectively with divalent cations such as calcium causing
hydrogel formation. Only bacterial alginates are being acetylated at C2 and C3 positions, leading to
increasing the interaction of polymer with water molecules and increasing water capacity and
polymer extension
1 Alginate Biosynthesis and Biotechnological Production 5
The presence of O-acetyl groups in bacterial alginates notably changes the prop-
erties of the alginates reflecting at polymer conformation, chain expansion,
solubility, water capacity, viscoelasticity, and molecular mass. An acetylated algi-
nate absorbs more water due to the better interaction of chains with water mole-
cules, leading to chain expansion and better solubility [22, 24, 32, 33].
Unique composition and properties of alginates led to their wide applications in
various industries including agriculture, food, textile, cosmetic, and pharmaceuti-
cal/biomedical industries. These natural polymers have been considered as thicken-
ers, stabilizers, viscosifiers, additives, gel and film formers, and fertilizers [2,
34–37]. Owing to non-toxicity, biocompatibility, non-immunogenicity, hydrophi-
licity, and biodegradability, alginates have been extensively applied for biomedical
and pharmaceutical uses [38–40]. They have been traditionally applied for generat-
ing materials in dental impression and wound dressing [41–44]. However, techno-
logical advancements such as the materials fabrication and processing technologies
have enhanced their biomedical applications via modifications and tailoring of algi-
nate compositions including interactions with other polysaccharides. Alginates have
been processed into nanoparticles, nanotubes, microspheres, microcapsules,
sponges, hydrogels, foams, elastomers, fibers, etc. [45–51]. Nowadays, various
types of alginate derivatives are considered as one of the highly valuable and bio-
compatible biopolymers for drug delivery [2, 37]; immobilization of enzymes [52–
54]; cancer therapy by functionalizing polymeric scaffolds for controlled release of
anticancer drugs [55, 56]; therapeutic cell entrapment [57–59]; protection of trans-
planted cells from the host immune system [60–62]; tissue engineering [63–65];
generation of three-dimensional cell culture matrices for different laboratory assess-
ments such as cell-drug interaction, cell growth, and cell biology [66–68]; and algi-
nate formulations for preventing gastric reflux [69, 70]. The next chapters will
present the current state-of-the-art review for biomedical application of alginates
and their derivatives with regard to drug delivery (Chaps. 3 and 4); tissue engineer-
ing and cell therapy (Chaps. 4 and 5); tumor studies (Chap. 6); heart and cardiovas-
cular diseases (Chap. 7); dressings and wound management (Chap. 8); metabolic
syndromes (Chap. 9); chronic lung diseases, biofilm infections, and antibiotic use
(Chap. 10); and developing anti-inflammatory drugs (Chap. 11).
To date, seaweeds have been the sole and relatively low-cost alginate producers for
all commercial purposes. These natural resources have been mainly brown algae
from the genera Laminaria, Macrocystis, Ascophyllum, Ecklonia, Lessonia, and
Durvillaea.
Variable composition of alginates is linked to their natural biological role for
producer (Table 1.1). Hence, seasonal and growth condition as well as geographical
distribution is critical for determining the composition of alginates as well as the
percentage of alginates in various parts of the algae. Generally speaking, apparently
6 M.F. Moradali et al.
Fig. 1.2 Biosynthesis pathway of alginates in bacteria and algae. This pathway is mainly under-
stood in bacteria P. aeruginosa and A. vinelandii, and some homologous genes which have been
hypothetically reported in brown alga Ectocarpus siliculosus (based on reference [113]) are pre-
sented. Gene clusters encoding different proteins accomplishing different steps of alginate biosyn-
thesis are presented and functionally assigned in the frame (TCA cycle the tricarboxylic acid cycle,
MPI mannose-6-phosphate isomerase, PMM phosphomannomutase, MPG mannose-1-phosphate
guanylyltransferase, GMD GDP-mannose 6-dehydrogenase, MC5E mannuronate C5-epimerase,
PolyM poly-mannuronate, polyMG poly-mannuronate/guluronate)
are variably acetylated, and contrary to Azotobacter alginates which contain all
types of block structures, Pseudomonas alginates only possess M and MG blocks,
but not G blocks, indicating their different biological role for different species
(Fig. 1.1, Table 1.1).
For many years, understanding the biosynthesis of alginates has been of great
importance for scientific community in order to inform drug development for treat-
ment of P. aeruginosa infections exacerbated by alginate overproduction as well as
establishing the production of bacterial and tailor-made alginates. This is particu-
larly important for high-value purposes which require defined properties of algi-
nates. Therefore, in contrast to biosynthesis of algal alginates, alginate biosynthesis,
modification, and secretion events are relatively well understood in bacteria.
Generally, P. aeruginosa and A. vinelandii share a similar biosynthesis gene
cluster which is conserved among alginate producing bacteria (Fig. 1.2). However
they differ in regard to epimerization as well as regulatory mechanisms. Most
genes involved in alginate production are clustered in bacterial genomes. Except
for algC, the genes including algD, alg8, alg44, algK, algE (algJ), algG, algX,
algL, algI, algJ (algV), algF, algA are clustered within the alginate operon [81, 82]
(gene names in parentheses correspond gene names in Azotobacter). Transcription
of these genes is under the tight control of a promoter upstream of algD (Fig. 1.2)
[83, 84] and two internal promoters [85]. Hence, resulting differential transcrip-
tion of downstream genes to internal promoters was proposed as a mechanism
which may control the stoichiometry of protein subunits within the multiprotein
complex resulting in varying composition of produced alginates under different
environmental conditions [85].
The synthesis of alginate starts with the provision of the active precursor guanosine
di-phosphate (GDP)-mannuronic acid. This requires a series of cytosolic enzymatic
steps mediated by AlgA (phosphomannoseisomerase/GDP-mannose), AlgC (phos-
phomannomutase), and AlgD (GDP-mannose dehydrogenase) which catalyze four
biosynthesis steps to convert fructose-6-phosphate originating from the gluconeo-
genesis pathway to GDP-mannuronic acid [86–90] (Fig. 1.2). The last enzymatic
event catalyzed by AlgD, leading to GDP-mannuronic acid formation, is irrevers-
ible and is thought to be a key rate-limiting reaction in the alginate synthesis path-
way (Fig. 1.2) [91–94].
1 Alginate Biosynthesis and Biotechnological Production 9
1.3.2 P
rotein-Protein Interactions Constituting Alginate
Biosynthesis/Modification/Secretion Multi-protein
Complex
Fig. 1.3 Proposed model of alginate biosynthesis machinery complex and experimentally demon-
strated protein-protein interactions (marked with white triangles). This model shows alginate pro-
duction is positively regulated by c-di-GMP binding to Alg44 (S1) which targets the catalytic site
of Alg8 polymerase (S2). Then, translocation across the periplasmic scaffold is coupled with modi-
fication events (S3 to S5). AlgL is responsible for degrading misguided alginate accumulating in
the periplasm (S6). MucD protein links the complex with the posttranslational alginate regulatory
network via an interaction with AlgX (OM outer membrane, CM cytoplasmic membrane (Adapted
from Ref. [22]))
1 Alginate Biosynthesis and Biotechnological Production 11
1.3.4 Epimerization
1.4 A
lginate Polymerizing and Modifying Enzymes
and Their Uses for Tailor-Made Alginate Production
Hitherto, alginates extracted from brown algae have been solely applied for generat-
ing desired alginate derivatives mainly via in vitro chemical modifications and treat-
ments. However, since such approaches are often less controllable, achieving the
alginates with specific and defined physicochemical properties is hard and some-
times impossible. Furthermore, in vitro chemical approaches may result in unde-
sired changes in other polymer characteristics such as degradation of polymer chain.
Hence, understanding molecular mechanisms of alginate polymerization/modifica-
tion and unraveling their correlation as well as functional relationships of involved
proteins can build up a critical foundation for production of tailor-made alginates.
In a recent study, the functional and structural relationship of these mechanisms was
investigated and resulted in the production of various alginates from engineered P.
aeruginosa (Fig. 1.4) [22]. For many years, it was known that Alg8 (polymerase)/
Alg44 (co-polymerase) and AlgX (acetyltransferase)/AlgG (epimerase) are respon-
sible for alginate polymerization and modifications, respectively, while their possi-
ble interaction to constitute the proposed multi-protein complex remained unclear.
However, functional relationships of these mechanisms and proteins were based on
assumptions and indirect evidences. In recent years, studies have shed light on func-
tional and structural relationships of proteins involved in constituting the alginate
biosynthesis/modification/secretion multiprotein complex. Studies showed that the
processivity of alginate polymerization was interrupted by epimerization, resulting
in alginates with lower molecular mass (2755 kDa) (Fig. 1.4, Table 1.2) [22]. Upon
removal of epimerization by generating a catalytically inactive variant of AlgG
(Fig. 1.4 No. 8, Table 1.2), processivity of alginate polymerization was increased
leading to a very high molecular mass alginate (4653 kDa) (Figs. 1.4 and 1.5,
Table 1.2) [22]. Furthermore, the productivity of alginate (yield) was about three-
fold higher when epimerization was eliminated when compared with the presence
of epimerization. Hence, it was concluded that polymerization has a negative
14 M.F. Moradali et al.
Fig. 1.4 Impact of putative alginate polymerase subunits on alginate polymerase activity, alginate
polymerization, and composition and correlation between polymerization and modification. (a)
The values of molar fraction of G residue (FG), acetylation degrees (Ac. %), mean molecular
masses, and alginate yield are aligned with the strains producing the respective alginates. (b)
Correlation between degree of acetylation, epimerization, and molecular mass of alginate.
Presumable features (No. 1 to 11) show protein complexes constituted by Alg8, Alg44, AlgG, and
AlgX (see the legend at the top left corner of the plot). The subunit produced upon in trans comple-
mentation is shown as darker shape(s). Inactive AlgX(S269A) and AlgG(D324A) proteins are
labeled as (Ac) and (Ep), respectively. The length of various alginates (PD) with respect to acetyla-
tion (Ac. %) and epimerization (FG) degrees are presented and proportionally illustrated for each
feature (300 PDO300, MCS5 pBBR1MCS-5, PD polymerization degree, OM outer membrane,
CM cytoplasmic membrane (Adapted from Ref. [22]))
correlation with epimerization event [22]. Furthermore, evidence was provided that
elimination of acetylation by producing a catalytically inactive variant of AlgX (i.e.,
AlgX (S269A)) resulted in lowering the molecular mass (2086 kDa) (No. 6 in
Fig. 1.4, Table 1.2), while upon its presence higher molecular mass (2460 kDa) was
produced [22]. This result showed that acetylation did not disrupt processivity of
alginate polymerization indicative of a positive correlation. Importantly, epimeriza-
tion and acetylation did not show any competitive relationship, contrary to previous
assumptions, and removal of one (i.e., replacement with catalytically inactive vari-
ants of AlgG or AlgX) did not impact the other one (Table 1.2) [22]. Interestingly,
AlgX and AlgG proteins showed mutual auxiliary behavior as the overproduction of
AlgX boosted epimerization (FG = 0.36; acetylation = 9.8%) and AlgG overproduc-
tion increased the acetylation degree (FG = 0.32; acetylation = 23.3%) (Table 1.2).
However, the elimination of two modification events resulted in the alginate with
Table 1.2 Composition and molecular mass analyses of alginates produced by different P. aeruginosa (Pa) PDO300 mutants complemented with plasmid-
borne genes encoding Alg8 (polymerase), Alg44 (co-polymerase), AlgX (acetyltransferase), AlgG (epimerase), and inactive variantsa
Alginate yield
Mw Mn
Mutant FG FM FGM/MG FMM Ac.% (kDa) (kDa) pI (g)/ CDM(g)
1 Wild type (Pa) +empty plasmid 0.3 0.7 0.29 0.41 32 3927 (±0.864%) 3832 (±0.842%) 1.025 (±1.2%) 1.3 ± 0.03
2 Pa ∆alg8+ MCS5:alg8 0.18 0.82 0.17 0.65 11.3 3045 (±0.556%) 3037 (±0.551%) 1.003 (±0.7%) 12.8 ± 1.03
3 Pa ∆alg44+ MCS5:alg44 0.18 0.82 0.18 0.64 26.8 3831 (±0.963%) 3650 (±0.950%) 1.05 (±1.3%) 8.7 ± 0.53
4 Pa ∆alg44∆alg8+ MCS5:alg44:alg8 0.17 0.83 0.17 0.66 11 3369 (±0.839%) 3352 (±0.821%) 1.005 (±1.1%) 41.5 ± 4.9
5 Pa ∆algX+ MCS5:algX 0.36 0.64 0.36 0.28 9.8 2460 (±0.932%) 2447 (±0.913%) 1.005 (±1.3%) 104.1 ± 5.5
6 Pa ∆algX+ MCS5:algX(S269A)† 0.36 0.64 0.36 0.28 0 2086 (±0.960%) 2065 (±0.944%) 1.010 (±1.3%) 125.8 ± 9.9
7 Pa ∆algG+ MCS5:algG 0.32 0.68 0.32 0.36 23.3 2755 (±1.041%) 2726 (±0.986%) 1.011 (±1.4%) 2.6 ± 0.04
8 Pa ∆algG+ MCS5:algG(D324A)‡ 0 1 0 1 25.2 4653 (±1.097%) 4575 (±1.117%) 1.017 (±1.5%) 7.6 ± 0.57
9 Pa ∆algX∆algG+ MCS5:algX:algG 0.34 0.66 0.34 0.32 28.4 3076 (±1.051%) 3044 (±1.029%) 1.011 (±1.4%) 67.42 ± 4.8
10 Pa ∆algX∆algG+ 0 1 0 1 0 1811 (±0.884%) 1716 (±0.888%) 1.055 (±1.2%) 8.7 ± 0.42
1 Alginate Biosynthesis and Biotechnological Production
MCS5:algX(S269A): algG(D324A)
11 Pa ∆alg44∆algG+ 0.22 0.78 0.22 0.56 14.5 2907 (±0.966%) 2861 (±0.944%) 1.016 (±1.3%) 9.0 ± 0.3
MCS5:alg44+algG
300 PDO300 strain, MCS5 pBBR1MCS-5 plasmid, FG molar fraction of guluronate (G) residue, FM molar fraction of mannuronate (M) residue, FGM/MG molar
fraction of two consecutive G and M residues, FMM molar fraction of two consecutive M residues, Ac. acetylation, M w weight-average molecular weights, M n
number-average molecular weights, pI polydispersity index, CDM cell dry mass. † catalytically inactive variant of AlgX acetyltransferase, ‡ catalytically inac-
tive variant of AlgG epimerase
a
Adapted from Ref. [22]
15
16 M.F. Moradali et al.
Fig. 1.5 Based on experimental results, interactive performances of protein functionality over
alginate polymerization, acetylation, epimerization, and length determination are modeled. In this
model, translocation of polymerized nascent alginate across the periplasmic scaffold is coupled
with interactive functional performances of modification events where alginate molecular mass or
polymerization is inversely correlated with alginate epimerization but positively correlated with
acetylation. Also, Alg44 boost acetylation and AlgG and AlgX proteins display mutual auxiliary
function for each other (Adapted from reference [22])
the lowest molecular mass (1811 kDa vs. 3076 kDa when both modification events
present) (No. 10 in Figs. 1.4, 1.5, and Table 1.2).
Furthermore, the overproduction of Alg8 and Alg44 impacted on polymerization
event by increasing the ratio of M residue and M blocks, while Alg44 boosted the
acetylation degree (Figs. 1.4, 1.5, and Table 1.2) [22, 103].
Importantly, all tested strains harboring various combinations of Alg8, Alg44,
AlgG, and AlgX and their catalytically inactive variants produced monodisperse
alginates (a monodisperse polymer has a polydispersity index equal or close to 1.0)
[22]. This study showed that bacterial alginate is an ideal source of monodisperse
alginates versus algal alginates which are polydisperse.
Also, particle tracking microrheology was applied to assess the viscoelastic
properties of the various resulting alginates [22]. All alginates showed viscoelastic
properties in which the solid-like elastic modulus G′ was greater than the liquid-like
viscous modulus G″ (G′ > G″), but significantly different from each other. Generally,
the alginates without G residues or with the highest molar fraction of MM blocks
1 Alginate Biosynthesis and Biotechnological Production 17
(FM = 0.82 to 1.0), which possessed higher molecular mass, showed the highest and
quite similar viscoelastic properties (G′ = 0.41, G″ = 0.28–0.3) [22]. Lower visco-
elastic properties were found for alginates with a molecular mass of ≤2000 kDa.
Surprisingly, introducing high copy of alginate acetyltransferase AlgX in P. aerugi-
nosa resulted in alginate with the lowest viscoelastic property among all analyzed
samples because of boosting epimerization event and higher occurrence of G resi-
dues [22]. These results suggested that viscoelasticity was positively impacted by
the molecular mass combined with high M content, while the presence of G resi-
dues and acetyl groups in the alginate chain lowered viscoelasticity [22].
Overall, this study had demonstrated the production of various alginates through
manipulating the activity and production of various protein subunits in bacteria.
Application of these biopolymers for specific purposes particularly in the context
of biomedical application and generation of high-value products may require spe-
cific and desired properties of alginates. Therefore, molar fraction of M and G resi-
dues, the ratio of M and G blocks, acetylation degree, molecular mass (or
polymerization degree), polydispersity, viscoelasticity, and other physicochemical
properties are critical parameters which are obtainable by engineering bacterial
alginate producers.
Another important step in the production of tailor-made alginate is to understand
minimal protein requirements for alginate polymerization and modifications. For
example, bacterial alginate polymerization is necessarily activated by the second
messenger c-di-GMP, and MucR protein is specific c-di-GMP provider in alginate
biosynthesis. Our recent analysis showed that site-specific mutagenesis of Alg8 at
specific amino acid residues surrounding the catalytic sites could decouple alginate
polymerization from MucR activity and presumably from proposed pool of c-di-
GMP in the cell [103]. This finding may facilitate the production of alginate in
nonpathogenic bacteria which may not generate the required localized pool of c-di-
GMP for inducing alginate production. In addition, the production of polymannuro-
nate (PolyM) alginates may be achievable by establishing only the polymerizing
unit (Alg8-Alg44) as the minimal protein requirement in suitable Gram-positive
bacteria which do not possess the outer membrane and the periplasm.
Production of alginates may be achievable through in vitro settings as previously
reported [128]. However, tailoring alginates by combining minimal protein require-
ments alone or within particular lipid scaffolds such as proteoliposomes, lipid rafts,
lipid discs, and inverted membrane vesicles may be approachable. Likewise, the
combination of various alginate-modifying enzymes produced by various organ-
isms will expand their applicability as well as the production of tailor-made algi-
nates. For example, by applying the secreted mannuronan C-5-epimerases in in vitro
reactions, the production of alginates with long G blocks and/or replacing stretches
of M blocks with MG blocks has been achievable [129]. To this end, understanding
the molecular mechanisms of various alginate-modifying enzymes such as alginate
lyases and epimerases will provide further opportunities for bioengineering toward
tailored alginates.
18 M.F. Moradali et al.
Alginates have been one of the most widely applied polysaccharides with uses in
various industries due to their unique physicochemical properties. Traditionally,
they have been largely utilized in food, cosmetic, and pharmaceutical industries as
thickeners, stabilizers, viscosifiers, additives, gel and film formers, and fertilizers.
Of the most relevant criteria which have made alginates particularly suitable for
biomedical and pharmaceutical purposes are associated with their non-toxicity, bio-
compatibility, inertness, and modifiability. Hence, nowadays, various types of algi-
nates and their derivatives have earned their reputation in drug delivery, cell
encapsulation, and enzyme immobilization. In addition, technological advancement
in three-dimensional printing and material fabrication as well as processing tech-
nologies harnessed unique properties of alginates for expanding their application
for biomedical purposes. However, development of some high-value products from
alginates and their derivatives may require defined compositions and material prop-
erties which might not exist in algal alginates or may not be achievable via chemical
modifications. Therefore, bacterial alginates display different characteristics from
algal alginates such as acetylation, higher molecular mass, monodispersity, differ-
ent viscoelastic properties, and possibility to engineer the producer to produce tai-
lored alginates with specifications for advanced biomedical uses. Most importantly,
in contrast to algal alginates whose biosynthesis pathways and production in con-
trolled environments have not been achieved yet, obtaining desired alginates from
bacteria is achievable in a controlled environment (e.g., bioreactor) without exces-
sive effort while eliminating oceanic impurities coming from algal resources.
Understanding functional and structural relationships of various protein subunits
involved in alginate polymerization and modification is important for establishing
bacterial production of tailored alginates. Recently we demonstrated production of
various alginates via engineering P. aeruginosa. The unraveled interplay of polym-
erization with epimerization and acetylation offers bioengineering opportunities to
produce various alginate compositions, molecular masses, and viscoelastic proper-
ties, while retaining monodispersity. However, bacterial production of alginates is
still at early stage, as the molecular mechanism of biosynthesis/modification path-
ways need to be fully elucidated. In addition, these pathways are tightly controlled
by bacterial regulatory systems which may be an obstacle for establishment of bio-
technological production of alginates. Therefore, understanding the minimal pro-
tein requirements for bacterial production of alginates in homologous and
heterologous hosts is of particular importance.
Currently most knowledge about alginate production was obtained in opportu-
nistic human pathogen P. aeruginosa. Therefore, establishing the production of
various alginates by nonpathogenic bacteria and non-virulent strains must be con-
sidered. Another potential challenge could be the purification process associated
with alginates produced by bacteria. Existing commercial methods applied for puri-
fication of algal alginates to eliminate unwanted and immunogenic impurities may
not be sufficient for removing impurities from bacterial production process. Overall,
1 Alginate Biosynthesis and Biotechnological Production 19
bacterial alginates have largely remained unexplored for advanced biomedical uses
and the production of high-value products. They exhibit different material proper-
ties when compared to algal alginates and can be tailored via bioengineering/syn-
thetic biology toward the production of novel and advanced alginates.
References
1. Stanford ECC (1883) On align: a new substance obtained from some of the commoner spe-
cies of marine algae. Chem News 47:254–257
2. Lesser MA (1947) Alginates in drugs and cosmetics. Drug Cosmet Ind 61(6):761–842
3. Woodward F (1951) The Scottish seaweed research association. J Mar Biol Assoc UK
29(03):719–725
4. Steiner AB, McNeely WH (1951) Organic derivatives of alginic acid. Ind Eng Chem
43(9):2073–2077
5. Krefting A (1896) An improved method of treating seaweed to obtain valuable products
therefrom. Br Patent 11:538
6. Krefting A (1898) Axel krefting. Google Patents
7. Atsuki K, Tomoda Y (1926) Studies on seaweeds of Japan I. The chemical constituents of
Laminaria. J Soc Chem Ind Japan 29:509–517
8. Nelson WL, Cretcher LH (1929) The alginic acid from macrocystis pyrifera. J Am Chem Soc
51(6):1914–1922
9. Nelson WL, Cretcher LH (1930) The isolation and identification of D-mannuronic acid lac-
tone from the Macrocystis pyrifera. J Am Chem Soc 52(5):2130–2132
10. Nelson WL, Cretcher LH (1932) The properties of D-mannuronic acid lactone. J Am Chem
Soc 54(8):3409–3412
11. Bird GM, Haas P (1931) On the nature of the cell wall constituents of Laminaria spp.
Mannuronic acid. Biochem J 25(2):403
12. Miwa T (1930) Alginic acid. J Chem Soc Japan 51:738–745
13. Schoeffel E, Link KP (1933) Isolation of α-and β, D-Mannuronic acid. J Biol Chem
100(2):397–405
14. Astbury W (1945) Structure of alginic acid. Nature 155:667–668
15. Hirst E, Jones J, Jones WO (1939) Structure of alginic acid. Nature 143:857
16. Fischer F, Dörfel H (1955) Die polyuronsäuren der braunalgen (Kohlenhydrate der Algen I).
Hoppe-Seyler’s Zeitschrift für physiologische Chemie 302(1-2):186–203
17. Linker A, Jones RS (1964) A polysaccharide resembling alginic acid from a Pseudomonas
microorganism. Nature 204:187–188
18. Linker A, Jones RS (1966) A new polysaccharide resembling alginic acid isolated from
Pseudomonads. J Biol Chem 241(16):3845–3851
19. Gorin P, Spencer J (1966) Exocellular alginic acid from Azotobacter vinelandii. Can J Chem
44(9):993–998
20. Govan JR, Fyfe JA, Jarman TR (1981) Isolation of alginate-producing mutants of
Pseudomonas fluorescens, Pseudomonas putida and Pseudomonas mendocina. J Gen
Microbiol 125(1):217–220
21. Clare K (1993) Algin. Ind Gums:105–143
20 M.F. Moradali et al.
22. Moradali MF, Donati I, Sims IM, Ghods S, Rehm BH (2015) Alginate polymerization and
modification are linked in Pseudomonas aeruginosa. MBio 6(3):e00453-00415
23. Douthit SA, Dlakic M, Ohman DE, Franklin MJ (2005) Epimerase active domain of
Pseudomonas aeruginosa AlgG, a protein that contains a right-handed β-helix. J Bacteriol
187(13):4573–4583
24. SkjÅk-Bræk G, Paoletti S, Gianferrara T (1989) Selective acetylation of mannuronic acid
residues in calcium alginate gels. Carbohydr Res 185(1):119–129
25. Windhues T, Borchard W (2003) Effect of acetylation on physico-chemical properties of bac-
terial and algal alginates in physiological sodium chloride solutions investigated with light
scattering techniques. Carbohydr Polym 52(1):47–52
26. Mørch ÝA, Donati I, Strand BL, Skjåk-Bræk G (2006) Effect of Ca2+, Ba2+, and Sr2+ on algi-
nate microbeads. Biomacromolecules 7(5):1471–1480
27. Haug A, Smidsrod O (1970) Selectivity of some anionic polymers for divalent metal ions.
Acta Chem Scand 24(3):843–854
28. Haug A, Smidsrød O (1967) Strontium–calcium selectivity of alginates. Nature
215(5102):757–757
29. Ouwerx C, Velings N, Mestdagh M, Axelos M (1998) Physico-chemical properties and
rheology of alginate gel beads formed with various divalent cations. Polym Gels Networks
6(5):393–408
30. Braccini I, Pérez S (2001) Molecular basis of Ca2+-induced gelation in alginates and pectins:
the egg-box model revisited. Biomacromolecules 2(4):1089–1096
31. Sikorski P, Mo F, Skjåk-Bræk G, Stokke BT (2007) Evidence for egg-box-compatible
interactions in calcium− alginate gels from fiber X-ray diffraction. Biomacromolecules
8(7):2098–2103
32. Straatmann A, Windhues T, Borchard W (2004) Effects of acetylation on thermodynamic
properties of seaweed alginate in sodium chloride solutions. In: Analytical ultracentrifuga-
tion VII. Springer, Berlin, pp 26–30
33. Delben F, Cesaro A, Paoletti S, Crescenzi V (1982) Monomer composition and acetyl content
as main determinants of the ionization behavior of alginates. Carbohydr Res 100(1):C46–C50
34. Onsøyen E (1997) Alginates. In: Thickening and gelling agents for food. Springer, Boston,
pp 22–44
35. McHugh DJ (1987) Production, properties and uses of alginates. Production and utilization
of products from commercial seaweeds. FAO Fish Tech Pap 288:58–115
36. Smith AM, Miri T (2010) 6 alginates in foods. Practical food rheology: an interpretive
approach:113
37. Tønnesen HH, Karlsen J (2002) Alginate in drug delivery systems. Drug Dev Ind Pharm
28(6):621–630
38. Skaugrud Ø, Hagen A, Borgersen B, Dornish M (1999) Biomedical and pharmaceutical
applications of alginate and chitosan. Biotechnol Genet Eng Rev 16(1):23–40
39. Bhattarai N, Li Z, Edmondson D, Zhang M (2006) Alginate-based nanofibrous scaffolds:
structural, mechanical, and biological properties. Adv Mater 18(11):1463–1467
40. Douglas KL, Piccirillo CA, Tabrizian M (2006) Effects of alginate inclusion on the vector
properties of chitosan-based nanoparticles. J Control Release 115(3):354–361
41. Cook W (1986) Alginate dental impression materials: chemistry, structure, and properties.
J Biomed Mater Res 20(1):1–24
42. Craig R (1988) Review of dental impression materials. Adv Dental Res 2(1):51–64
43. Groves A, Lawrence J (1986) Alginate dressing as a donor site haemostat. Ann R Coll Surg
Engl 68(1):27
44. Barnett S, Varley S (1987) The effects of calcium alginate on wound healing. Ann R Coll
Surg Engl 69(4):153
45. Hrynyk M, Martins-Green M, Barron AE, Neufeld RJ (2012) Alginate-PEG sponge architec-
ture and role in the design of insulin release dressings. Biomacromolecules 13(5):1478–1485
1 Alginate Biosynthesis and Biotechnological Production 21
46. Barbetta A, Barigelli E, Dentini M (2009) Porous alginate hydrogels: synthetic methods for
tailoring the porous texture. Biomacromolecules 10(8):2328–2337
47. Andersen T, Melvik JE, Gåserød O, Alsberg E, Christensen BE (2012) Ionically gelled algi-
nate foams: physical properties controlled by operational and macromolecular parameters.
Biomacromolecules 13(11):3703–3710
48. Shin S-J, Park J-Y, Lee J-Y, Park H, Park Y-D, Lee K-B, Whang C-M, Lee S-H (2007) “On
the fly” continuous generation of alginate fibers using a microfluidic device. Langmuir
23(17):9104–9108
49. Daemi H, Barikani M, Barmar M (2013) Highly stretchable nanoalginate based polyurethane
elastomers. Carbohydr Polym 95(2):630–636
50. Senuma Y, Lowe C, Zweifel Y, Hilborn J, Marison I (2000) Alginate hydrogel microspheres
and microcapsules prepared by spinning disk atomization. Biotechnol Bioeng 67(5):616–622
51. Bodmeier R, Chen H, Paeratakul O (1989) A novel approach to the oral delivery of micro-or
nanoparticles. Pharm Res 6(5):413–417
52. Kierstan M, Bucke C (1977) The immobilization of microbial cells, subcellular organelles,
and enzymes in calcium alginate gels. Biotechnol Bioeng 19(3):387–397
53. Palmieri G, Giardina P, Desiderio B, Marzullo L, Giamberini M, Sannia G (1994) A new
enzyme immobilization procedure using copper alginate gel: application to a fungal phenol
oxidase. Enzym Microb Technol 16(2):151–158
54. Fukushima Y, Okamura K, Imai K, Motai H (1988) A new immobilization technique of whole
cells and enzymes with colloidal silica and alginate. Biotechnol Bioeng 32(5):584–594
55. Zhang W, Zhang Z, Zhang Y (2011) The application of carbon nanotubes in target drug deliv-
ery systems for cancer therapies. Nanoscale Res Lett 6(1):555
56. Barreto JA, O’Malley W, Kubeil M, Graham B, Stephan H, Spiccia L (2011) Nanomaterials:
applications in cancer imaging and therapy. Adv Mater 23(12)
57. Serp D, Cantana E, Heinzen C, Von Stockar U, Marison I (2000) Characterization of an
encapsulation device for the production of monodisperse alginate beads for cell immobiliza-
tion. Biotechnol Bioeng 70(1):41–53
58. Baruch L, Machluf M (2006) Alginate–chitosan complex coacervation for cell encapsulation:
effect on mechanical properties and on long-term viability. Biopolymers 82(6):570–579
59. Orive G, Hernandez R, Gascon A, Igartua M, Pedraz J (2003) Survival of different cell lines
in alginate-agarose microcapsules. Eur J Pharm Sci 18(1):23–30
60. Shapiro L, Cohen S (1997) Novel alginate sponges for cell culture and transplantation.
Biomaterials 18(8):583–590
61. de Vos P, Faas MM, Strand B, Calafiore R (2006) Alginate-based microcapsules for immuno-
isolation of pancreatic islets. Biomaterials 27(32):5603–5617
62. Kulseng B, Skjåk-Bræk G, Ryan L, Andersson A, King A, Faxvaag A, Espevik T (1999)
Transplantation of alginate microcapsules: generation of antibodies against alginates and
encapsulated porcine islet-like cell clusters. Transplantation 67(7):978–984
63. Li Z, Ramay HR, Hauch KD, Xiao D, Zhang M (2005) Chitosan–alginate hybrid scaffolds
for bone tissue engineering. Biomaterials 26(18):3919–3928
64. Kuo CK, Ma PX (2001) Ionically crosslinked alginate hydrogels as scaffolds for tis-
sue engineering: part 1. Structure, gelation rate and mechanical properties. Biomaterials
22(6):511–521
65. Drury JL, Mooney DJ (2003) Hydrogels for tissue engineering: scaffold design variables and
applications. Biomaterials 24(24):4337–4351
66. Lee J, Cuddihy MJ, Kotov NA (2008) Three-dimensional cell culture matrices: state of the
art. Tissue Eng B Rev 14(1):61–86
67. Perka C, Spitzer RS, Lindenhayn K, Sittinger M, Schultz O (2000) Matrix-mixed culture:
new methodology for chondrocyte culture and preparation of cartilage transplants. J Biomed
Mater Res A 49(3):305–311
68. Murphy WL, Mooney DJ (1999) Controlled delivery of inductive proteins, plasmid DNA and
cells from tissue engineering matrices. J Periodontal Res 34(7):413–419
22 M.F. Moradali et al.
69. Kwiatek MA, Roman S, Fareeduddin A, Pandolfino JE, Kahrilas PJ (2011) An alginate-
antacid formulation (Gaviscon Double Action Liquid) can eliminate or displace the postpran-
dial ‘acid pocket’ in symptomatic GERD patients. Aliment Pharmacol Ther 34(1):59–66
70. Washington N (1990) Investigation into the barrier action of an alginate gastric reflux sup-
pressant, liquid Gaviscon®. Drug Investig 2(1):23–30
71. Andresen I-L, Smidsørod O (1977) Temperature dependence of the elastic properties of algi-
nate gels. Carbohydr Res 58(2):271–279
72. Indergaard M, Skjåk-Bræk G (1987) Characteristics of alginate from Laminaria digitata
cultivated in a high-phosphate environment. In: Twelfth international seaweed symposium,
Springer, pp 541–549
73. Kloareg B, Quatrano R (1988) Structure of the cell walls of marine algae and ecophysiologi-
cal functions of the matrix polysaccharides. Oceanogr Mar Biol 26:259–315
74. Lin T-Y, Hassid W (1966) Pathway of alginic acid synthesis in the marine brown alga, Fucus
gardneri Silva. J Biol Chem 241(22):5284–5297
75. Haug A, Larsen B (1969) Biosynthesis of alginate. Epimerisation of D-mannuronic
to L-guluronic acid residues in the polymer chain. Biochim Biophy Acta Genl Subj
192(3):557–559
76. Ryder C, Byrd M, Wozniak DJ (2007) Role of polysaccharides in Pseudomonas aeruginosa
biofilm development. Curr Opin Microbiol 10(6):644–648
77. Moradali MF, Ghods S, Rehm BHA (2017) Pseudomonas aeruginosa lifestyle: a paradigm
for adaptation, survival, and persistence. Front Cell Infect Microbiol 7:39
78. Ghafoor A, Hay ID, Rehm BH (2011) Role of exopolysaccharides in Pseudomonas aerugi-
nosa biofilm formation and architecture. Appl Environ Microbiol 77(15):5238–5246
79. Clementi F (1997) Alginate production by Azotobacter vinelandii. Crit Rev Biotechnol
17(4):327–361
80. Costerton JW, Cheng KJ, Geesey GG, Ladd TI, Nickel JC, Dasgupta M, Marrie TJ (1987)
Bacterial biofilms in nature and disease. Annu Rev Microbiol 41:435–464
81. Chitnis CE, Ohman DE (1993) Genetic analysis of the alginate biosynthetic gene clus-
ter of Pseudomonas aeruginosa shows evidence of an operonic structure. Mol Microbiol
8(3):583–590
82. Hay ID, Wang Y, Moradali MF, Rehman ZU, Rehm BH (2014) Genetics and regulation of
bacterial alginate production. Environ Microbiol 16(10):2997–3011
83. Schurr M, Martin D, Mudd M, Hibler N, Boucher J, Deretic V (1992) The algD promoter:
regulation of alginate production by Pseudomonas aeruginosa in cystic fibrosis. Cell Mol
Biol Res 39(4):371–376
84. Shankar S, Ye RW, Schlictman D, Chakrabarty A (1995) Exopolysaccharide alginate syn-
thesis in Pseudomonas seruginosa: enzymology and regulation of gene expression. Adv
Enzymol Relat Areas Mol Biol 70:221–255
85. Paletta JL, Ohman DE (2012) Evidence for two promoters internal to the alginate biosynthe-
sis operon in Pseudomonas aeruginosa. Curr Microbiol 65(6):770–775
86. Lynn A, Sokatch J (1984) Incorporation of isotope from specifically labeled glucose into algi-
nates of Pseudomonas aeruginosa and Azotobacter vinelandii. J Bacteriol 158(3):1161–1162
87. Narbad A, Russell N, Gacesa P (1987) Radiolabelling patterns in alginate of Pseudomonas
aeruginosa synthesized from specifically-labelled 14C-monosaccharide precursors. Microbios
54(220-221):171–179
88. May TB, Shinabarger D, Boyd A, Chakrabarty AM (1994) Identification of amino acid
residues involved in the activity of phosphomannose isomerase-guanosine 5′-diphospho-D-
mannose pyrophosphorylase. A bifunctional enzyme in the alginate biosynthetic pathway of
Pseudomonas aeruginosa. J Biol Chem 269(7):4872–4877
89. Zielinski NA, Chakrabarty AM, Berry A (1991) Characterization and regulation of the
Pseudomonas aeruginosa algC gene encoding phosphomannomutase. J Biol Chem
266(15):9754–9763
1 Alginate Biosynthesis and Biotechnological Production 23
90. Shinabarger D, Berry A, May TB, Rothmel R, Fialho A, Chakrabarty AM (1991) Purification
and characterization of phosphomannose isomerase-guanosine diphospho-D-mannose pyro-
phosphorylase. A bifunctional enzyme in the alginate biosynthetic pathway of Pseudomonas
aeruginosa. J Biol Chem 266(4):2080–2088
91. Hay ID, Rehman ZU, Moradali MF, Wang Y, Rehm BH (2013) Microbial alginate produc-
tion, modification and its applications. Microb Biotechnol 6(6):637–650
92. Roychoudhury S, May T, Gill J, Singh S, Feingold D, Chakrabarty AM (1989) Purification
and characterization of guanosine diphospho-D-mannose dehydrogenase. A key enzyme in
the biosynthesis of alginate by Pseudomonas aeruginosa. J Biol Chem 264(16):9380–9385
93. Tatnell PJ, Russell NJ, Gacesa P (1994) GDP-mannose dehydrogenase is the key regulatory
enzyme in alginate biosynthesis in Pseudomonas aeruginosa: evidence from metabolite stud-
ies. Microbiology 140(7):1745–1754
94. Tavares IM, Leitão JH, Fialho AM, Sá-Correia I (1999) Pattern of changes in the activity of
enzymes of GDP-D-mannuronic acid synthesis and in the level of transcription of algA, algC
and algD genes accompanying the loss and emergence of mucoidy in Pseudomonas aerugi-
nosa. Res Microbiol 150(2):105–116
95. Rehman ZU, Wang Y, Moradali MF, Hay ID, Rehm BH (2013) Insights into the assembly of
the alginate biosynthesis machinery in Pseudomonas aeruginosa. Appl Environ Microbiol
79(10):3264–3272
96. Hay ID, Schmidt O, Filitcheva J, Rehm BH (2012) Identification of a periplasmic AlgK–
AlgX–MucD multiprotein complex in Pseudomonas aeruginosa involved in biosynthesis and
regulation of alginate. Appl Microbiol Biotechnol 93(1):215–227
97. Franklin MJ, Douthit SA, McClure MA (2004) Evidence that the algI/algJ gene cassette,
required for O acetylation of Pseudomonas aeruginosa alginate, evolved by lateral gene
transfer. J Bacteriol 186(14):4759–4773
98. Oglesby LL, Jain S, Ohman DE (2008) Membrane topology and roles of Pseudomonas aeru-
ginosa Alg8 and Alg44 in alginate polymerization. Microbiology 154(6):1605–1615
99. Remminghorst U, Rehm BH (2006) Alg44, a unique protein required for alginate biosynthe-
sis in Pseudomonas aeruginosa. FEBS Lett 580(16):3883–3888
100. Remminghorst U, Hay ID, Rehm BH (2009) Molecular characterization of Alg8, a putative
glycosyltransferase, involved in alginate polymerisation. J Biotechnol 140(3):176–183
101. Merighi M, Lee VT, Hyodo M, Hayakawa Y, Lory S (2007) The second messenger bis-(3′-
5′)-cyclic-GMP and its PilZ domain-containing receptor Alg44 are required for alginate bio-
synthesis in Pseudomonas aeruginosa. Mol Microbiol 65(4):876–895
102. Hay ID, Remminghorst U, Rehm BH (2009) MucR, a novel membrane-associated regulator of
alginate biosynthesis in Pseudomonas aeruginosa. Appl Environ Microbiol 75(4):1110–1120
103. Moradali MF, Ghods S, Rehm BH (2017) Activation mechanism and cellular localization
of membrane-anchored alginate polymerase in Pseudomonas aeruginosa. Appl Environ
Microbiol 83:03499–03416
104. Smidsrød O, Glover R, Whittington SG (1973) The relative extension of alginates having
different chemical composition. Carbohydr Res 27(1):107–118
105. Jain S, Franklin MJ, Ertesvåg H, Valla S, Ohman DE (2003) The dual roles of AlgG in C-5-
epimerization and secretion of alginate polymers in Pseudomonas aeruginosa. Mol Microbiol
47(4):1123–1133
106. Gimmestad M, Sletta H, Ertesvåg H, Bakkevig K, Jain S, S-j S, Skjåk-Bræk G, Ellingsen
TE, Ohman DE, Valla S (2003) The Pseudomonas fluorescens AlgG protein, but not its
mannuronan C-5-epimerase activity, is needed for alginate polymer formation. J Bacteriol
185(12):3515–3523
107. Gimmestad M, Steigedal M, Ertesvåg H, Moreno S, Christensen BE, Espín G, Valla S
(2006) Identification and characterization of an Azotobacter vinelandii type I secretion
system responsible for export of the AlgE-type mannuronan C-5-epimerases. J Bacteriol
188(15):5551–5560
24 M.F. Moradali et al.
108. Ertesvåg H, Valla S (1999) The A modules of the Azotobacter vinelandii mannuronan-C-
5-epimerase AlgE1 are sufficient for both epimerization and binding of Ca2+. J Bacteriol
181(10):3033–3038
109. Ullrich MS, Schergaut M, Boch J, Ullrich B (2000) Temperature-responsive genetic loci in
the plant pathogen Pseudomonas syringae pv. glycinea. Microbiology 146(10):2457–2468
110. Bjerkan TM, Bender CL, Ertesvåg H, Drabløs F, Fakhr MK, Preston LA, Skjåk-Bræk G,
Valla S (2004) The Pseudomonas syringae genome encodes a combined mannuronan C-5-
epimerase and O-acetylhydrolase, which strongly enhances the predicted gel-forming prop-
erties of alginates. J Biol Chem 279(28):28920–28929
111. Ertesvåg H (2015) Alginate-modifying enzymes: biological roles and biotechnological uses.
Front Microbiol 6
112. Nyvall P, Corre E, Boisset C, Barbeyron T, Rousvoal S, Scornet D, Kloareg B, Boyen C
(2003) Characterization of mannuronan C-5-epimerase genes from the brown alga Laminaria
digitata. Plant Physiol 133(2):726–735
113. Michel G, Tonon T, Scornet D, Cock JM, Kloareg B (2010) The cell wall polysaccharide
metabolism of the brown alga Ectocarpus siliculosus. Insights into the evolution of extracel-
lular matrix polysaccharides in Eukaryotes. New Phytol 188(1):82–97
114. Tonon T, Rousvoal S, Roeder V, Boyen C (2008) Expression profiling of the mannuronan
C5 epimerase multigenic family in the brown alga Laminaria digitata (Phaeophyceae) under
biotic stress condition. J Phycol 44(5):1250–1256
115. Baker P, Ricer T, Moynihan PJ, Kitova EN, Walvoort MT, Little DJ, Whitney JC, Dawson K,
Weadge JT, Robinson H (2014) P. aeruginosa SGNH hydrolase-like proteins AlgJ and AlgX
have similar topology but separate and distinct roles in alginate acetylation. PLoS Pathog
10(8):e1004334
116. Franklin MJ, Ohman DE (2002) Mutant analysis and cellular localization of the AlgI, AlgJ,
and AlgF proteins required for O-acetylation of alginate in Pseudomonas aeruginosa.
J Bacteriol 184(11):3000–3007
117. Franklin MJ, Ohman DE (1996) Identification of algI and algJ in the Pseudomonas aeru-
ginosa alginate biosynthetic gene cluster which are required for alginate O-acetylation.
J Bacteriol 178(8):2186–2195
118. Franklin MJ, Ohman DE (1993) Identification of algF in the alginate biosynthetic gene
cluster of Pseudomonas aeruginosa which is required for alginate acetylation. J Bacteriol
175(16):5057–5065
119. Wong TY, Preston LA, Schiller NL (2000) Alginate lyase: review of major sources and
enzyme characteristics, structure-function analysis, biological roles, and applications. Annu
Rev Microbiol 54(1):289–340
120. Jain S, Ohman DE (2005) Role of an alginate lyase for alginate transport in mucoid
Pseudomonas aeruginosa. Infect Immun 73(10):6429–6436
121. Wang Y, Moradali MF, Goudarztalejerdi A, Sims IM, Rehm BH (2016) Biological function
of a polysaccharide degrading enzyme in the periplasm. Sci Rep 6
122. Bakkevig K, Sletta H, Gimmestad M, Aune R, Ertesvåg H, Degnes K, Christensen BE,
Ellingsen TE, Valla S (2005) Role of the Pseudomonas fluorescens alginate lyase (AlgL) in
clearing the periplasm of alginates not exported to the extracellular environment. J Bacteriol
187(24):8375–8384
123. Jain S, Ohman DE (1998) Deletion of algK in mucoid Pseudomonas aeruginosa blocks algi-
nate polymer formation and results in uronic acid secretion. J Bacteriol 180(3):634–641
124. Robles-Price A, Wong TY, Sletta H, Valla S, Schiller NL (2004) AlgX is a periplasmic protein
required for alginate biosynthesis in Pseudomonas aeruginosa. J Bacteriol 186(21):7369–7377
125. Rehm B, Boheim G, Tommassen J, Winkler U (1994) Overexpression of algE in Escherichia
coli: subcellular localization, purification, and ion channel properties. J Bacteriol
176(18):5639–5647
1 Alginate Biosynthesis and Biotechnological Production 25
126. Whitney JC, Hay ID, Li C, Eckford PDW, Robinson H, Amaya MF, Wood LF, Ohman DE,
Bear CE, Rehm BH, Lynne Howell P (2011) Structural basis for alginate secretion across the
bacterial outer membrane. Proc Natl Acad Sci U S A 108(32):13083–13088
127. Keiski C-L, Harwich M, Jain S, Neculai AM, Yip P, Robinson H, Whitney JC, Riley L,
Burrows LL, Ohman DE (2010) AlgK is a TPR-containing protein and the periplasmic com-
ponent of a novel exopolysaccharide secretin. Structure 18(2):265–273
128. Remminghorst U, Rehm BH (2006) In vitro alginate polymerization and the functional
role of Alg8 in alginate production by Pseudomonas aeruginosa. Appl Environ Microbiol
72(1):298–305
129. Skjåk-Bræk G, Donati I, Paoletti S (2015) Alginate hydrogels: properties and applications.
In: Matricardi FA P, Coviello T (eds) Polysaccharide hydrogels: characterization and bio-
medical applications. Pan Stanford Publishing Pte Ltd, Singapore
130. Donati I, Paoletti S (2009) Material properties of alginates. In: Alginates: biology and appli-
cations. Springer, Berlin, pp 1–53
131. Conti E, Flaibani A, O’Regan M, Sutherland IW (1994) Alginate from Pseudomonas fluore-
scens and P. putida: production and properties. Microbiology 140(5):1125–1132
132. Gacesa P (1988) Alginates. Carbohydr Polym 8(3):161–182
Chapter 2
Alginate Production from Marine Macroalgae,
with Emphasis on Kelp Farming
César Peteiro
Abstract Alginates are produced industrially from marine macroalgae (also called
seaweeds) belonging to the taxonomic group of brown algae (phylum Ochrophyta,
class Phaeophyceae). In particular, the seaweeds commonly known as kelps (order
Laminariales) are the most widely exploited worldwide as raw materials for alginate
production. Alginophytes (i.e. alginate-yielding seaweeds) are mainly harvested
from wild populations, although some of the raw material that is used in the alginate
industry comes from the cultivation of the kelp Saccharina japonica. The demand
for alginate production has increased over time, and it is likely to increase signifi-
cantly in the future, particularly for the use of alginates in current and future bio-
medical and bioengineering applications. However, alginophyte resources are
limited, and the natural kelp resources have declined worldwide in recent years.
One way to meet the current and future demands of alginate-using industries is to
encourage alginate production via kelp farming. The mariculture of the kelp S.
japonica has already been well developed in Asia, and the cultivation of other kelp
species is currently also being attempted in Europe and the Americas. This chapter
provides an overview of seaweeds as a feedstock for alginate production, with
emphasis on kelp farming to ensure a sustainable supply of alginates required for
many applications. It describes the major stages for the cultivation of Saccharina
and any other kelp, as well as the economic and environmental benefits of integrated
kelp aquaculture to produce alginates, in addition to other value-added products.
C. Peteiro (*)
Seaweed Culture Center, Oceanographic Center of Santander, Spanish Institute of
Oceanography (IEO), Santander, Spain
e-mail: peteiro@st.ieo.es; cpeteiro@gmail.com
2.1 Introduction
Alginate, also called algin, is the generic name for the salts of alginic acid or any of
the derivatives of this compound. It belongs to the family of linear unbranched poly-
saccharides, which consists of binary copolymers of β-D-mannuronic acid (M) and
α-L-guluronic acid (G) units linked together by 1 → 4 glycosidic bonds (see repre-
sentation of the two monomeric units of alginic acid in Fig. 2.1). The monomers are
mainly arranged in sequences of homopolymeric blocks (MM and GG blocks) and
heteropolymeric blocks (MG or GM blocks) [1–3]. The block types and their respec-
tive chair conformations are shown in Fig. 2.2. The monomer sequence distribution
in the copolymer gives rise to a flat ribbonlike structure for the MM blocks, a buck-
led ribbonlike structure for the GG blocks and a helix-like structure for the MG or
GM blocks. The differences in conformation are due to the existence of a linkage in
diequatorial position for the MM blocks, a linkage in diaxial position for the GG
blocks and an equatorial/axial or axial/equatorial linkage for the MG or GM blocks,
respectively. The linkage in the block structure results in varying degrees of stiffness
or flexibility in alginates due to a greater or lesser hindrance of rotation around the
glycosidic bonds. The polymer chains of alginates containing predominantly GG
blocks are stiffer and possess a more extended chain conformation than those con-
taining MM blocks, which in turn are stiffer than MG or GM blocks (i.e. the relative
flexibility increasing in the order G block < M block < MG or GM block) [4–6].
Alginate structure depends fundamentally on the monomer composition, sequen-
tial structure and molecular weight of the polymeric chain. These structural param-
eters affect the chemical and physical properties of alginate, and these properties in
turn have both biological and industrial significance [7–9]. Generally, chemical
structure of alginate is typically described by the frequencies of monads (one mono-
mer unit: M or G), dyads (blocks containing two monomer units: MM, GG, or
MG = GM) and sometimes triads (blocks containing three monomer units: GGG,
Fig. 2.2 Principal block structures in alginate chair conformation: M block, G block and MG or
GM block
MGM, or GGM = MGG) [10–14]. The monad frequencies (FM and FG), the dyad
frequencies (FMM, FGG and FMG = FGM) and the triad frequencies (FGGG, FMGG and
FGGM, = FMGG) are preferably expressed as a mole fraction [15–17], although it has
previously been reported as a percentage [8, 15, 18]. In addition, commercial algi-
nate is traditionally characterized by the ratio of mannuronic to guluronic acid
(M/G), which is also currently estimated from monad frequencies [3, 17, 19]. These
30 C. Peteiro
key structural elements of alginate are obtained by applying various methods (for
more details, see review in ref. [15]). Among all techniques used for the description
of alginates, proton nuclear magnetic resonance (1H–NMR) spectroscopy is the
most accurate method currently employed to determine both the composition and
sequential structure of alginates.
One of the most important and useful properties of alginates is their ability to
form gels and stabilize emulsions in the presence of certain metal cations, particu-
larly divalent cations such as calcium (Ca2+), through a cross-linking reaction [20–
22]. This ability is conventionally described in terms of the so-called “egg-box”
model proposed by Grant and co-workers [20]. According to this model, the diva-
lent cations are embedded into cavities formed naturally by two adjacent polymer
chains containing GG blocks in a helical conformation. The alginate chains thereby
adopt a structure that resembles an “egg-box”, hence the name given to this model.
The mechanism for the alginate gelation may involve the ionic-bonding interaction
of cations with carboxyl groups and the hydrogen-bonding interaction of these
cross-linking agents with oxygen atoms, in both cases between the guluronic acid
blocks of two adjacent polymer chains [5, 21, 22]. The alginate gelation process
with divalent calcium cations is depicted in Fig. 2.3.
Alginate gel formation is mainly dependent on the type and concentration of
cross-linking agents, as well as the composition, sequence and polymer chain length
of the alginate; these features determine the physical properties of the gels formed
[23–25]. For example, the binding affinity of alginates for different divalent cations
has been shown to increase in the following order: barium (Ba2+) > strontium
(Sr2+) > calcium (Ca2+) > magnesium (Mg2+), as well as increasing with the density
of the crosslinkers. In addition, alginates with a high guluronic acid (G) content
display a higher affinity towards these crosslinkers than do alginates with high man-
nuronic acid (M) content [20, 25–27]. Essentially, gel strength and viscosity are the
two most important physical properties used to assess the gelling capability of algi-
nates [23, 28, 29]. While the gel strength is mainly dependent on the content and
length of the guluronic acid (G) in the alginate [26, 28, 30], the viscosity of an
alginate solution is directly determined by the alginate concentration and the chain
length of the alginate polymer, which is proportional to its molecular weight [17,
31, 32]. Generally, alginates rich in guluronic acid are known to form strong but
brittle gels, whereas those rich in mannuronic acid or mixed sequences form weaker
but more flexible gels [27, 33, 34]. Thus, gel strength has also been shown to
increase in the order of GG block > MG block > MM block [26, 27].
The physical and chemical properties vary considerably among different com-
mercial alginates. This natural variability in alginates provides a wide range of func-
tional properties that determine their use in specific applications and thus also their
commercial value [8, 9, 34]. Furthermore, enzymatic and chemical modifications
have been used to manipulate the composition, sequential structure and molecular
weights of alginates, and their derivatives exhibit novel or improved functional
properties for specific high-value applications [35–37].
Alginate was discovered in 1881 by the British pharmacist Stanford [38, 39], and
it has since become one of the most useful and versatile polymers, used in a wide
range of industries. Because of their gelling, thickening, emulsifying and stabilizing
2 Alginate Production from Marine Macroalgae, with Emphasis on Kelp Farming 31
Fig. 2.3 Schematic representation of the egg-box model for calcium alginate gelation. (a)
Illustration of the binding of polymer chains and (b) the formation of junction zones in alginate gels
properties, alginates have been commonly employed in the food, textile printing,
papermaking and pharmaceutical industries, as well as for many other purposes.
Alginates are especially important in the food and beverage industry, in which they
are used as food additives or functional food ingredients in a vast array of different
dairy products [9, 33, 40]. Alginates are internationally accepted food additives and
are therefore explicitly listed as human food ingredients by the European Union
(EU) and as “generally recognized as safe” (GRAS) by the US Food and Drug
Administration (FDA), as well as being recognized as such in the United Nations
32 C. Peteiro
Codex Alimentarius (Latin for “Food Code”) established by the Food and Agriculture
Organization (FAO) and the World Health Organization (WHO). In particular, the
reference codes of the European Union for the different alginates used in the food
industries are E400 (alginic acid), E401 (sodium alginate), E402 (potassium algi-
nate), E403 (ammonium alginate), E404 (calcium alginate) and E405 (propylene
glycol alginate, usually abbreviated as PGA) [15, 40].
More recently, the alginates have found a wide variety of applications in bio-
medical and bioengineering fields. The interest in and use of alginates for biomedi-
cal applications has expanded considerably in recent years because of alginates’
unique and favourable properties such as gelling capacity, biocompatibility, biode-
gradability and lack of toxicity as well as their biological and pharmacological
activities [7, 8, 41]. The current biomedical applications of alginates are the focus
of this book, and an updated and detailed review of the subject may be found in the
different chapters. Although the food and textile uses are still the most important
markets worldwide for alginates, there are growing markets in the bioscience, bio-
engineering and medical fields. The demand for alginate production has increased
during recent years, and it is likely to increase significantly in the future, particu-
larly for their use in current and future biomedical and bioengineering applications
worldwide [42–44].
Alginates occur naturally as a major structural component in marine macroalgae
(also called seaweeds) belonging to the taxonomic group of brown algae (phylum
Ochrophyta, class Phaeophyceae) [42, 44, 45] and are also produced as extracellular
polysaccharides (exopolysaccharides) by some bacteria belonging to the genera
Pseudomonas and Azotobacter [46–48]. Currently, all commercial alginates are
produced solely from brown seaweeds [43, 44] because most species contain large
amounts of alginate [44, 49, 50] and because of the availability of seaweed resources,
as they can be harvested from natural populations and farmed in the sea [42, 51, 52].
The industrially most important seaweeds used worldwide for alginate production
are the species commonly known as kelp (order Laminariales) [42–44]. This review
focuses on the source of seaweed alginate, the process for alginate extraction and
the availability of seaweed resources and their exploitation. It also summarizes the
techniques that have been developed for the commercial-scale farming of kelps as
well as describes the important environmental benefits associated with their
cultivation.
The term algae (singular, alga) is commonly used to refer to a large and diverse
group of aquatic photosynthetic organisms that can grow in marine, brackish and
freshwater environments. Based on morphology and size, algae are generally
grouped into two categories: macroalgae and microalgae. Macroalgae are
2 Alginate Production from Marine Macroalgae, with Emphasis on Kelp Farming 33
multicellular forms, often with plant-like structures, ranging in length from a few
millimetres up to 50 m, which typically live on hard-bottom substrates (i.e. benthic)
of coastal marine habitats. In contrast, microalgae are unicellular or simple forms
with a size range of a few micrometres up to hundreds of millimetres, which typi-
cally grow suspended in water [53, 54]. Marine macroalgae, so-called seaweeds, are
classified primarily on the basis of their photosynthetic pigment composition into
three different phyla (taxonomic groups): Ochrophyta (brown algae), Rhodophyta
(red seaweed) and Chlorophyta (green algae) [55]. For example, the presence of the
pigment fucoxanthin is responsible for the characteristic yellow-brown colour of
brown algae. In addition, these taxonomic groups also differ in many ways, particu-
larly in their morphology, life history, storage compounds and cell wall polysac-
charides [53, 54].
Alginate is characteristically present in most or all species of brown algae, which
belong to the class Phaeophyceae (phylum Ochrophyta, formerly named
Phaeophyta), as a structural component of the matrix of the cell wall and intercel-
lular regions. In these seaweeds, alginate is found in the form of insoluble mixed
salts of alginic acid, mainly with calcium and to a lesser extent with sodium, potas-
sium, and magnesium, strontium and barium, among other ions naturally found in
seawater [1, 56, 57]. Its biological function is primarily skeletal, giving the algae
both the mechanical strength and the flexibility necessary to withstand the force of
the sea. Indeed, functional differences in the alginate content and structure of sea-
weeds have been reported. For example, the seaweeds growing in more wave-
exposed habitats have alginates with higher mannuronic acid content than those in
wave-sheltered habitats, providing greater flexibility to withstand the wave action
[58–60]. It has also been observed that the part of the thallus (plural, thalli) that
attaches the algae to a hard substrate (the so-called holdfast) contained more gulu-
ronic acid than in the rest of the thallus, giving it more rigidity and thereby affixing
it more firmly to the rock [59–61]. In addition, alginate plays important roles in high
ion-exchange equilibrium with seawater as well as functioning in retarding desicca-
tion when the seaweeds are exposed to the air during low tide [56, 57, 62].
Brown seaweeds of the class Phaeophyceae (Ochrophyta) and in particular some
species of the orders Laminariales and Fucales (commonly known as kelps and
fucoids, respectively) have large amounts of alginate, comprising up to 55% of their
dry weight (Table 2.1). Both kelps and fucoids are the largest and most structurally
complex brown seaweeds. Generally, and in particular in kelps, the body or thallus
of the macroalgae consists of a holdfast (root-like), stipe (stem-like) and blade
(leaf-like) (see Fig. 2.4, in which some kelp species are illustrated). At present, com-
mercial alginates are produced mainly from brown seaweeds of the genera
Laminaria, Saccharina, Lessonia, Macrocystis, Durvillaea, Ecklonia and
Ascophyllum [42, 43] (Fig. 2.4). Specifically, the industrially most important algi-
nate-yielding species (alginophytes) are currently the kelps (Laminariales)
Macrocystis pyrifera, Laminaria hyperborea, Laminaria digitata, Saccharina
japonica, Lessonia nigrescens species complex, Lessonia trabeculata, Ecklonia
arborea and Ecklonia radiata as well as the fucoids (Fucales) Durvillaea potatorum
and Ascophyllum nodosum [42, 43] (more information on these alginophyte
resources will be described in Sect. 2.2.3).
34 C. Peteiro
Table 2.1 Alginate yields from the brown seaweeds used for industrial production
Alginate content
(DW) Country of
Seaweed species Range Mean origin Sampling month References
Macrocystis pyrifera 26–37% n.d. Mexico Feb.–Nov. [63]
(monthly)
18–45% n.d. Chile Year-round [50]
Laminaria digitata 18–26% n.d. United Year-round [60]
Kingdom
16–36% n.d. Denmark Year-round [64]
Laminaria 14–21% n.d. United Year-round [60]
hyperborea Kingdom
Saccharina japonica 15–20% n.d. China Mar., Apr., May [65]
17–25% n.d. Japan Mar.–Oct. [66]
(monthly)
Saccharina latissima 16–34% n.d. Denmark Year-round [64]
Lessonia trabeculata 13–29% n.d. Chile Jul. [67]
Ecklonia arborea 24–28% n.d. Mexico Feb., May., Aug., [63]
Nov.
Durvillaea potatorum n.d. 55% Australia Mar. [68]
n.d. 45% New Zealand n.d. [69]
Ascophyllum 12–16% n.d. Russia Apr., Aug., Dec. [70]
nodosum
Alginate yield based on the dry seaweed weight (DW). Data obtained from the seaweed thallus and
using similar alginate extraction methods
n.d no available data
Tables 2.1 and 2.2 present the alginate yield and chemical composition of the
main kelp and fucoid species used worldwide for alginate production. These param-
eters vary considerably between and within brown seaweed species. Based on these
data, the highest levels of alginate are found in Durvillaea potatorum and
Macrocystis pyrifera, in which alginate represents up to 55% and 45% of the dry
seaweed weight, respectively, while the lowest levels are in Ascophyllum nodosum
and Laminaria hyperborea, in which alginate represents 12% and 13% of the dry
weight. Regarding structural parameters, Lessonia trabeculata and Laminaria
hyperborea have the highest fraction of guluronic acid (M/G ratio of <1), while the
lowest proportions of guluronic acid are observed in Durvillaea potatorum and D.
antarctica and to a lesser extent in Saccharina japonica, Macrocystis pyrifera and
Ascophyllum nodosum (all with M/G ratios of >1.5). Brown seaweeds are well
known to exhibit some seasonal variation both in alginate chemical structure and
alginate content, which can be higher or lower depending on the species [49, 50,
75]. For example, the content of alginate in Macrocystis pyrifera is highly variable
throughout the year, in contrast to that in Laminaria digitata, which is much more
stable (Fig. 2.5). Similarly, there may also be seasonal differences in the structural
characteristics of alginate, as in the case of Saccharina latissima, whose M/G ratio
varies seasonally. However, there is hardly any variation in the M/G ratio over the
2 Alginate Production from Marine Macroalgae, with Emphasis on Kelp Farming
Fig. 2.4 The industrially most important alginate-yielding seaweeds worldwide (Data from Refs. [42, 43]. Taxonomic classification based on Algaebase [55].
35
The drawings are not completely to scale between the different seaweeds in order to improve the view. Note that, when referring collectively to some or all of
the species in a genus, the generic name is followed by spp)
36 C. Peteiro
Table 2.2 Composition and sequence of alginates obtained from various brown seaweeds
Composition Sequence Country
Seaweed species FM FG M/G FMM FGG FMG, GM of origin References
Macrocystis pyrifera 0.62 0.38 1.63 0.42 0.18 0.20 Mexico [10]
Laminaria digitata 0.59 0.41 1.43 0.43 0.25 0.16 Norway [17]
Laminaria 0.45 0.55 0.81 0.28 0.38 0.17 n.d. [17]
hyperborea
Laminaria 0.32 0.68 0.47 0.20 0.56 0.12 n.d. [17]
hyperborea (stipe)
Saccharina japonica 0.65 0.35 1.85 0.48 0.18 0.17 China [71]
Saccharina latissima 0.45 0.55 0.82 0.33 0.43 0.12 Norway [72]
Saccharina 0.41 0.59 0.69 0.07 0.25 0.34 Canada, [16]
longicruris May
Lessonia nigrescens 0.59 0.41 1.43 0.40 0.22 0.19 n.d. [13]
species complex
Lessonia trabeculata 0.38 0.62 0.61 0.21 0.47 0.15 Chile [58]
Lessonia trabeculata 0.22 0.78 0.28 0.10 0.67 0.11 Chile [73]
(stipe)
Ecklonia arborea 0.52 0.48 1.08 0.37 0.33 0.15 Mexico [10]
Ecklonia maxima 0.55 0.45 1.22 0.32 0.22 0.32 n.d. [15]
Durvillaea potatorum 0.76 0.24 3.17 0.58 0.06 0.18 New [69]
Zealand
Durvillaea antarctica 0.68 0.32 2.15 0.51 0.16 0.17 n.d. [13]
Ascophyllum nodosum 0.61 0.39 1.56 0.46 0.23 0.16 n.d. [74]
Structural parameters determined by proton nuclear magnetic resonance (1H–NMR) spectroscopy.
Data were obtained from the same tissue type, the thallus (except where specified), and using
similar alginate extraction methods
course of a year in species such as Laminaria digitata (Fig. 2.6). The seasonal vari-
ability of alginate is related mainly to seasonal changes in temperature as well as to
nutrient and light availability, most often influencing seaweed growth [76–78]. It
has been reported that the highest values of alginate in some kelps and fucoids occur
in the summer months [49, 50, 70]. However, in general, it appears that there is no
overall pattern of seasonal variation, and the same applies to the composition of
alginates. Thus, it is essential to know the seasonal composition of alginates in
brown seaweeds in order to determine the optimal harvesting time by which to
obtain not only higher quantities but, above all, alginates of better quality, i.e. those
with high G content.
Nevertheless, alginate content and structure also depend on the age and part of
the seaweed used [50, 59, 61] as well as on the environmental conditions of the
habitat in which the seaweed grew [58, 59, 64]. In general terms, thalli from older
seaweeds are richer in mannuronic acid than those from younger specimens [17,
33]. In addition, compared with the blades, the stipes of kelp species generally con-
tain a higher amount of alginate rich in guluronic acid [62, 75, 79]. Indeed, the
highest content of guluronic acid in alginate is obtained from stipes of the kelps
2 Alginate Production from Marine Macroalgae, with Emphasis on Kelp Farming 37
Fig. 2.5 Seasonal variation in alginate content of the brown seaweeds Macrocystis pyrifera (MP),
Saccharina latissima (SL) and Laminaria hyperborea (LH) (Data from Refs. [49, 50])
Fig. 2.6 Seasonal variation in mannuronic to guluronic acid ratios (M/G) of alginate from the
brown seaweeds Saccharina latissima (SL) and Laminaria digitata (LD) (Data from Ref. [64])
[80]. All these derivatives of alginic acid are industrially produced following the
same manufacturing process that will be described here based on sodium alginate
extraction. The process of alginate extraction from seaweeds is based on the conver-
sion in an alkaline medium of the water-insoluble mixed salts of alginic acid from
algal cell wall matrix to water-soluble salts, normally sodium alginate, followed by
precipitation and purification [42, 81, 82]. This conventional procedure for alginate
extraction has been well studied during the last decade to optimize the yield and
quality of alginate for various applications [81, 83, 84]. Today, it is commonly used
in the industry to produce alginates from brown seaweeds [42, 80].
The following will present in detail each of the steps constituting the process of
producing seaweed alginate. The alginate extraction procedure is schematically
illustrated in Fig. 2.7. Generally, it consists of three major steps: (1) pre-extraction,
(2) neutralization and (3) precipitation/purification [42, 81, 82].
In the first step, the seaweeds (usually dried) are washed with distilled water and
then ground to speed up the chemical reactions for the extraction of alginic acid.
Further, 0.1% formaldehyde solution may be added in order to avoid pigments in
alginate; this has been seen to increase the alginate yield. The milled algal biomass
is then dissolved and stirred with a dilute mineral acid up to pH 4 (usually hydro-
chloric acid (HCl) or calcium chloride (CaCl2) at 0.1–0.2 M) to remove counter ions
(Ca2+, Na+, Mg2+, Sr2+, etc.) of algal alginate by ion exchange with protons from the
acid [81, 85]. The acid treatment is also effective in removing potential contami-
nants or impurities (fucoidans, laminarins, proteins and polyphenols), leading to a
higher final yield and purity of alginate. In addition, 85% ethanol can be used to
extract pigments and proteins in this process [16]. This pretreatment is often
repeated several times to ensure full extraction of alginic acid. At the end of this
process, the supernatant (residual algal particles) is eliminated [81, 85].
2 Alginate Production from Marine Macroalgae, with Emphasis on Kelp Farming 39
Fig. 2.7 Schematic flow chart of the process of extracting sodium alginate and other forms of
alginates from seaweeds
40 C. Peteiro
In the second step, the insoluble alginic acid in the seaweed-water mixture is
brought to pH 9–10 with an alkaline solution (usually sodium carbonate (Na2CO3)
or sodium hydroxide (NaOH)) to form water-soluble sodium alginate. In this pro-
cess, the mixture is mechanically stirred, and the temperature is maintained at
60–80 °C. The insoluble algal residues are removed by extensive centrifugation and
subsequent filtration (up to 0.2 μm pore size), thereby obtaining the sodium alginate
in aqueous solution [16, 81].
In the third step, the sodium alginate solution can be precipitated into sodium
alginate, calcium alginate or alginic acid by the addition, respectively, of alcohol
(usually ethanol (C2H6O)), calcium chloride (CaCl2) and hydrochloric acid (HCl).
These three methods are therefore known as the sodium alginate process or ethanol
route, the calcium alginate process or CaCl2 route and the alginic acid process or
HCl route [42, 81, 85]. The sodium alginate precipitated via the ethanol route is
obtained directly by the addition of ethanol, and it is separated by solvent extrac-
tion/evaporation. The CaCl2 route first produces a precipitated calcium alginate that
is isolated by sieving and rinsed with distilled water to remove the excess calcium.
It is then converted to alginic acid by acid treatment, generally using hydrohydro-
chloric acid (HCl) as described above in the first step. The HCl route directly yields
alginic acid, which is separated from the solution by simple flotation and centrifuga-
tion. The resulting alginic acid can also be reconverted by alkaline neutralization to
any of the commercial forms of alginate in the same manner as described in the
second step. Specifically, sodium carbonate (Na2CO3), potassium carbonate
(K2CO3), ammonium carbonate ((NH4)2CO3), magnesium carbonate (MgCO3), cal-
cium carbonate (CaCO3) or propylene oxide (C3H6O) is added in order to obtain the
following alginates, respectively: sodium alginate (Na-alginate), potassium alginate
(K-alginate), ammonium alginate (NH4-alginate), magnesium alginate
(Mg-alginate), calcium alginate (Ca-alginate) and propylene glycol alginate (PGA).
Finally, all alginates form a paste that is separated, dried and milled. Commercial
alginates produced specifically for biomedical purposes (e.g. ultrapure and amito-
genic alginates) are prepared using more rigorous extraction processes to remove
any biological and inorganic impurities, and companies consider these processes
confidential. However, the alginates obtained from the described methods usually
contain some impurities, making them unsuitable for some biomedical applications.
In this case, an alternative extraction method using barium ions (Ba2+) is used in the
precipitation process due to their high binding affinity and selectivity towards algi-
nates. Subsequently, the Ba2+ from the alginate is exchanged for sodium ions to
form sodium alginate, which can be precipitated using ethanol [36].
It is known that the alginate extraction process can influence the yield and chemi-
cal compositions as well as rheological properties of the isolated alginates [42, 81,
85]. To illustrate these effects, Table 2.3 summarizes the comparative results of
three precipitation methods in the process of alginate extraction from the brown
seaweed Macrocystis pyrifera. According to these data, the sodium alginate process
or ethanol route gives the highest yield and rheological properties of alginates,
although very similar results can be obtained from the alginic acid process or HCl
route. Clearly, the calcium alginate process or CaCl2 route results in an alginate with
poor viscoelastic properties and low toughness [85].
2 Alginate Production from Marine Macroalgae, with Emphasis on Kelp Farming 41
Table 2.3 Yields and properties of alginate obtained from the seaweed Macrocystis pyrifera using
three precipitation methods in the alginate extraction process
Precipitation Extraction yield Composition Sequence Molar mass Viscosity
methods (% DW) M/G FGG Mw (kg/mol) [ŋ] (mL/g)
CaCl2 route 28 1.045 0.730 75 160
HCl route 27 1.295 0.675 220 505
Ethanol route 33 1.170 0.690 297 575
Data from Ref. [85]. Alginate yield based on the dry weight of the seaweed samples (DT). If there
are several values from different samples, the average is shown
Overall, it is also important to control both pH and temperature during all extrac-
tion processes to improve the yield and quality of the alginate obtained. For exam-
ple, it is well known that acid treatment at a pH lower than 4 may break bonds of the
polymer chain, decreasing the alginate viscosity [81, 84, 85]. In addition, it has been
shown that high temperatures (higher than 80 °C) and longer extraction processes
increase yield but decrease viscosity [81, 84].
Brown seaweeds have been worldwide resources for alginate extraction since indus-
trial alginate production began in 1929 in California, USA, shortly thereafter, begin-
ning in 1939 in several European countries and Japan and, more recently in the
1980s, in China [42, 43, 81]. Currently, the alginate industry is concentrated into 15
factories in 6 different countries (China, the USA, the United Kingdom, Japan,
Chile and Germany), and most of these factories are now in China (see compilation
in ref. [86] for more details on commercial companies selling alginates). According
to the latest available data for 2009, the world market for alginates is approximately
26,500 tons (dry weight), with an estimated sale value of US$ 318 million annually
[43]. Alginate production worldwide is derived from seaweed resources, most of
which are harvested from the wild, except in China, where seaweed is sourced
mostly from aquaculture [42, 43, 81].
Although there are over 2000 species of brown seaweeds (phylum Ochrophyta,
class Phaeophyceae) [55], only some species of the orders Laminariales and Fucales
(kelp and fucoid) are exploited worldwide as raw material for alginate production.
It is estimated that there currently are at least 38 species of kelps and fucoids grown
in 24 countries that are used worldwide to produce alginates [51, 52, 87]. A list of
alginate-yielding seaweeds (or alginophytes) and their countries of origin is pro-
vided in Table 2.4, which includes the full names of the species and their taxonomic
classifications (updated). However, many of these seaweeds are harvested from
small local stocks and are therefore not available on the international market. In
addition, some seaweed species (e.g. Sargassum species) are only used occasionally
for alginate production, when the main commercial sources are unavailable, because
their alginate is usually judged to be of “borderline” quality [42, 43]. Thus, the
42 C. Peteiro
Table 2.4 Species of brown seaweeds used in various countries for alginate production
Seaweed species (arranged by order and family) Country
Order: Laminariales
Family: Laminariaceae
Macrocystis pyrifera (Linnaeus) C.Agardh 1820 PE, CL, MX, US, CA,
NZ
Laminaria digitata (Hudson) J.V.Lamouroux 1813 FR, IS, DK
Laminaria hyperborea (Gunnerus) Foslie 1884 ES, FR, IE, NO, RU
Laminaria longipes Bory 1826 RU
Laminaria ochroleuca Bachelot de la Pylaie 1824 ES
Saccharina angustata (Kjellman) C.E.Lane, C.Mayes, Druehl & JP, RU
G.W.Saunders 2006
Saccharina bongardiana (Postels & Ruprecht) Selivanova, Zhigadlova RU
& G.I.Hansen 2007
Saccharina cichorioides (Miyabe) C.E.Lane, C.Mayes, Druehl & RU
G.W.Saunders 2006
Saccharina gurjanovae (A.D.Zinova) Selivanova, Zhigadlova & RU
G.I.Hansen 2007
Saccharina japonica (Areschoug) C.E.Lane, C.Mayes, Druehl & CN, KR, JP, RU
G.W.Saunders 2006
Saccharina latissima (Linnaeus) C.E.Lane, C.Mayes, Druehl & ES, FR, RU, CA
G.W.Saunders
Family: Lessoniaceae
Lessonia nigrescens Bory 1826 CL, PE
Lessonia trabeculata Villouta & Santelices 1986 CL
Lessonia berteroana Montagne 1842 CL, PE
Lessonia spicata (Suhr) Santelices 2012 CL, PE
Ecklonia arborea (Areschoug) M.D.Rothman, Mattio & J.J.Bolton MX
2015
Ecklonia maxima (Osbeck) Papenfuss 1940 ZA
Ecklonia radiata (C.Agardh) J.Agardh 1848 AU, NZ
Order: Fucales
Family: Durvillaeaceae
Durvillaea potatorum (Labillardière) Areschoug 1854 AU
Durvillaea antarctica (Chamisso) Hariot 1892 CL
Family: Fucaceae
Ascophyllum nodosum (Linnaeus) Le Jolis 1863 FR, IE, IS, NO, US,
CA
Fucus serratus Linnaeus 1753 IE
Fucus vesiculosus Linnaeus 1753 IE
Family: Sargassaceae
Sargassum swartzii C.Agardh Saunders 1820 IN
Sargassum aquifolium (Turner) C.Agardh 1820 PH
Sargassum cinctum J.Agardh 1848 PH
Sargassum hemiphyllum (Turner) C.Agardh 1820 PH
(continued)
2 Alginate Production from Marine Macroalgae, with Emphasis on Kelp Farming 43
Table 2.4 (continued)
Seaweed species (arranged by order and family) Country
Sargassum ilicifolium (Turner) C.Agardh 1820 PH
Sargassum feldmannii Pham-Hoàng Hô PH
Sargassum paniculatum J.Agardh 1848 PH
Sargassum polycystum C.Agardh 1824 CN, PH, ID, TH
Sargassum siliquosum J.Agardh 1848 PH, VN
Sargassum graminifolium C.Agardh 1820 VN
Sargassum henslowianum C.Agardh 1848 VN
Sargassum mcclurei Setchell 1933 VN
Family: Sargassaceae
Turbinaria conoides (J.Agardh) Kützing 1860 IN
Turbinaria decurrens Bory 1828 IN
Turbinaria ornata (Turner) J.Agardh 1848 IN
Data from Refs. [42, 51, 52, 80, 87, 88]. Full names of species and their taxonomic classifications
were validated with AlgaeBase [55] and updated when necessary. Country abbreviations: Australia
(AU), Canada (CA), Chile (CL), China (CN), Denmark (DK), Spain (ES), France (FR), India (IN),
Indonesia (ID), Ireland (IE), Iceland (IS), Japan (JP), Korea (KR), Mexico (MX), NO (Norway),
New Zealand (NZ), Peru (PE), Philippines (PH), Russia (RS), Thailand (TH), United Kingdom
(UK), United States of America (US), Vietnam (VN) and South Africa (ZA)
global alginate production worldwide comes from a small number of seaweed spe-
cies, specifically the kelps Macrocystis pyrifera, Laminaria hyperborea, L. digitata,
Saccharina japonica, Lessonia nigrescens species complex, L. trabeculata, Ecklonia
arborea, and Ecklonia radiata and the fucoids Durvillaea potatorum and
Ascophyllum nodosum [42, 43, 80]. These alginophytes, in addition to containing
large amounts of alginate [43, 44, 51], form dense stands on shallow rocky shores
commonly referred to as forests or beds. These seaweed species are distributed in
cold-temperate waters around the world, and as photosynthetic organisms, they are
restricted to habitats with appropriate light levels, primarily from the intertidal zone
to a depth of 50 m in the sublittoral zone.
The quantities of alginophytes harvested worldwide in 2009 are summarized in
Table 2.5, including the main producer countries and the quality or type of alginate
obtained from seaweed species. These statistics are based on the most up-to-date
and reliable estimates available [43], but some species names have been updated to
the current taxonomy [55]. Indeed, recent revision of kelps based on the application
of molecular techniques has greatly improved the taxonomic understanding of this
group, resulting changes in the taxonomic identity and nomenclature of some com-
mercialized species. For example, some species of the former Laminaria sensu lato
have been transferred to the new genus Saccharina, including the cultivated
Laminaria japonica (now Saccharina japonica) [89]. Lessonia species have also
undergone taxonomic changes, and the former Lessonia nigrescens is now a species
complex, i.e. encompassing multiple species that were hidden under a single name.
Currently, there is genetic evidence of the presence of at least two cryptic species:
Lessonia berteroana and Lessonia spicata [90, 91]. Moreover, Lessonia flavicans,
44 C. Peteiro
Table 2.5 Alginate production from brown seaweeds (alginophytes) in the world
Alginate Producer Harvested Contribution to
Major alginophytes (species) typea countries (tonnes DW) total (%)
Macrocystis: M. pyrifera Low G US, MX, 5000 5
CL
Laminaria: L. digitata and L. Med/high G FR, IE, 30,500 32
hyperborea UK, NO
Saccharina: S. japonica Med G CN, JP 20,000 21
Lessonia: L. nigrescens Med/high G CL, PE 31,000 33
species complex and L.
trabeculata
Ecklonia: E. maxima Med G ZA 2000 2
Durvillaea: D. potatorum High G AU 4500 5
Ascophyllum: A. nodosum Low G FR, IS, IE, 2000 2
UK
World total production 95,000 100
Quantities harvested of seaweeds are expressed on the basis of their dry weight (DW). Data yield
for the year 2009 from Ref. [43] and commercial species used for alginate production from Refs.
[42, 43, 80]. Seaweed groups and species names were updated to current taxonomy
a
Type of alginate by measuring the guluronic acid content: low, medium and high. For country
abbreviations, see Table 2.4
which has been reported traditionally in the alginate marketplace [43], actually cor-
responds to either Lessonia nigrescens or L. trabeculata [88, 92]. Finally, all species
of the genus Macrocystis are now considered taxonomic synonyms of Macrocystis
pyrifera [93, 94]. Independent of the current taxonomic status of seaweeds, the
commercial companies selling seaweed or algal products generally continue to
maintain the commercial names of their raw materials or products [92]. However, an
adequate specific identification of seaweed species used as alginate sources is
particularly important because of effects on chemical compositions and properties
of the isolated alginate (see Sect. 2.2.1).
Based on data for 2009 from Table 2.5, the world harvest of alginophytes is esti-
mated to be approximately 95,000 tonnes (dry seaweed weight) annually. The main
alginophytes harvested worldwide are Lessonia and Laminaria, accounting for 65%
of the total production, followed by Saccharina with 21% of the total. It is worth
highlighting the drastic reduction in the harvest of Macrocystis and Ascophyllum,
which were previously important raw materials for alginate production.
In the year 2009, these seaweeds supplied only 2% of the worldwide production,
down from 58% in 1999 [43]. The reason for this change is the marked decrease in
the use of these species because their alginate has low guluronic acid (G) content,
while the current market is demanding alginates with intermediate or high G content
[43, 81]. These commercial-grade alginates are now mainly used in biomedical
applications and novel therapies [7, 36, 37], which have increased considerably in
the last decade and are currently the most profitable, selling at the high end of the
alginate market [43]. Regarding the kelp Macrocystis, harvesting has also been cur-
tailed for ecological reasons due to concern about potential environmental effects of
exploitation of kelp forests that provide habitat for many species [43].
2 Alginate Production from Marine Macroalgae, with Emphasis on Kelp Farming 45
Fig. 2.8 Worldwide distribution of seaweed resources harvested industrially for alginate produc-
tion (Data from Refs. [43, 81])
Fig. 2.9 Vessel for harvesting the kelp Macrocystis pyrifera for alginate production along the
coast of California (Photo by C. Peteiro)
is performed using trawler boats with a cutting dredge that is towed through the kelp
beds to cut the seaweeds. In France, Laminaria digitata is harvested by boats with
a hydraulic arm fitted with an iron hook on the end (called a “scoubidou”) that
rotates to wrap the kelp around itself [42, 99].
It has been demonstrated that kelp harvesting in various parts of the world may
cause deterioration of natural resources or habitats or disturbance of species [100–
102]. Kelps act as ecosystem engineers or foundation species, providing habitat,
protection and food for numerous organisms in coastal ecosystems, in the same way
as terrestrial forests [103, 104]. Over the last several years, there has been an
increase in governmental control over the exploitation of natural seaweeds popula-
tions to limit their misuse. Some countries, depending on the state of the specific
resource, allocate harvest quotas and/or establish different management and control
measures to ensure the conservation of kelp forests and to lower the impact of their
exploitation on marine ecosystems. Such measures may include establishing fallow
periods of several years for areas subject to harvest, allowing only the collection of
the upper part of the thallus from perennial seaweeds, limiting harvesting during
non-reproductive periods, or even prohibiting the exploitation of endangered spe-
cies and/or those with high ecological value [102, 105, 106]. In addition, kelp for-
ests are also exposed to a range of disturbances of natural and/or anthropogenic
origins. Particularly in recent years, kelp populations have declined in many areas
of the world due to environmental stress caused by climate change, among other
factors, especially by the increase in sea temperature and disruption in the natural
nutrient availability patterns [107, 108].
Seaweed exploitation in Asia intended for alginate production mainly comes
from commercial cultivation of the kelp Saccharina japonica (kombu) in China
[42–44]. Kelp cultivation techniques have been well developed in Japan and China,
where several Asian species have been cultivated on a large scale since the 1960s
2 Alginate Production from Marine Macroalgae, with Emphasis on Kelp Farming 47
[109, 110]. Today, aquaculture in Asia provides almost all of the global production
of S. japonica, reaching over 900 thousand tonnes dry weight of seaweeds (esti-
mated from fresh weight data) in 2014 [111], of which until now only a small part
has been used for alginate extraction, but, as we have seen (Table 2.5), accounting
for more than 20% of world alginate production [43]. Current practices of kelp
farming in Asia have contributed not only to significantly increasing production to
meet commercial demands for various uses, including alginate production but also
to conserving natural populations and the ecosystems that they produce. Recently,
mariculture of kelp species has also generated great interest in Europe and the
Americas, as it may lead to increased production for commercial uses and potential
applications; in addition, it may help protect the kelp forests from overharvesting
[112–114]. The current limitations on the availability and use of natural kelp
resources are expected to become the major driving force for the growth and devel-
opment of farming of kelp species for alginate production, as has already been the
case in Asia. In fact, it is now widely recognized that a transition from seaweed
extraction to aquaculture is needed to meet the growing demand and to avoid the
decline or loss of natural populations [115, 116]. At present, cultivation of kelp spe-
cies is currently also being attempted in several countries of Europe and the
Americas [112–114]. The techniques and biological basis required for full-cycle
cultivation of Saccharina, as well as for other kelp species, will be described in the
next section.
Kelps are characterized by a heteromorphic life cycle that alternates between a hap-
loid generation formed by microscopic filaments (known as the gametophyte
because it produces gametes) and a diploid generation formed by a macroscopic
thallus (called the sporophyte because it produces spores) [117]; the different life-
history stages are shown in Fig. 2.10. Most kelp species have a perennial (i.e. lasting
several years) sporophytic phase, during which the sporophyte may reach a length
of several metres depending on the specific seaweed taxon (e.g. up to 50 m in
Macrocystis, up to 15 m in Ecklonia, up to 10 m in Durvillaea, up to 4 m in Lessonia,
up to 5 m in Saccharina japonica, up to 2 m in Laminaria hyperborea and up to 1 m
in Laminaria digitata [55]).
Large sporophytes are commercially exploited for different uses, such as the
extraction of alginates [42, 43, 99]. Sporophyte morphology of kelps varies depend-
ing on the species and the environmental conditions in which they grow. However,
three parts are typically recognized: the holdfast, stipe and blade (described in Sect.
2.2.1; see Figs. 2.4 and 2.10) [55, 118]. The gametophytic phase, in contrast, con-
sists of slightly branched male and female filaments composed of round-shaped
48 C. Peteiro
Fig. 2.10 Saccharina life-history stages, which are the same for all brown seaweeds of the order
Laminariales (kelp)
2 Alginate Production from Marine Macroalgae, with Emphasis on Kelp Farming 49
cells smaller than 50 μm in diameter. Microscopic gametophytes are a survival strat-
egy for the sporophyte, enabling long-term resistance to adverse environmental
conditions while it waits to reproduce and form new sporophytes. The filaments of
gametophytes may remain dormant or grow vegetatively, although their growth is
generally much reduced [119–121].
Most kelps are distributed along the rocky shores of the Arctic and the cold-
temperate regions of the Northern and Southern Hemispheres, where temperatures
are generally below 20 °C. Temperature is therefore a key environmental factor that
affects not only the distribution of kelp species but also their growth [122–124]. The
perennial sporophytes of kelps generally exhibit strong seasonality in their develop-
ment with a period of rapid growth during winter and spring and a period of mini-
mal growth during the summer and fall, which coincides with the seasonal
temperature and nutrient cycles in cold-temperate waters. In winter, temperatures
are lower, and nitrogen levels are higher. In contrast, temperatures are higher and
nitrogen levels are often negligible during summer [118, 125]. Another important
environmental factor influencing the development of kelp sporophytes is water
movement, which affects nutrient assimilation and gas exchange by determining
passive transport across the diffusion boundary layer of the algal surface [126].
Cultivation practices are rooted in Asia in the eighteenth century, during which dif-
ferent methods were used to expand the populations of edible kelps as a natural
resource [127, 128]. However, these practices depended entirely on the natural envi-
ronment since there was no control over the biological cycle of these seaweeds. The
scientific basis for the development of the full-cycle cultivation of Saccharina and
other kelp species was first established in the middle of the twentieth century, and
since then, techniques have been established in Asia to obtain seedlings from spores
under more-or-less controlled laboratory conditions. This Asian technique of seed-
ling production enabled the subsequent development of different types of floating
rafts for cultivating kelp in the sea, and beginning in the 1960s, commercial kelp
mariculture extended to different regions of Japan, China and Korea, where it was
promoted by different governments to meet the demand for human consumption in
a context of insufficient natural resources [110, 127–129]. Currently, Saccharina
japonica is the most extensively cultivated kelp species in these countries.
In Europe and the Americas, different cultivation practices were initiated in the
1980s and 1990s to study the viability of native kelps [120, 130–133]. As an alterna-
tive to the Asian method of seedling production, a European technique was devel-
oped to produce kelp seedlings from gametophyte cultures [120, 134]. Research
showed that kelp cultivation using simple, relatively low-cost techniques was bio-
logically and technically feasible, but these early attempts at cultivation did not
continue due to a lack of interest since the available wild kelp stocks were sufficient
50 C. Peteiro
to meet commercial demand and their exploitation was considered more profitable
than cultivation. However, there is currently a growing interest in the development
and optimization of kelp species cultivation in several European and American
countries; in fact, early cultivation practices have already begun on a commercial
scale. This change is due to the growing demand for these species for different high-
value commercial uses as well as the important environmental benefits that their
cultivation would provide [112, 113]. Marine macroalgae use carbon dioxide and
nutrients to grow and may thus contribute to the reduction of atmospheric CO2,
which is a contributing factor to climate change [135–138], and of the amount of
inorganic waste that is discharged into marine coastal areas [139–142]. In particular,
seaweed farming is considered to be the basis for the development of sustainable
aquaculture because they can absorb some of the inorganic nutrients that are pro-
duced, for example, in the aquaculture of mussels and fish [114, 143–148]
(Fig. 2.11). To date, sea farming has been tested with success for the following kelp
species: Macrocystis pyrifera in Chile [149]; Laminaria digitata in Ireland [150];
Saccharina latissima in several countries in Europe [113], Canada [142] and the
USA [139]; Saccharina longicruris in Canada [151]; Lessonia trabeculata in Chile
[152]; Alaria esculenta in Canada [142] and in Ireland [153]; and finally Undaria
pinnatifida in Japan, China and Korea [110] and in France and Spain [154]. In gen-
eral, the techniques developed for the commercial-scale farming of the kelp
Saccharina in Asia [109] have been adapted to the cultivation of other kelp species
in Europe and the Americas [113, 149].
As with other kelps, full-cycle cultivation of Saccharina consists of two very differ-
ent phases associated with their characteristic life cycle (Fig. 2.10). In the first step
(laboratory-culture stage), kelp seedlings are produced on strings (commonly
Fig. 2.11 A bay along the European coast where kelp cultivation (1) occurs along with the cultiva-
tion of mussels (2) and fish (3) (Photo by C. Peteiro)
2 Alginate Production from Marine Macroalgae, with Emphasis on Kelp Farming 51
Fig. 2.12 Diagram summarizing the steps in kelp farming, which include the production of seed-
lings on strings (laboratory-culture stage) and their subsequent attachment to culture ropes for
growth in a floating raft culture (sea-culture stage) (Adapted and reprinted from Ref. [113],
Copyright 2016, with permission from Elsevier)
52 C. Peteiro
gametophytes through vegetative growth that can produce seedlings at any time of
the year [119]. This method is currently being successfully used to cultivate differ-
ent kelp species in Europe [119, 150, 153], the Americas [151, 155] and in Asia,
albeit in a more limited way [156, 157]. Given its extensive role in cultivation, the
production of kelp seedlings from gametophyte cultures is specifically described
here.
The production of seedling strings under laboratory conditions is divided into
two phases: a first phase to create a culture of free-living gametophytes maintained
with aeration under controlled environmental conditions (Fig. 2.13) and a second
phase where gametophytes are sown on strings and grown in tanks, in which game-
togenesis is induced so that, after sexual reproduction, seedlings develop (Fig. 2.14).
Gametophyte cultures are obtained from the germination of spores extracted
from the fertile parts (i.e. reproductive structures) of mature sporophytes that are
generally obtained from natural populations or cultures. The formation of reproduc-
Fig. 2.13 The environmental culture chambers or incubators (above) used to maintain culture
flasks (bottom left) containing microscopic kelp gametophytes (bottom right), growing under free-
living conditions (Photos by C. Peteiro)
2 Alginate Production from Marine Macroalgae, with Emphasis on Kelp Farming 53
tive structures (called sori or sporophylls) in kelp sporophytes can also be induced
in some species under short-day or long-day photoperiods [158, 159]. Spore germi-
nation and the subsequent development of gametophyte cultures are performed in
culture flasks containing sterile, nutrient-enriched seawater under environmental
conditions specific to each kelp species [119, 149, 150, 156]. The entire process is
carried out in environmental culture chambers or incubators that are designed to
rigorously control temperature and light (considering irradiance, the light spectrum
and photoperiod) and to aerate the cultures (Fig. 2.13).
Adequate aeration of free-living gametophytes is maintained by bubbling within
the culture flasks (Fig. 2.13), which homogenizes light and promotes nutrient avail-
ability in the culture medium. Gametophytes are usually conserved in their dormant
or slow-growth states, although vegetative growth can be promoted by filament
fragmentation under specific environmental conditions to increase their biomass.
However, the growth rate of kelp gametophytes is generally very low, so large quan-
tities of spores are usually obtained and germinated to have sufficient reserves of
gametophytes [119]. It is important to note that the gametophyte culture also acts as
a germplasm bank for ex situ kelp conservation, as it can be indefinitely maintained
in vivo under suitable environmental conditions [119, 121, 160]. Seedlings or early
sporophytes can be obtained from the gametophyte collections of a germplasm kelp
Fig. 2.14 Images of the embryogenic tanks (above) where gametophyte reproduction is induced
after the fertilization of seedling strings (bottom) (Photos by C. Peteiro)
54 C. Peteiro
bank for sea culture and for the repopulation of coastal areas that have been degraded
by human activities and/or natural processes [161–164].
For indoor production of seedlings, a selection of cultures with a suitable propor-
tion of male and female gametophytes (generally a 1:1 sex ratio) are sprayed onto
strings that are wound around a rigid coil or frame called a collector. To promote the
attachment of gametophytes, the strings are pretreated by boiling followed by succes-
sive washes with distilled water, bidirectional sanding and, finally, a surface burn
using hot air guns to remove any filaments produced by sanding [119, 165]. The col-
lectors with the strings seeded with gametophytes are immersed in embryogenesis
tanks, in which temperature, light (irradiance, the light spectrum and photoperiod) and
water movement (by aeration) are under absolute control (Fig. 2.14). In these tanks,
which contain sterile seawater enriched with nitrates and phosphates, sexual repro-
duction is induced under environmental conditions specific to each kelp species [119,
149, 150, 156], which allows zygotes to be fertilized, first giving rise to embryos and
later to seedlings (i.e. early sporophytes) (Figs. 2.10 and 2.14). Generally, seedlings
are embryos with polystromatic thalli (i.e. composed of many layers of cells) that are
normally more than 2 mm long, in which the stipe-blade area begins to differentiate.
To grow young sporophytes in the sea, the seedling strings are primarily attached to
the culture ropes by two methods (Fig. 2.15). In the first, a continuous string is heli-
cally wound around the culture rope, while in the second, pieces of cut string are
woven into the structure of the culture rope at regular intervals.
The culture ropes are deployed in the sea in floating culture rafts, the main ele-
ments of which include an anchoring system, a floating structure and culture lines.
Figure 2.16 shows the different culture rope arrangements that are usually used in
kelp mariculture [110, 113, 127–129, 149]. The hanging rope culture is used in pro-
tected areas, while the horizontal rope culture is used in the most exposed areas, as it
better resists strong waves and currents. In horizontal culture, the light is homoge-
neous along the rope; thus, production is greater, and it may be used in shallow areas
and highly turbid conditions. In the hanging culture, light decreases with depth, so
growth is irregular. To minimize this effect, this approach is usually used in clear
water and within the optimal depth range of the species. Culture rafts may be config-
ured to regulate the depth of the culture ropes and thus control light conditions, so
rafts can be adapted to the needs of the kelp species or the individual culture.
In Saccharina mariculture in Asia, three different methods have been used: two-
year cultivation, forced cultivation and cultivation by transplanting [128, 129, 166].
Due to the natural biannual growth cycle of these kelps, two-year cultivation requires
approximately 20 months of cultivation in the sea to obtain sporophytes of com-
mercial size. An alternative method has been developed in Asia that involves the
production of seedlings during the summer, allowing earlier cultivation and thus
reducing the growth period in the sea to 10 months to obtain adult sporophytes. This
method of sea cultivation, termed forced cultivation, expanded rapidly and became
the main method for the commercial cultivation of Saccharina in Asia. In cultiva-
tion by transplanting, young sporophytes obtained by thinning cultures are normally
2 Alginate Production from Marine Macroalgae, with Emphasis on Kelp Farming 55
Fig. 2.15 Methods of attaching kelp seedling strings to culture ropes for later sea culture
Fig. 2.16 Floating raft culture with different rope arrangements: hanging rope method (vertical
type or garland type) and horizontal rope method (long-line type) (Adapted and reprinted from
Ref. [113], Copyright 2016, with permission from Elsevier)
apical part of the blade, leaving the basal part, which may regrow apically from the
basal blade meristem. However, the most common harvesting method is the collec-
tion of all sporophytes when most have reached commercial size [113, 127, 128].
The productivity of kelp cultures varies with the species, cultivation method,
growing season, environmental conditions at the cultivation site and many other
factors. However, the approximate average wet weight biomass yield per hectare of
cultivation on a commercial farm has been, for example, 70 tons fresh weight for
Macrocystis pyrifera [149], 26 tons for Saccharina japonica [129] and 25 tons for
Saccharina latissima [169].
The commercial alginates that have numerous applications are exclusively extracted
from brown seaweeds, and it is estimated that the global production of alginates pri-
marily involves the exploitation of only 9 seaweed species, of which kelps Lessonia,
2 Alginate Production from Marine Macroalgae, with Emphasis on Kelp Farming 57
Fig. 2.17 Time frames for Saccharina latissima mariculture in southern Europe. As with peren-
nial kelp species, sporophyte growth occurs when the sea temperature is lowest and nitrogen avail-
ability is high
58 C. Peteiro
Laminaria and Saccharina are the most commercially important, accounting for 86%
of the worldwide production. Most of these marine macroalgae are currently har-
vested from native populations (over 80%), while Saccharina farming in Asia pro-
vides the rest of the resources (approximately 20%) for alginate production.
The demand for alginates is expected to increase in the future; however, natural
resources are limited, and kelp forests are decreasing worldwide. Nevertheless, it is
expected that the contribution of kelp cultures to global alginate production in the
coming years will increase in volume and in the number of species used for this
purpose, which would also provide greater security and stability to the market sup-
ply of alginates, as it would no longer depend on natural populations. Additionally,
this cultivation would enable alginates of commercial interest to be obtained from
species whose extraction has not been previously possible due to a lack of available
natural resources.
Finally, kelp farming provides significant environmental benefits by capturing
atmospheric carbon and recycling inorganic nutrients from the marine environment.
Additionally, since kelps have many other applications, the uses of the biomass
harvested in cultures could be integrated in biofactories so that other products of
commercial value, besides alginates, could be obtained (Fig. 2.18).
Fig. 2.18 Kelp sea farming scheme to produce alginates as well as other value-added bioproducts
from the integrated use of the kelp biomass harvested in biofactories (Adapted and reprinted from
Ref. [113], Copyright 2016, with permission from Elsevier)
2 Alginate Production from Marine Macroalgae, with Emphasis on Kelp Farming 59
Acknowledgements We would like to acknowledge the assistance of Dr. A. Secilla in the elabo-
ration of the figures. The English language has been edited by American Journal Experts (AJE).
References
1. Haug A (1964) Composition and properties of alginates. Rep Norw Inst Seaweed Res No. 30.
Norwegian Institute of Seaweed Research, Trondheim, pp 1–123
2. Percival E (1979) The polysaccharides of green, red and brown seaweeds: their basic struc-
ture, biosynthesis and function. Br Phycol J 14(2):103–117
3. Grasdalen H (1983) High-field, 1H-n.m.r. spectroscopy of alginate: sequential structure and
linkage conformations. Carbohydr Res 118:255–260
4. Smidsrød O, Glover RM, Whittington SG (1973) The relative extension of alginates having
different chemical composition. Carbohydr Res 27(1):107–118
5. Braccini I, Grasso RP, Pérez S (1999) Conformational and configurational features of acidic
polysaccharides and their interactions with calcium ions: a molecular modeling investigation.
Carbohydr Res 317(1/4):119–130
6. Draget KI, Gåserød O, Aune I, Andersen PO, Storbakken B, Stokke BT, Smidsrød O (2001)
Effects of molecular weight and elastic segment flexibility on syneresis in Ca-alginate gels.
Food Hydrocoll 15(4/6):485–490
7. Lee KY, Mooney DJ (2012) Alginate: properties and biomedical applications. Prog Polym
Sci 37(1):106–126
8. Andersen T, Strand BL, Formo K, Alsberg E, Christensen BE (2012) Alginates as biomateri-
als in tissue engineering. Carbohydr Chem 37:227–258
9. Helgerud T, Gåserød O, Fjæreide T, Andersen PO, Larsen CK (2009) Alginates. In: Imeson
A (ed) Food stabilisers, thickeners and gelling agents. Wiley-Blackwell, Oxford, pp 50–72
10. Murillo-Álvarez JI, Hernández-Carmona G (2007) Monomer composition and sequence of
sodium alginate extracted at pilot plant scale from three commercially important seaweeds
from Mexico. J Appl Phycol 19(5):545–548
11. Indergaard M, Skjåk-Bræk G (1987) Characteristics of alginate from Laminaria digitata cul-
tivated in a high-phosphate environment. Hydrobiologia 151(152):541–549
12. Reyes Tisnado R, Hernández Carmona G, Rodríguez Montesinos E, Arvizu Higuera DL,
López Gutiérrez F (2005) Food grade alginates extracted from the giant kelp Macrocystis
pyrifera at pilot-plant scale. Rev Invest Mar 26(3):185–192
13. Hartmann M, Dentini M, Ingar Draget K, Skjåk-Bræk G (2006) Enzymatic modification
of alginates with the mannuronan C-5epimerase AlgE4 enhances their solubility at low
pH. Carbohydr Polym 63(2):257–262
14. Smidsrød O, Christensen BE (1991) Molecular structure and physical behavior of seaweed
colloids as compared with microbial polysaccharides. In: Guiry MD, Blunden G (eds)
Seaweeds resources in europe: uses and potential. Wiley, West Sussex, pp 185–217
15. Draget KI, Moe ST, Skjak-Bræk G, Smidsrød O (2006) Alginates. In: Stephen AM, Phillips
GO, Williams PA (eds) Food polysaccharides and their applications (Chapter 9). CRC Press/
Taylor & Francis Group, Boca Raton, pp 289–334
16. Rioux LE, Turgeon SL, Beaulieu M (2007) Characterization of polysaccharides extracted
from brown seaweeds. Carbohydr Polym 69(3):530–537
17. Smidsrød O, Draget KI (1996) Chemistry and physical properties of alginates. Carbohydr
Eur 14:6–13
18. Moral CK, Dogan O, Sanin FD (2013) Comparison of chemical fractionation method and
H-1-NMR spectroscopy in measuring the monomer block distribution of algal alginates.
J Polym Eng 33(3):239–246
19. Salomonsen T, Jensen HM, Stenbæk D, Engelsen SB (2008) Chemometric prediction of algi-
nate monomer composition: a comparative spectroscopic study using IR, Raman, NIR and
NMR. Carbohydr Polym 72(4):730–739
60 C. Peteiro
20. Grant GT, Morris ER, Rees DA, Smith PJC, Thom D (1973) Biological interactions between
polysaccharides and divalent cations: the egg-box model. FEBS Lett 32(1):195–198
21. Braccini I, Pérez S (2001) Molecular basis of Ca2+-induced gelation in alginates and pectins:
the egg-box model revisited. Biomacromolecules 2(4):1089–1096
22. Plazinski W (2011) Molecular basis of calcium binding by polyguluronate chains. Revising
the egg-box model. J Comput Chem 32(14):2988–2995
23. Mørch ÝA, Donati I, Strand BL (2006) Effect of Ca2+, Ba2+, and Sr2+ on alginate microbeads.
Biomacromolecules 7(5):1471–1480
24. Smidsrød O, Haug A (1972) Properties of poly(1,4-hexuronates) in the gel state.
II. Comparison of gels of different chemical composition. Acta Chem Scand 26:79–88
25. Haug A, Smidsrød O (1965) The effect of divalent metals on the properties of alginate solu-
tions. II. Comparison of different metal ions. Acta Chem Scand 69(2):341–351
26. Smidsrød O, Haug A (1968) Dependence upon uronic acid composition of some ion-exchange
properties of alginate. Acta Chem Scand 22(6):1989–1997
27. Draget KI, Strand B, Hartmann M, Valla S, Smidsrød O, Skjåk-Bræk G (2000) Ionic
and acid gel formation of epimerised alginates; the effect of AlgE4. Int J Biol Macromol
27(2):117–122
28. Mancini M, Moresi M, Rancini R (1999) Mechanical properties of alginate gels: empirical
characterisation. J Food Eng 39(4):369–378
29. Skjåk-Bræk G, Grasdalen H, Smidsrød O (1989) Inhomogeneous polysaccharide ionic gels.
Carbohydr Polym 10(1):31–54
30. Draget KI, Simensen MK, Onsøyen E, Smidsrød O (1993) Gel strength of Ca-limited algi-
nate gels made in situ. Hydrobiologia 260-261(1):563–565
31. Storz H, Zimmermann U, Zimmermann H, Kulicke W-M (2009) Viscoelastic properties of
ultra-high viscosity alginates. Rheol Acta 49(2):155–167
32. Senuma Y, Lowe C, Zweifel Y, Hilborn JG, Marison I (2000) Alginate hydrogel microspheres
and microcapsules prepared by spinning disk atomization. Biotechnol Bioeng 67(5):616–622
33. Moe ST, Draget KI, Skjåk-Braek G, Smidsrød O (1995) Alginates. In: Stephen AM (ed) Food
polysaccharides and their applications. Marcel Dekker, New York, pp 245–286
34. Draget KI, Taylor C (2011) Chemical, physical and biological properties of alginates and
their biomedical implications. Food Hydrocoll 25(2):251–256
35. Ertesvåg H (2015) Alginate-modifying enzymes: biological roles and biotechnological uses.
Front Microbiol 6:523. article 523
36. Pawar SN, Edgar KJ (2012) Alginate derivatization: a review of chemistry, properties and
applications. Biomaterials 33(11):3279–3305
37. d’Ayala GG, Malinconico M, Laurienzo P (2008) Marine derived polysaccharides for bio-
medical applications: chemical modification approaches. Molecules 13(9):2069–2106
38. Stanford ECC (1883) On algin: a new substance obtained from some of the commoner spe-
cies of marine algae. Chem News 47:254–257
39. Stanford ECC (1881) Improvements in the manufacture of useful products from seaweeds.
British Patent No 142
40. Smith AM, Miri T (2010) Alginates in foods. In: Norton IT, Spyropoulos F, Cox P (eds)
Practical food rheology: an interpretive approach. Wiley-Blackwell, Oxford, pp 113–132
41. Szekalska M, PuciBowska A, SzymaNska E, Ciosek P, Winnicka K (2016) Alginate: current
use and future perspectives in pharmaceutical and biomedical applications. Int J Polym Sci
2016:1–17. article ID 7697031
42. McHugh DJ (2003) A guide to the seaweed industry. FAO fisheries technical paper no. 441.
FAO, Rome
43. Bixler HJ, Porse H (2011) A decade of change in the seaweed hydrocolloids industry. J Appl
Phycol 23(3):321–335
44. Kraan S (2012) Algal polysaccharides, novel applications and outlook. In: Chang C-F
(ed) Carbohydrates: comprehensive studies on glycobiology and glycotechnology. InTech,
Maastricht, pp 489–532
45. Rioux LE, Turgeon SL (2015) Seaweed carbohydrates. In: Tiwari BK, Troy DJ (eds) Seaweed
sustainability: food and non-food applications. Academic, London, pp 141–192
2 Alginate Production from Marine Macroalgae, with Emphasis on Kelp Farming 61
46. Remminghorst U, Rehm BHA (2006) Bacterial alginates: from biosynthesis to applications.
Biotechnol Lett 28(21):1701–1712
47. Rehm BHA, Valla S (1997) Bacterial alginates: biosynthesis and applications. Appl Microbiol
Biotechnol 48(3):281–288
48. Hay ID, Rehman ZU, Moradali MF, Wang Y, Rehm BHA (2013) Microbial alginate produc-
tion, modification and its applications. Microb Biotechnol 6:637–650
49. Schiener P, Black KD, Stanley MS, Green DH (2015) The seasonal variation in the chemical
composition of the kelp species Laminaria digitata, Laminaria hyperborea, Saccharina latis-
sima and Alaria esculenta. J Appl Phycol 27(1):363–373
50. Westermeier R, Murúa P, Patiño DJ, Muñoz L, Ruiz A, Müller DG (2012) Variations of
chemical composition and energy content in natural and genetically defined cultivars of
Macrocystis from Chile. J Appl Phycol 24:1191–1201
51. Critchley AT, Ohno M, Largo DB (2006) The seaweed resources of the world. A CD-rom
project. Expert Centre for Taxonomic Identification (ETI), Amsterdam
52. Zemke-White WL, Ohno M (1999) World seaweed utilisation: an end-of-century summary.
J Appl Phycol 11:369–376
53. Graham L, Graham J, Wilcox L (2009) Algae, 2nd edn. Pearson Benjamin Cummings, San
Francisco
54. Barsanti L, Gualtieri P (2014) Algae: anatomy, biochemistry, and biotechnology. CRC Press,
Taylor & Francis Group, Boca Raton
55. Algaebase (2017) World-wide electronic publication. National University of Ireland, Galway.
http://www.algaebase.org. Accessed 8 Jan 2017
56. Synytsya A, Čopíková J, Kim WJ, Park YI (2015) Cell wall polysaccharides of marine
algae. In: Kim S-K (ed) Handbook of marine biotechnology. Springer, Berlin/Heidelberg,
pp 543–590
57. Kloareg B, Quatrano RS (1988) Structure of the cell walls of marine algae and ecophysiologi-
cal functions of the matrix polysaccharides. Oceanogr Mar Biol Annu Rev 26:259–315
58. Venegas M, Matsuhiro B, Edding ME (1993) Alginate composition of Lessonia trabeculata
(Phaeophyta: Laminariales) growing in exposed and sheltered habitats. Bot Mar 36(1):47–51
59. Craigie JS, Morris ER, Rees DA, Thom D (1984) Alginate block structure in phaeophyceae
from Nova Scotia: variation with species, environment and tissue-type. Carbohydr Polym
4(4):237–252
60. Black WAP (1950) The seasonal variation in weight and chemical composition of the com-
mon British Laminariaceae. J Mar Biol Assoc UK 29:45–72
61. Kelly BJ, Brown MT (2000) Variations in the alginate content and composition of Durvillaea
antarctica and D. willana from southern New Zealand. J Appl Phycol 12(3/5):317–324
62. Andresen IL, Skipnes O, Smidsrød O, Ostgaard K, Hemmer P (1977) Some biological func-
tions of matrix components in benthic algae in relation to their chemistry and the composition
of seawater. Cell Chem Technol 48:361–381
63. Hernández-Carmona G (1985) Variación estacional del contenido de alginatos en tres espe-
cies de feofitas de Baja California Sur. Invest Marinas CICIMAR 2:29–45
64. Manns D, Nielsen MM, Bruhn A, Saake B, Meyer AS (2017) Compositional variations
of brown seaweeds Laminaria digitata and Saccharina latissima in Danish waters. J Appl
Phycol 29:1493–1506. in press
65. Minghou J, Yujun W, Zuhong X, Yucai G (1984) Studies on the M:G ratios in alginate. In:
Eleventh international seaweed symposium. Dr W. Junk Publishers, Dordrecht, pp 554–556
66. Honya M, Kinoshita T, Ishikawa M, Mori H, Nisizawa K (1993) Monthly determination of
alginate, M/G ratio, mannitol, and minerals in cultivated Laminaria japonica. Nippon Suisan
Gakk 59(2):295–299
67. Chandía N (2001) Alginic acids in Lessonia trabeculata: characterization by formic acid
hydrolysis and FT-IR spectroscopy. Carbohydr Polym 46(1):81–87
68. Lorbeer AJ, Charoensiddhi S, Lahnstein J, Lars C, Franco CMM, Bulone V, Zhang W (2017)
Sequential extraction and characterization of fucoidans and alginates from Ecklonia radi-
ata, Macrocystis pyrifera, Durvillaea potatorum, and Seirococcus axillaris. J Appl Phycol
29:1515–1526. in press
62 C. Peteiro
69. Panikkar R, Brasch DJ (1996) Composition and block structure of alginates from New
Zealand brown seaweeds. Carbohydr Res 293(1):119–132
70. Obluchinskaya ED, Voskoboinikov GM, Galynkin VA (2002) Contents of alginic acid and
fuccidan in Fucus algae of the Barents Sea. Appl Biochem Microbiol 38(2):186–188
71. Nai-yu Z, Yan-xia Z, Xiao F, Li-jun H (1994) Effects of composition and structure of algi-
nates on adsorption of divalent metals. Chin J Oceanol Limnol 12(1):78–83
72. Indergaard M, Skjak-Braek G, Jensen A (1990) Studies on the influence of nutrients on the
composition and structure of alginate in Laminaria saccharina (L.) Lamour. (Laminariales,
Phaeophyceae). Bot Mar 33:277–288
73. Storz H, Müller KJ, Ehrhart F, Gómez I, Shirley SG, Gessner P, Zimmermann G, Weyand E,
Sukhorukov VL, Forst T, Weber MM, Zimmermann H, Kulicke W-M, Zimmermann U (2009)
Physicochemical features of ultra-high viscosity alginates. Carbohydr Res 344(8):985–995
74. Yuguchi Y, Urakawa H, Kajiwara K, Draget KI, Stokke BT (2000) Small-angle X-ray scatter-
ing and rheological characterization of alginate gels. 2. Time-resolved studies on ionotropic
gels. J Mol Struct 554(1):21–34
75. Rosell K-G, Srivastava LM (1984) Seasonal variation in the chemical constituents of the
brown algae Macrocystis integrifolia and Nereocystis luetkeana. Can J Bot 62(11):2229–2236
76. Franklin LA, Forster RM (1997) The changing irradiance environment: consequences for
marine macrophyte physiology, productivity and ecology. Eur J Phycol 32:207–232
77. Harrison PJ, Hurd CL (2001) Nutrient physiology of seaweeds: application of concepts to
aquaculture. Cah Biol Mar 42:71–82
78. Davison IR (1991) Environmental effects on algal photosynthesis: temperature. J Phycol
27:2–8
79. Cheshire AC, Hallam ND (1985) The environmental role of alginates in Durvillaea potato-
rum (Fucales, Phaeophyta). Phycologia 24(2):147–153
80. FMC (2016) FMC Biopolymer FMC Health & Nutrition – Pharmaceutical FMC Corporation
http://www.fmcbiopolymer.com. Accessed 25 Dec 2016
81. Hernández-Carmona G, Freile-Pelegrín Y, Hernández-Garibay E (2013) Conventional
and alternative technologies for the extraction of algal polysaccharides. In: Dominguez H
(ed) Functional ingredients from algae for foods and nutraceuticals. Woodhead Publishing
Limited, Cambridge, pp 475–516
82. McHugh DJ (1987) Production, properties and uses of alginates. In: DJ MH (ed) Production
and utilization of products from commercial seaweeds. FAO fisheries technical paper, 288.
Food and Agriculture Organization of the United Nations (FAO), Fishery and Aquaculture
Economics and Policy Division, Rome
83. Hernández-Carmona G, McHugh DJ, Arvizu-Higuera DL, Rodríguez-Montesinos YE (2002)
Pilot plant scale extraction of alginates from Macrocystis pyrifera. 4. Conversion of alginic
acid to sodium alginate, drying and milling. J Appl Phycol 14(6):445–451
84. Hernández-Carmona G, McHugh DJ, López-Gutiérrez F (1999) Pilot plant scale extraction
of alginates from Macrocystis pyrifera. 2. Studies on extraction conditions and methods of
separating the alkaline-insoluble residue. J Appl Phycol 11(6):493–502
85. Gomez CG, Pérez Lambrecht MV, Lozano JE, Rinaudo M, Villar MA (2009) Influence of the
extraction–purification conditions on final properties of alginates obtained from brown algae
(Macrocystis pyrifera). Int J Biol Macromol 44(4):365–371
86. Hallmann A (2007) Algal transgenics and biotechnology. Transgenic Plant J 1(1):81–98
87. White WL (2015) World seaweed utilization. In: Tiwari BK, Troy DJ (eds) Seaweed sustain-
ability: food and non-food applications. Academic Press/Elsevier, Oxford, pp 7–25
88. Vásquez JA (2016) The brown seaweeds fishery in Chile. In: Mikkola H (ed) Fisheries and
aquaculture in the modern world. InTechOpen, Rijeka, pp 123–141
89. Lane CE, Mayes C, Druehl LD, Saunders GW (2006) A multi-gene molecular investigation
of the kelp (Laminariales, Phaeophyceae) supports substantial taxonomic re-organization.
J Phycol 42(2):493–512
90. Tellier F, Tapia J, Faugeron S, Destombe C, Valero M (2011) The Lessonia nigrescens spe-
cies complex (Laminariales, Phaeophyceae) shows strict parapatry and complete reproduc-
tive isolation in a secondary contact zone. J Phycol 47(4):894–903
2 Alginate Production from Marine Macroalgae, with Emphasis on Kelp Farming 63
113. Peteiro C, Sánchez N, Martínez B (2016) Mariculture of the Asian kelp Undaria pinnatifida
and the native kelp Saccharina lattisima along the Atlantic coast of southern Europe: an
overview. Algal Res 15:9–23
114. Chopin T, Robinson S, Reid G, Ridler N (2013) Prospects for Integrated Multi-Trophic
Aquaculture (IMTA) in the open ocean. Bull Aquacul Assoc Canada 111(2):28–35
115. Rebours C, Marinho-Soriano E, Zertuche-González J, Hayashi L, Vásquez J, Kradolfer P,
Soriano G, Ugarte R, Abreu M, Bay-Larsen I, Hovelsrud G, Rødven R, Robledo D (2014)
Seaweeds: an opportunity for wealth and sustainable livelihood for coastal communities.
J Appl Phycol 26(5):1939–1951
116. Cottier-Cook EJ, Nagabhatla N, Badis Y, Campbell M, Chopin T, Dai W, Fang J, He P, Hewitt
CL, Kim GH, Huo Y, Jiang Z, Kema G, Li X, Liu F, Liu H, Liu Y, Lu Q, Luo Q, Mao Y, Msuya
FE, Rebours C, Shen H, Stentiford GD, Yarish C, Wu H, Yang X, Zhang J, Zhou Y, Gachon
CMM (2016) Safeguarding the future of the global seaweed aquaculture industry. United
Nations University, Institute for Water, Environment and Health (UNU-INWEH) & Scottish
Association for Marine Science (SAMS), Hamilton
117. Kanda T (1936) On the gametophytes of some japanese species of Laminariales. Sci Pap Inst
Algol Res, Fac Sci, Hokkaido Imp Univ 1(2):221–260
118. Schiel DR, Foster MS (2015) The biology and ecology of giant kelp forests. University of
California Press, Oakland
119. Peteiro C (2015) Open-sea cultivation of commercial kelps in the Atlantic coast of southern
Europe. Ph.D. thesis, King Juan Carlos University (URJC), Madrid
120. Perez R, Kaas R, Barbaroux O (1984) Culture expérimentale de l’algue Undaria pinnatifida
sur les côtes de France. Sci Pêche 343:3–15
121. Barrento S, Camus C, Sousa-Pinto I, Buschmann AH (2016) Germplasm banking of the giant
kelp: our biological insurance in a changing environment. Algal Res 13:134–140
122. Lüning K (1990) Seaweeds: their environment, biogeography and ecophysiology. Wiley,
New York
123. Breeman AM (1988) Relative importance of temperature and other factors in determining
geographic boundaries of seaweeds: experimental and phenological evidence. Helgoländer
Meeresunters 42(2):199–241
124. van den Hoek C (1982) The distribution of benthic marine algae in relation to the temperature
regulation of their life histories. Biol J Linnean Soc 18:81–144
125. Lüning K (1982) Seasonality of larger brown algae and its possible regulation by the environ-
ment. In: Srivastava LM (ed) Synthetic and degradative processes in marine macrophytes.
Walter de Gruyter, Berlin, pp 47–67
126. Hurd CL (2000) Water motion, marine macroalgal physiology and production. J Phycol
36(3):453–472
127. Saito Y (1975) Undaria. In: Tokida J, Hirose H (eds) Advance in phycology in Japan. Dr.
W. Junk, Hague, pp 304–320
128. Kawashima S (1984) Kombu cultivations in Japan for human foodstuff. Jpn J Phycol
32:379–394
129. Chen J (2004) Laminaria japonica (Areschoug, 1851). In: Cultured aquatic species informa-
tion programme. FAO Fisheries and Aquaculture Department, Rome
130. Kain (Jones) JM (1991) Cultivation of attached seaweeds. In: Guiry MD, Blunden G (eds)
Seaweeds resources in Europe: uses and potential. Wiley, West Sussex, pp 309–378
131. Brinkhuis BH, Levine HG, Schlenk CG, Tobin S (1987) Laminaria cultivation in the far
east and north America. In: Bird KT, Benson PH (eds) Seaweed cultivation for renew-
able resources. Developments in aquaculture and fisheries science. Elsevier, Amsterdam,
pp 107–146
132. Druehl LD, Baird R, Lindwall A, Lloyd KE, Pakula S (1988) Longline cultivation of some
Laminariaceae in British Columbia, Canada. Aquacult Fish Manag 19:253–263
133. Pérez-Cirera JL, Salinas JM, Cremades J, Bárbara I, Granja A, Veiga AJ, Fuertes C (1997)
Cultivo de Undaria pinnatifida (Laminariales, Phaeophyta) en Galicia. NACC Biol 7:3–28
2 Alginate Production from Marine Macroalgae, with Emphasis on Kelp Farming 65
134. Perez R, Kaas R, Defend JF, Le Bayon N, Planchon G (1991) Production de gamétophytes
en “free living” pour la culture de l’algue brune Laminaria japonica (Aresch.) Aqua Revue
37:30–37
135. Ritschard RL (1992) Marine algae as a CO2 sink. Water Air Soil Pollut 64(1/2):289–303
136. Chung I, Beardall J, Mehta S, Sahoo D, Stojkovic S (2011) Using marine macroalgae for
carbon sequestration: a critical appraisal. J Appl Phycol 23(5):877–886
137. Gao K, McKinley KR (1994) Use of macroalgae for marine biomass production and CO2
remediation: a review. J Appl Phycol 6:45–60
138. Muraoka D (2004) Seaweed resources as a source of carbon fixation. Bull Fish Res Agen
(Jpn) 1:59–64
139. Kim JK, Kraemer GP, Yarish C (2015) Use of sugar kelp aquaculture in Long Island Sound
and the Bronx River Estuary for nutrient extraction. Mar Ecol Prog Ser 531:155–166
140. Handå A, Forbord S, Wang X, Broch OJ, Dahle SW, Størseth TR, Reitan KI, Olsen Y,
Skjermo J (2013) Seasonal- and depth-dependent growth of cultivated kelp (Saccharina
latissima) in close proximity to salmon (Salmo salar) aquaculture in Norway. Aquaculture
414(415):191–201
141. Sanderson JC, Dring MJ, Davidson K, Kelly MS (2012) Culture, yield and bioremedia-
tion potential of Palmaria palmata (Linnaeus) Weber & Mohr and Saccharina latissima
(Linnaeus) C.E.Lane, C.Mayes, Druehl & G.W.Saunders adjacent to fish farm cages in north
west Scotland. Aquaculture 354:128–135
142. Reid GK, Chopin T, Robinson SMC, Azevedo P, Quinton M, Belyea E (2013) Weight ratios
of the kelps, Alaria esculenta and Saccharina latissima, required to sequester dissolved inor-
ganic nutrients and supply oxygen for Atlantic salmon, Salmo salar, in Integrated Multi-
Trophic Aquaculture systems. Aquaculture 408(409):34–46
143. Chopin T, MacDonald B, Robinson S, Cross S, Pearce C, Knowler D, Noce A, Reid G,
Cooper A, Speare D, Burridge L, Crawford C, Sawhney M, Ang KP, Backman C, Hutchinson
M (2013) The canadian integrated multi-trophic aquaculture network (CIMTAN)-a network
for a new era of ecosystem responsible aquaculture. Fisheries 38(7):297–308
144. Neori A (2007) Essential role of seaweed cultivation in integrated multi-trophic aquaculture
farms for global expansion of mariculture: an analysis. J Appl Phycol 20(5):117–120
145. Neori A, Chopin T, Troell M, Buschmann AH, Kraemer GP, Halling C, Shpigel M, Yarish C
(2004) Integrated aquaculture: rationale, evolution and state of the art emphasizing seaweed
biofiltration in modem mariculture. Aquaculture 231(1/4):361–391
146. Buschmann AH, Hernández-González MC, Astudillo C, de la Fuente L, Gutierrez A, Aroca
G (2005) Seaweed cultivation, product development and integrated aquaculture studies in
Chile. World Aquac 36(3):51–53
147. Buschmann AH, Stead RA, Hernández-González MC, Pereda SV (2013) Un análisis crítico
sobre el uso de macroalgas como base para una acuicultura sustentable. Rev Chil Hist Nat
86:251–264
148. Chopin T, Robinson SMC, Troell M, Neori A, Buschmann AH, Fang JG (2008) Multitrophic
integration for sustainable marine aquaculture. In: Jørgensen SE, Fath BD (eds) Ecological
engineering, encyclopedia of ecology, vol 3. Elsevier, Oxford, pp 2463–2475
149. Camus C, Infante J, Buschmann AH (2017) Overview of 3 year precommercial seafarming of
Macrocystis pyrifera along the Chilean coast. Rev Aquac (in press)
150. Edwards M, Watson L (2011) Cultivating Laminaria digitata. aquaculture explained, no. 26.
Bord Iascaigh Mhara, Dublin
151. Tamigneaux E, Licois A, Bourdages D, Leblanc MJ (2013) Protocoles pour la culture de la
laminaire à long stipe (Saccharina longicruris) et de la laminaire sucrée (Saccharina latis-
sima) dans le contexte du Québec, vol Guide No. 13-01. Les publications MERINOV. Centre
d’Innovation de l’aquaculture et des pêches du Québec, Québec
152. Westermeier R, Patino D, Piel MI, Maier I, Mueller DG (2006) A new approach to kelp mari-
culture in Chile: production of free-floating sporophyte seedlings from gametophyte cultures
of Lessonia trabeculata and Macrocystis pyrifera. Aquac Res 37(2):164–171
66 C. Peteiro
153. Arbona JF, Molla M (2006) Cultivation of brown seaweed Alaria esculenta. Aquaculture
explained. Bord Iascaigh Mhara, Dublin
154. Peteiro C, Freire Ó (2011) Effect of water motion on the cultivation of the commercial
seaweed Undaria pinnatifida in a coastal bay of Galicia, Northwest Spain. Aquaculture
314(1/4):269–276
155. Camus C, Buschmann AH (2017) Macrocystis pyrifera aquafarming: Production optimiza-
tion of rope-seeded juvenile sporophytes. Aquaculture 468:107–114
156. Wang W, Sun X, Wang G, Xu P, Wang X, Lin Z, Wang F (2010) Effect of blue light on indoor
seedling culture of Saccharina japonica (Phaeophyta). J Appl Phycol 22:737–744
157. Zhang QS, Cong YZ, Qu SC, Luo SJ, Li XJ, Tang XX (2008) A simple and highly efficient
method for the cryopreservation of Laminaria japonica (Phaeophyceae) germplasm. Eur
J Phycol 42(2):209–213
158. Pang SJ, Lüning K (2004) Breaking seasonal limitation: year-round sporogenesis in the
brown alga Laminaria saccharina by blocking the transport of puntative sporulation inhibi-
tors. Aquaculture 240:531–541
159. Forbord S, Skjermo J, Arff J, Handå A, Reitan KI, Bjerregaard R, Lüning K (2012)
Development of Saccharina latissima (Phaeophyceae) kelp hatcheries with year-round pro-
duction of zoospores and juvenile sporophytes on culture ropes for kelp aquaculture. J Appl
Phycol 24(3):393–399
160. Peteiro C, Cortés B, Arroyo NL, García-Tasende M, Vergés A, Martínez B (2016) Creación
de bancos de germoplasma o “semillas” con algas laminarias para su conservación, restaura-
ción ecológica y/o cultivo comercial. Ipac Revista de acuicultura 105:15–19
161. Vasquez X, Gutierrez A, Buschmann AH, Flores R, Farias D, Leal P (2014) Evaluation of
repopulation techniques for the giant kelp Macrocystis pyrifera (Laminariales). Bot Mar
57(2):123–130
162. Correa JA, Lagos NA, Medina MH, Castilla JC, Cerda M, Ramirez M, Martinez E, Faugeron
S, Andrade S, Pinto R, Contreras L (2006) Experimental transplants of the large kelp
Lessonia nigrescens (Phaeophyceae) in high-energy wave exposed rocky intertidal habitats
of northern Chile: experimental, restoration and management applications. J Exp Mar Biol
Ecol 335(1):13–18
163. Carney LT, Waaland JR, Klinger T, Ewing K (2005) Restoration of the bull kelp Nereocystis
luetkeana in nearshore rocky habitats. Mar Ecol Prog Ser 302:49–61
164. Terawaki T, Hasegawa H, Arai S, Ohno M (2001) Management-free techniques for restora-
tion of Eisenia and Ecklonia beds along the central Pacific coast of Japan. J Appl Phycol
13(1):13–17
165. Peteiro C, Sánchez N, Dueñas-Liaño C, Martínez B (2014) Open-sea cultivation by trans-
planting young fronds of the kelp Saccharina latissima. J Appl Phycol 26(1):519–528
166. Kawashima S (1993) Cultivation of the brown alga, Laminaria ‘kombu’. In: Ohno M,
Critchley AT (eds) Seaweed cultivation and marine ranching, vol vol 4. Japan International
Cooperation Agency (JICA), Jokosuka, pp 25–40
167. Tseng CK (1987) Laminaria mariculture in China. In: Doty MS, Caddy JF, Santelices B (eds)
Case studies of seven commercial seaweed resources. FAO fisheries technical paper no. 281,
Rome, pp 239–263. Electronic edition. Available at: http://www.fao.org/docrep/X5819E/
x5819e09.htm#laminaria%20mariculture%20in%20china
168. Nanba N, Fujiwara T, Kuwano K, Ishikawa Y, Ogawa H, Kado R (2011) Effect of water
flow velocity on growth and morphology of cultured Undaria pinnatifida sporophytes
(Laminariales, Phaeophyceae) in Okirai Bay on the Sanriku coast, Northeast Japan. J Appl
Phycol 23(6):1023–1030
169. Peteiro C, Freire Ó (2013) Biomass yield and morphological features of the seaweed
Saccharina latissima cultivated at two different sites in a coastal bay in the Atlantic coast of
Spain. J Appl Phycol 25(1):205–213
Chapter 3
Alginate Microcapsules for Drug Delivery
Both, Edorta Santos-Vizcaino and Rosa Maria Hernandez are corresponding authors.
A. Gonzalez-Pujana • G. Orive • J.L. Pedraz • E. Santos-Vizcaino (*) • R.M. Hernandez (*)
NanoBioCel Group, Laboratory of Pharmaceutics, School of Pharmacy, University of the
Basque Country UPV/EHU, Vitoria, Spain
Biomedical Research Networking Centre in Bioengineering, Biomaterials and Nanomedicine
(CIBER-BBN), Vitoria, Spain
e-mail: edorta.santos@ehu.eus; rosa.hernandez@ehu.es
3.1 Introduction
Noncommunicable diseases (NCDs), also known as chronic diseases, are those not
passed from person to person and usually associated with long duration and slow
progression. NCDs include cardiovascular diseases, cancer, chronic respiratory
diseases, diabetes, or neurodegenerative diseases. According to the World Health
Organization, NCDs represent one of the major health challenges of the twenty-first
century, considering both patient suffering and the subsequent socioeconomic
burden [1, 2].
Their management is complex, as many of them demand a tight regulation of
therapeutic factors based on physiological requirements. Hence, conventional drug
administration offers poor control over this type of pathologies, leading in many
cases to non-efficient treatments and undesirable effects. Consequently, over the last
decades, new strategies have been thoroughly studied in order to develop new drug
delivery systems.
One of the concepts that have shown a high potential to become a viable thera-
peutic option for chronic diseases is cell microencapsulation. In this strategy, cells
that produce therapeutically active biomolecules are enveloped in a polymeric
matrix. The resulting microspheres are usually coated with a polycation in order to
form a semipermeable membrane. Thus, the system allows the ingress of nutrients
and oxygen and the egress of therapeutic factors. Furthermore, the passage of
immune cells and antibodies is restricted, leading to immunoprotection of the
encapsulated cells (Fig. 3.1). Hence, the transplantation of encapsulated cells may
Fig. 3.1 Scheme representing the main properties of alginate cell microcapsules
3 Alginate Microcapsules for Drug Delivery 69
Fig. 3.2 (a) Structure of β-D-mannuronic acid (M) (top left), structure of α-L-guluronic acid (G)
(top right). Representation of “egg-box” model binding of α-L-guluronic acid blocks to calcium
ions in alginic acid (bottom) (Reproduced from Ref. [11] by permission of the Royal Society of
Chemistry). (b) Graphical description of the three possible junctions in alginate gels. (a) GG/GG
junctions, (b) MG/MG junctions, and (c) mixed GG/MG junctions (Reprinted with permission
from Ref. [13]. Copyright 2005 American Chemical Society)
procedures may not tolerate cell survival to the same extent ionic cross-linking
does. Hence, ionotropic gelation represents a good alternative to entrap cells in
alginate hydrogels under mild conditions [21].
Consequently, sol-gel processes based on the original extrusion method devel-
oped by Lim and Sun [9] are the most widely employed to manufacture alginate
capsules. Cell entrapment within the hydrogel starts by suspending cells in an aque-
ous solution of the polymer, giving rise to the sol flowing phase. The suspension is
then extruded into a solution of the cross-linking agent, with the subsequent sol-gel
transition that forms the microbeads.
Although the obtained beads may protect allogeneic cells, alginates are too
porous to provide xenograft immunoprotection [6, 22]. Therefore, microbeads are
usually coated with a polycation that controls the molecular weight cutoff of the
biosystem by forming a semipermeable membrane. This permselective barrier
allows the diffusion of the therapeutic factor, oxygen, nutrients, and cellular waste
while it protects the implant from the host immune system and mechanical stress. In
this regard, poly-L-lysine (PLL) [23, 24] and poly-L-ornithine (PLO) [25, 26] are
the most frequently employed polycations. As positively charged ions, when
implanted in the organism, they may trigger a strong immunological reaction [27–
30]. For this reason, a second coating of diluted alginate is usually added with the
aim to mask the positive charges and improve graft biocompatibility. Therefore,
when performing both coatings “alginate-poly-L-lysine-alginate” and “alginate-
poly-L-ornithine-alginate” (APA), microcapsules are obtained. Nevertheless, it has
been reported that those two coatings are not multilayered but instead blend forming
an external layer of PLL and alginate that surrounds the inner calcium-alginate core
[31]. If the degree of interaction between the molecules is not sufficient, unbound
PLL may be exposed at the surface of microcapsules, fact that might explain the
immune response that APA microcapsules cause and opens the debate about the
importance of the second coating based on alginate [32, 33].
Therefore, these coatings are still nowadays one of the most limiting factors of
cell microencapsulation technology due to the poor biocompatibility, mechanical
strength, and stability that PLL and PLO provide [34]. Being the latter a relevant
drawback, it has widely been suggested that other coatings should be tested in order
to overcome these limitations [35]. Chitosan [36–38], modified chitosan [39],
poly(methylene-co-guanidine) [30, 40], and the application of diblock copolymers
of PEG and PLL [41, 42] have been some of the alternative approaches.
3.3 A
lginate Microcapsules as Platform for Controlled
Drug Delivery
supporting cell viability and, as a result, functionality. Thus, the immobilized cells
are able to produce the therapeutic factor de novo in a sustained way from weeks to
months, which may match treatment duration with disease longevity [43].
Consequently, the single application of the treatment would remarkably improve
patient comfort and compliance. Hence, it may overcome the huge problem of
adherence, particularly important in diseases where an exhaustive regulation of
drug delivery is mandatory to achieve treatment efficacy.
This implantation of cells is possible because of the immunoisolation that the
semipermeable membrane confers to the system. Indeed, it has been widely demon-
strated that microencapsulation protects allografts [44]. Furthermore, due to the
shortage of donor tissues for patients, xenograft transplantation has emerged as a
proper alternative. Despite being known that the host’s immune response for xeno-
geneic cells is usually more aggressive, cell encapsulation has been able to prolong
xenograft survival [45, 46]. However, there are still unsolved issues such as the
release of xenogeneic epitopes, which induce the formation of encapsulated tissue-
specific antibodies [47] or pro-inflammatory cytokines [48] that could lead to graft
rejection. For this reason, different approaches have been carried out in order to find
a solution. Examples are incorporating to the alginate the CXCL12 chemokine,
which can repel effector T cells while recruiting immune-suppressive regulatory T
cells [49], and conjugating to the hydrogel a peptide inhibitor for the cell surface
IL-1 receptor (IL-1R) thus blocking the interaction between the entrapped cells and
cytokines [50].
May these issues be overcome, chronic co-administration of immunosuppressant
drugs could be significantly reduced [51]. This fact would have a major impact from
a therapeutic standpoint since the severe side effects produced by these drugs may
be avoided. To date, almost 50 significant adverse effects have been related to the
use of immunosuppressant therapy, so immunoisolation may improve significantly
the life quality of patients [52].
In addition, microcapsules are three-dimensional (3D) scaffolds that mimic more
efficiently the tissues of the body than two-dimensional (2D) hydrogels. Another
advantage of this 3D structure is that the surface to volume ratio is increased com-
pared to conventional bulk hydrogels. This improves nutrient and oxygen supply
[53], overcoming a problem that in many cases leads to graft failure [54].
One of the advantages of using alginate for cell microencapsulation is the abun-
dance because of its natural origin. In particular, the most commonly applied algi-
nates for this purpose are derived from brown algae such as Laminaria digitata,
Laminaria hyperborea, Macrocystis pyrifera, Ascophyllum nodosum, Laminaria
japonica, Durvillaea Antarctica, Ecklonia maxima, Lessonia nigrescens, and
3 Alginate Microcapsules for Drug Delivery 73
Sargassum spp. [55]. Alginates with more defined chemical structures and, conse-
quently, physical properties may be obtained by bacterial biosynthesis. In particular,
alginate is produced by two genera: Pseudomonas and Azotobacter. The relatively
easy modification of bacteria and the recent advances in regulation of the polymer
biosynthesis may enable manufacture of tailor-made alginates with high-value bio-
medical applications [56].
As natural polymers, alginates may incorporate contaminants such as heavy met-
als, proteins, endotoxins, or polyphenolic compounds. Moreover, additional impuri-
ties could be introduced during the industrial extraction processes of raw alginates
[57]. Their presence compromises the biocompatibility of the graft. In particular, a
recent study pointed to endotoxins, or in other words pathogen-associated molecu-
lar patterns (PAMPs), as responsible for the pro-inflammatory responses in the host
when microcapsules are implanted [58]. PAMPs are small molecular motifs found
on pathogens that initiate immune responses to eliminate pathogenic bacteria and
thus protect the host from infections. Cells of the innate immune system recognize
PAMPs via toll-like receptors (TLRs) and other pattern recognition receptors
(PRRs). Some of the PAMPs that raw alginates contain are peptidoglycan, lipotei-
choic acid, or flagellin, and they predominantly activate TLR2, 5, 8, and 9 [58].
When present in alginates, PAMPs lead to immune activation with the subsequent
cytokine release. The majority of them are small enough to pass the membrane, hav-
ing a deleterious effect on cell function and viability. They may also limit the diffu-
sion of nutrients, oxygen, therapeutic molecules, and the waste products of
metabolism [59]. As a consequence of cell death, intracellular components are
released. These molecules are damage- or danger-associated molecules (DAMPs)
which also evoke an inflammation cascade when recognized by PRRs, mainly via
TLRs [53]. Therefore, it is believed that DAMPs also play a role in the responses
against encapsulated cells (Fig. 3.3a) [6].
In order to avoid these problems, purification of alginates has been intensively
studied. In fact, it has been reported that for manufacturing biocompatible alginate
microcapsules that are suitable for cell transplant, the use of ultrapurified alginate is
mandatory [60]. The efforts have given rise to different methods to obtain pure algi-
nates that meet clinically useful criteria [61].
This ultrapurified, “clinical-grade” alginate has been proven to reduce the for-
eign body reaction in many in vivo studies (Fig. 3.3b) [62–64]. Indeed, nowadays
there are commercially available highly purified alginates that evoke no immune
reaction when injected subcutaneously to mice [12]. Moreover, the modification of
alginates has been suggested as a tool to mitigate the foreign body response and
achieve a better biocompatibility [65]. In particular, recently, the modification of
alginates with triazole-thiomorpholine dioxide (TMTD) was able to achieve a long-
term functionality of the graft in immunocompetent animals with no need of immu-
nosuppression [66]. Going further, in a recent study, a technology platform has been
designed to predict whether the purification is efficacious [67]. Finally, it is relevant
to cite that alginate purification is not only beneficial to avoid the immune rejection
of the implant but also to improve viability of the encapsulated cells [68].
74 A. Gonzalez-Pujana et al.
Fig. 3.3 (a) PAMP and DAMP release from microencapsulated cells. (b) Microcapsules produced
with purified or crude alginate (Figure as originally published in [53])
3.3.2 T
herapeutic Factors Delivered Through Cell
Encapsulation
The major advantage of cell encapsulation lies beneath the adaptability of the bio-
system. The optimization of this technology is giving rise to promising treatments
for multiple disorders, since the secretion of a wide range of therapeutic factors is
possible by simply selecting the appropriate cell type or by bioengineering cells so
that they produce the biomolecule of interest. To date, many studies have focused
their attention in specific therapeutically active molecules and their delivery through
cell microcapsules. Table 3.1 classifies some of the most studied ones according to
their nature.
76 A. Gonzalez-Pujana et al.
3.3.2.1 Hormones
Hormones are secretory products that act in specific target cells where they elicit
physiological, morphological, or biochemical responses. Since they regulate funda-
mental bodily and biochemical processes, their use in clinic is extensive. Indeed,
being considered as important biological molecules in terms of clinical utility, hor-
mones as therapeutics are currently under intensive research, and cell encapsulation
represents a valuable strategy for their delivery.
Undoubtedly, one of the most thoroughly studied hormones in cell encapsulation
is insulin. Type 1 diabetes requires a strict exogenous insulin supplementation.
However, the current treatments, such as insulin injections or subcutaneous pumps,
are not able to reproduce physiological profiles of the protein, which results in sec-
ondary complications. With the aim to overcome this issue, multiple studies have
analyzed the possibility of cell microencapsulation as a tool for insulin delivery
[62, 78]. Encapsulation of β-cells from human and even xenogeneic donors (pigs, rats)
has given rise to intensive research. A recent study showed that alginate-encapsulated
3 Alginate Microcapsules for Drug Delivery 77
Neurotrophic factors play a key role in the maintenance of normal neuronal function
in adults and in neuronal survival and differentiation during the stages of develop-
ment. The local upregulation of these factors has been detected close to the site of
lesions such as acute brain injury and neurodegenerative diseases [96]. Since direct
injections of different factors have been demonstrated to improve recovery, the
transplantation of neurotrophic factor-producing cells close to the affected area may
serve as an attractive treatment for these pathologies.
An interesting approach has been the implantation of choroidal plexus (CP)
cells. CP cells are known to secrete multiple biologically active neurotrophic fac-
tors, and, thus, their encapsulation has derived promising systems that enable the
local delivery of these biomolecules for restoration of brain tissue [96, 97]. This
strategy has been successful in pathologies such as AD. AD is the most common
form of dementia, characterized by the presence of extensive deposition of amyloid-β
peptide (Aβ), abnormally phosphorylated tau, and neuronal loss. In a recent study,
alginate microcapsules containing CP were implanted overlying the cerebral cortex
of rats with exogenously induced Aβ memory impairment. Animals presented a
significant recovery, which was attributed to the decrease in apoptosis, and the
increase in neurogenesis, resulting in improved long-term memory [98].
Another example is the brain-derived neurotrophic factor (BDNF) delivery to
prevent deafness-induced auditory neuron degeneration. As it has already been
proven, the administration of exogenous neurotrophins to the deaf cochlea is a suc-
cessful treatment; nevertheless, the benefits are rapidly lost when the therapy stops.
Therefore, alginate microcapsules have been studied as a good alternative for
achieving a sustained release of the factor. In a recent research work, encapsulated
BDNF producing Schwann cells showed significant survival-promoting effects on
the auditory neurons of deaf guinea pigs. Moreover, it was suggested that this treat-
ment in combination with a cochlear implant might enhance and extend the benefits
of the latter [99]. This advantageous effect was demonstrated in a posterior work.
Specifically, it was observed that when cell-based therapy is combined with a
cochlear implant, the enhanced auditory neuron survival effects are translated in
important benefits with respect to electrical stimulation thresholds [100]. Altogether,
the results suggest that this technology may have important clinical benefits in this
area.
Another promising strategy has been the encapsulation of glial cell line-derived
neurotrophic factor (GDNF) and nerve growth factor (NGF) producing cells for the
treatment of neurodegenerative diseases such as Parkinson’s disease (PD). PD is a
80 A. Gonzalez-Pujana et al.
encapsulated cell behavior, giving rise to an easier and less time-consuming method
in comparison to in vivo angiogenesis assays [111].
Therapeutic antibodies are nowadays employed for the treatment of multiple dis-
eases. Therefore, encapsulated cells that produce them hold a great potential for
many applications in drug delivery. A current approach is the utilization of these
systems for cancer management, since the interaction of antibodies with cells from
the immune system modulates their response. This modulation may occur by tam-
pering with receptors involved in immune inhibition or by overstimulating receptors
that promote the cellular immune response. In particular, the encapsulation of
hybridoma cells has been widely studied as a platform for antibody delivery [112].
For instance, microcapsules containing hybridomas that produced anti-CD137 and
anti-OX40 antibodies elicited an efficacious antitumor response by enhancing
tumor-specific cellular immunity [113]. Recently, the capacity of encapsulated cells
producing bispecific antibodies against carcinoembryonic antigens, which are pres-
ent in most colon carcinomas, has been proven [114].
Another alternative is the secretion of endostatin, which is an endogenous anti-
angiogenic peptide that has shown potent antitumor activity. The approach of encap-
sulating endostatin-producing cells has been studied for more than 15 years. In fact,
in 2000, a promising study already showed the efficacy of the treatment obtaining
considerable survival benefits in the immunocompetent BT4C brain tumor model
[115]. Posterior studies confirmed the anti-angiogenic effect of endostatin by show-
ing a significantly reduced tumor vascularization. Nonetheless, the effect on tumor
growth was not observed [116].
The hypoimmunogenic and immunomodulatory properties of stem cells make
them an interesting alternative for cancer management. In particular, alginate-
encapsulated MSCs showed a threefold decrease in cytokine expression compared
to entrapped cell lines [117]. Furthermore, stem cells have inherent tumor-trophic
migratory properties, and the possibility to be modified to express diverse therapeu-
tic factors renders them as optimal vehicles for the targeted delivery to isolated
tumors and metastatic disease [118]. In a relevant example, MSCs were modified to
secrete the angiogenesis inhibitor hemopexin-like protein (PEX). Their administra-
tion adjacent to glioblastoma tumors resulted in a relevant reduction not only in
tumor volume (87%) but also in tumor weight (83%) [117].
The promising results obtained in experimental animal models have led to the con-
duction of several clinical trials in order to move this technology toward clinical
translation. In this regard, this section gathers some relevant examples of clinical
trials concerning cell encapsulation.
82 A. Gonzalez-Pujana et al.
3.4.1 Diabetes
Overall, the results obtained with insulin-producing cell encapsulation hold a great
potential. For this reason, this platform has been studied in clinical trials over the
last 20 years. The first human clinical trial in alginate-encapsulated islet transplanta-
tion dates back to 1994. In this work, encapsulated cadaveric human islets were
implanted intraperitoneally. Capsules were able to establish a glycemic control for
9 months in a type 1 diabetes patient who was on antirejection medications [119].
Since this successful approach, different clinical trials have been performed. Such
is the case of Calafiore et al. who carried out a study where alginate-PLO microen-
capsulated islets were intraperitoneally transplanted without the use of immunosup-
pression. The results did not show side effects of the grafting procedures or any
evidence of immune sensitization, confirming the technique as a powerful tool for
immunoprotection. Moreover, patient’s necessity of exogenous insulin intake
decreased approximately to half of the pretransplantation consumption levels
[120, 121].
Tuch and colleagues carried out a trial with four patients in which allogeneic
islets were transplanted without immunosuppression. Neither insulin requirement
nor glycemic control was altered, and C-peptide, an indirect measurement of insulin
levels, was undetectable by 1–4 weeks. However, in the particular case of a recipient
that received three separate islet infusions, the analysis of C-peptide revealed its
presence up to the next 2.5 years [122].
Recently, Jacobs-Tulleneeers-Thevissen et al. transplanted allogeneic islets in a
patient by means of Ca+2/Ba+2 alginate microcapsules. Three months after transplan-
tation capsules were retrieved and cells remained glucose responsive. However, the
transplant showed insufficient biocompatibility [79]. In another report, Sernova
Corp. announced the Cell Pouch® System, a commercial product of the macroen-
capsulation device that is currently ongoing a Phase I/II clinical trial in diabetic
patients where measures of safety and efficacy are the primary and secondary end-
points [123].
Likewise, the company Living Cell Technologies (LCT) offers the DIABECELL®
product, which is currently in late-stage clinical trials. In 1996, LCT initiated a
novel study addressing the efficacy of xenogeneic islets. In particular, porcine islets
were encapsulated and implanted in the peritoneal cavity of patients without the use
of immunosuppressive therapies. Nine and a half years after transplantation, lapa-
rotomy of one of the patients showed the presence of microcapsules in the perito-
neal cavity. Although the majority contained necrotic islets, impressively, some of
them still remained viable [124]. This fact demonstrated the potential of the
approach, and in consequence, the company has continued the research to achieve a
correct glycemic control without the need immunosuppressants [125]. In a Phase I/
II safety study, tolerability of the implant was confirmed [126]. Moreover, the trial
showed proof of principle of efficacy demonstrating improvement in blood glucose
control, even permitting the discontinuation of insulin injections entirely for up to
32 weeks. Later, a Phase IIa dose-finding trial was performed [127], followed by a
3 Alginate Microcapsules for Drug Delivery 83
Phase IIb safety and efficacy study [128], which resulted in a clinically significant
reduction of insulin dose and unaware hypoglycemia. As of now, a recent newsletter
from the website announced the launch of Phase IIb/III clinical trials [48].
Regarding ICH, a Phase I/II clinical trial has been conducted to evaluate the safety
of its treatment with encapsulated cells that secrete GLP-1 [129]. The goal of this
therapy is to improve the outcome after ICH surgery by enabling the local delivery
of the neuroprotective and anti-inflammatory factor. Thus, the treatment may pro-
mote the healing of the secondary neuronal injury that occurs in the first week after
the bleeding. With this purpose, stroke patients with space-occupying intracerebral
hemorrhage were selected. After surgical evacuation of the hematoma, alginate
microcapsules containing allogeneic mesenchymal cells transfected to produce
GLP-1 were implanted in patients. Grafts were removed by a second surgery after
14 days of treatment. Preliminary results revealed that neither side effects from the
surgical intervention nor implant-related side effects were shown in the interim
evaluation of the first 11 patients. Furthermore, 30% of the implanted cells were
viable and maintained their secretory capacity after explantation [130].
The LCT company is also carrying out clinical trials regarding neurological dis-
eases. In particular, the aim is to encapsulate clusters of neonatal porcine CP cells in
alginate matrices. As already mentioned, CP cells produce a wide variety of neuro-
trophic and neuroprotective factors that support brain health. This product, branded
NTCELL®, is intended for the treatment, without the implementation of immuno-
suppression, of different neurological diseases such as PD, AD, ALS, or Huntington
disease. After obtaining promising preclinical data in a model of PD [131], the com-
pany conducted a Phase I/IIa clinical trial [132]. The study, completed in June 2015,
investigated the safety and clinical effect of the capsules in four patients that had
been diagnosed with PD at least 5 years before. The implants were safe and well
tolerated and improved the clinical symptoms of PD in all of the patients, maintain-
ing the effect for 26 weeks post-implant. In March 2016, a Phase IIb study com-
menced with the purpose of confirming the most effective dose [133]. The company
has claimed that if the obtained results are positive, they will apply for provisional
consent to launch NTCELL as the first disease-modifying treatment for PD in 2017
[134].
84 A. Gonzalez-Pujana et al.
The multiple clinical trials conducted in the field demonstrate the applicability and
potential of cell encapsulation. Nonetheless, the sole alginate capsule does not ful-
fill all the requirements the cells need to survive and accomplish a correct physio-
logical activity. The same happens with crucial aspects such as the host immune
response and the safety issues of the implant. Consequently, supplementary strate-
gies need to be implemented to broaden the possibilities of the system and thus
develop a technology that succeeds in the clinical area.
In their natural niches, cells receive and process information from the ECM [135].
Specifically, cells bind to ECM molecules via integrins giving rise to biochemical
and mechanical signals that regulate myriad cellular processes such as proliferation,
migration, or differentiation. Regarding cell encapsulation in alginate matrices, it is
noticeable that this polymer by itself does not promote cell adhesion. Thus, control-
ling the intracapsular microenvironment by the inclusion of appropriate ligands is of
paramount importance to maintain the viability and correct function of the entrapped
cells.
The biofunctionalization of alginates with ECM motifs can be achieved by incor-
porating full-length ECM molecules as fibronectin, collagen, or laminin [34].
Nonetheless, a promising alternative involves the isolation of functional domains
from these large ECM molecules and their use as short peptide sequences to form
biofunctional matrices. These cell adhesive peptides have the advantage of being
relatively stable and can be easily attached to the hydrogel [136]. On this point,
carbodiimide chemistry has been the most commonly applied method for alginate
functionalization. However, lately, new options have been suggested as efficient
procedures for coupling bioactive molecules to alginates, such as the partial peri-
odate oxidation followed by reductive amination [137]. The most widely employed
motif is the arginine-glycine-aspartic acid (RGD), present in fibronectin, laminin, or
collagen, among other ECM proteins. RGD has been extensively used in the field of
cell microencapsulation with promising results [138–141]. Interestingly, these stud-
ies have highlighted critical aspects such as the ligand density, which has been
described as a crucial factor to consider when elaborating the biomimetic capsules.
Indeed, cytoskeleton organization may differ when encapsulating cells in alginate
matrices with differing degree of substitution (DS). Since DS is defined as the total
number of RGD peptides per alginate chain, different alginate types can be obtained.
In particular, Fig. 3.5 shows a study in which cells were entrapped in four different
alginates: DS 0 (No modified alginate), DS 1 (0.112 mM), DS 5 (0.5 mM), and DS
10 (1.12 mM) [141]. Moreover, the necessity of evaluating the role of RGD for each
cell line has also been stated, since the effects of the peptide sequence may vary.
3 Alginate Microcapsules for Drug Delivery 85
Other isolated ECM moieties that have been employed in alginate encapsulation are
the laminin-derived peptides YIGSR (Tyr-Ile-Gly-Ser-Arg) and IKVAV (Ile-Lys-
Val-Ala-Val) [73, 142].
Considering the importance of matrix functionalization, several novel approaches
are being carried out [143, 144]. An interesting example is the introduction of galac-
tosylated chitosan (GC) in the alginate core. GC provides cells with multiple bind-
ing domains that promote the cell-matrix interactions, improving viability and
specific cell functions [144].
A different concept is the inclusion of growth factor-binding domains. Multiple
signaling molecules that elicit essential cellular responses are present in the ECM
given their non-covalent interactions with heparin sulfate proteoglycans. Therefore,
the covalent tethering of heparan sulfate-containing molecules provides the hydro-
gel with binding domains that may sequestrate different growth factors for their
posterior-controlled release [145]. Therefore, the encapsulated cell microenviron-
ment may mimic more efficiently the ECM and, thus, improve cell fate.
One of the strategies to mitigate this unsolved issue has been the co-administration
of anti-inflammatory drugs. The in vivo screening of different anti-inflammatory
molecules has pointed out to dexamethasone (DXM) and curcumin as the most
effective drugs to inhibit reactive oxygen species (ROS) and early inflammatory
proteases. Indeed, the co-encapsulation of cells with the latter has shown to reduce
fibrosis and improve cell function [146]. The effect of DXM has also been studied.
In an interesting approach, different composite drug delivery systems have been
designed to combine APA cell microcapsules with DXM-loaded poly(lactic-co-
glycolic acid) (PLGA) microspheres. The combination has shown an enhanced per-
formance of the system [147, 148], resulting in a promising alternative to prevent
inflammation and, thus, improve the long-term efficacy of the graft, even when
implanting xenogeneic cells [148]. More recently, the co-encapsulation of pancre-
atic islets with the anti-inflammatory drug pentoxifylline (PTX) was demonstrated
to improve the resistance of these cells against host immune cells such as lympho-
cytes [149]. Similarly, antagonists of inflammation-mediators have also been
entrapped in cell microcapsules. With the aim to reduce the inflammation induced
by the high-mobility group box 1 (HMGB1), an antagonist of its receptor in immune
cells, HMGB1 A box, was co-encapsulated. Results revealed that the amount of
TNF-α secreted from macrophages was significantly attenuated. Moreover, an
in vivo study showed a significant improvement in the survival rate of implanted
cells [150].
Another alternative is the modification of the capsule with different motifs that
exert an anti-inflammatory effect. That is the case of sulfated alginates, which are
used as a secondary coating on alginate-PLL microcapsules or mixed in the gel core
of non-coated microbeads, resulting in the reduction of inflammatory cytokines
such as TNF, IL-6, IL-8, or IL-1b [151]. In another example, the attachment of an
inhibitory peptide for cell surface IL-1β enabled the maintenance of cell viability in
the presence of a combination of different cytokines including IL-1β or TNF-α.
Contrarily, cells encapsulated in unmodified hydrogels were unable to survive to the
exposure of these cytokines [50].
A newer strategy to reduce inflammation, and consequently PFO, is the co-
encapsulation of effector-target cells with MSCs. MSCs are known to be hypoim-
munogenic and exert an immunomodulatory effect because of the production of
factors such as prostaglandin E2 or nitric oxide that modulate the immune response
[152–154]. A recent study demonstrated the beneficial effects of co-encapsulated
MSCs in an aggressive xenotransplantation model of mice by showing a dose-
dependent reduction of PFO with the subsequent improvement on graft survival
[155]. This approach has resulted beneficial in different applications including pan-
creatic islet transplantation, where MSCs present a high potential to overcome some
of the current limitations of the system by suppressing inflammatory damage and
immune-mediated rejection [81]. Moreover, it is possible to take advantage of the
important characteristics of MSCs to modulate the neuro-inflammatory response.
Therefore, it has been suggested that alginate encapsulation of MSCs may not only
provide an auxiliary strategy to improve biocompatibility when co-encapsulated
with other effector cells but also a valuable treatment for CNS trauma [156].
3 Alginate Microcapsules for Drug Delivery 87
Despite cell encapsulation holds a great potential, the fact that it implies the use of
living cells leads to safety concerns, particularly considering that in many cases
these cells have been genetically modified. Moreover, once microcapsules are
administered, the determination of their position and integrity is complex. In order
to address this issue, different approaches have been proposed in order to monitor
or inactivate the implants.
Diverse noninvasive imaging techniques have been developed with the aim of
tracking the implants. Cell encapsulation is an asset for monitoring, as the contrast
agents are contained in the hydrogel instead of directly labeling the cells. This
reduces toxicity and promotes cell survival and function. Widely used modalities
are X-ray/computed tomography (CT) [159, 160], magnetic resonance imaging
(MRI) [161], and ultrasound imaging [162]. For instance, in a recent study, alginate
microcapsules with a self-assembled gold nanoparticle (AuNPs) coating resulted in
distinctive contrast and enabled the identification by using a conventional small
animal X-ray micro-CT scanner. In particular, AuNPs were modified with the cat-
ionic 2-(methacryloyloxy)ethyl trimethylammonium chloride (METAC) polymer in
order to enable the electrostatic interaction of AuNPs with the negatively charged
alginate microcapsules. The resulting PMETAC_SH-Au nanoparticles were coated
onto preformed alginate microcapsules (PMETAC_SH-Au MCs) with successful
imaging results (Fig. 3.6a) [160].
The limitation of these techniques is that they do not provide information about
cell viability or functionality. To overcome this issue, reporter genes have been
employed [163–165]. They are typically based on bioluminescence and/or fluores-
cence and provide us with a helpful tool to perform noninvasive, quantitative, and
real-time live image determinations. Regarding imaging techniques, recent
approaches have also studied the possibility of analyzing the host immune response
while the capsules are still implanted [166, 167]. This progress may represent a
pivotal advantage, as it would permit to adjust the therapy if needed. In other words,
in cases when a significant immune response is detected, the pertinent therapy could
be administered in order to alleviate acute inflammation. Thus, the transplantation
regimen would improve, restricting the possibilities of graft failure. This advance is
especially relevant considering that nowadays it is not possible to analyze the host
immune response unless biopsies are obtained.
88 A. Gonzalez-Pujana et al.
Fig. 3.6 (a) X-ray micro-CT reconstructed data of a mouse injected with PMETAC_SH-Au-MCs
(4.7 g L−1) (CTDI 28 mGy) in coronal (a), transverse (b), and sagittal (c) views with the magnifica-
tion of the detail in the square shown in (d). Arrows indicate the MCs. 3D rendering of the same
mouse with the MCs artificially colored in yellow (e) and the detail in the square magnified in (f).
Scale bar 1 cm (Reproduced from Ref. [160] by permission of the Royal Society of Chemistry.
(b–d) Behavior of C2C12-TGL microencapsulated cells in vivo. Microencapsulated C2C12-TGL
cells exhibited light emission after being injected subcutaneously in C57BL/6J mice. Mice treated
3 Alginate Microcapsules for Drug Delivery 89
A determinant aspect regarding the safety of the system is the possibility to inac-
tivate the graft. This inactivation might be useful when the therapy reaches its final
goal or when the system causes undesirable effects. To this end, an interesting
approach is the inclusion of suicide genes in the genome of the encapsulated cells.
These suicide genes induce apoptosis upon external drug administration, which
allows the inactivation of the implant when necessary. Interesting studies were car-
ried out regarding ganciclovir (GCV)-mediated inactivation in encapsulated cells
bearing TGL triple fusion reporter gene, which codifies for the suicide gene herpes
simplex virus type 1 thymidine kinase (HSV1-TK), green fluorescent protein (GFP),
and firefly luciferase (SFGNESTGL) [163, 165]. The in vivo studies demonstrated the
potential of the procedure by providing information about the enclosed cells at
desired time points in a noninvasive way, including the additional advantage of cell
inactivation if required (Fig. 3.6b–d). The design of inducible systems may also
dramatically improve safety. Recently, inducible hepatocyte growth factor (HGF)-
secreting human umbilical cord blood (hUCB)-derived MSCs were produced via
TALEN-mediated genome editing. The TetOn-HGF/hUCB-MSCs were encapsu-
lated in alginate microcapsules and demonstrated an improvement in angiogenesis
in a mouse hind limb ischemia model, proving that these inducible cells are able to
deliver the therapeutic factor in a controlled and effective manner [168]. Therefore,
these strategies may take control over the therapeutic effects exerted by the implanted
cells leading to safer treatments.
The retention of microcapsules within the tissue they are implanted and the suitable
retrieval of the whole implant are necessary premises in this technology. For this
reason, several attempts have been carried out in order to optimize administration
and extraction protocols. Different systems have been proposed to envelop the algi-
nate cell microcapsules, such as calcium phosphate cements [169, 170], mesh bags
[130], or hydrogel-based scaffolds [90]. The latter presents advantageous properties
that permit to select between an invasive administration (as a preformed scaffold)
and a noninvasive injection (as an in situ formed scaffold). Moreover, the scaffolds
may be multifunctionalized by combining in the hydrogel the alginate microcapsules
and other systems containing molecules such as anti-inflammatory drugs [148].
Fig. 3.6 (continued) with 150 mg/kg/day ganciclovir (GCV) for a week showed almost no signal
(b–c). Luciferase activity could be found in cells within the microcapsules 255 days after injection
(b). Quantification of light emission demonstrates an increase in the normalized photon flux during
the first 2 weeks after implantation (c–d), probably due to vascularization of microcapsule plugs.
Non-microencapsulated C2C12-TGL cells displayed a marked decrease in light emission between
days 1 and 15 (d). Conversely, microencapsulated myoblasts increased the emission during that
period of time (d). μE microencapsulated cells, NμE non-microencapsulated cells (Reprinted from
Ref. [163], Copyright 2010, with permission from Elsevier)
90 A. Gonzalez-Pujana et al.
Alginate is the most widely employed biomaterial for cell encapsulation as it pro-
vides the technology with pivotal advantages. In particular, the intrinsic characteris-
tics of alginates make them biocompatible and allow the rapid and simple gelation
under mild conditions, giving rise to scalable encapsulation methods. The obtained
alginate microcapsules represent a potential alternative for the treatment of multiple
chronic diseases that nowadays lack an adequate management with conventional
drug delivery systems. That is the case of disorders with a high prevalence in our
society such as diabetes, Alzheimer’s disease, Parkinson’s disease, chronic anemia,
or cancer.
Over the last decades, important advances have been achieved in the field; how-
ever, there are still major challenges to face in order to move toward a definitive
clinical translation. To this end, several attempts are being carried out, which include
the search for suitable and biocompatible coatings [35], the biofunctionalization of
the matrices, the optimization of safety measures such as inducible Tet-on/off sys-
tems [168], the synthesis of capsules with suitable size and shape [173], or the
selection of the most appropriate administration site [34]. May these hurdles be
overcome, therapeutic cell encapsulation would significantly evolve becoming an
invaluable tool in the medical practice.
References
1. World Health Organization (2014) Global status report on noncommunicable diseases. World
Health Organization, Geneva. doi: ISBN 978 92 4 156485 4
2. Shakir R (2015) Neurodegenerative noncommunicable diseases (Neurology NCDs). Where
are we now? J Neurol Sci 356(1–2):1–2. https://doi.org/10.1016/j.jns.2015.07.005
3. Hunt NC, Grover LM (2010) Cell encapsulation using biopolymer gels for regenerative med-
icine. Biotechnol Lett 32(6):733–742. https://doi.org/10.1007/s10529-010-0221-0
4. Chang TM (2005) Therapeutic applications of polymeric artificial cells. Nat Rev Drug Discov
4(3):221–235. https://doi.org/10.1038/nrd1659
5. Mazzitelli S, Capretto L, Quinci F et al (2013) Preparation of cell-encapsulation devices
in confined microenvironment. Adv Drug Deliv Rev 65(11–12):1533–1555. https://doi.
org/10.1016/j.addr.2013.07.021
6. de Vos P, Lazarjani HA, Poncelet D et al (2014) Polymers in cell encapsulation from an
enveloped cell perspective. Adv Drug Deliv Rev 67-68:15–34. https://doi.org/10.1016/j.
addr.2013.11.005
7. Bisceglie V (1934) Über die antineoplastische Immunität. Z Krebsforsch 40(1):122–140
8. Chang TM (1964) Semipermeable microcapsules. Science 146(3643):524–525
9. Lim F, Sun AM (1980) Microencapsulated islets as bioartificial endocrine pancreas. Science
210(4472):908–910
10. Grant GT, Morris ER, Rees DA et al (1973) Biological interactions between polysaccha-
rides and divalent cations: the egg-box model. FEBS Lett 32(1):195–198. https://doi.
org/10.1016/0014-5793(73)80770-7
11. Marriott AS, Bergstrom E, Hunt AJ et al (2014) A natural template approach to mesoporous
carbon spheres for use as green chromatographic stationary phases. RSC Adv 4(1):222–228.
https://doi.org/10.1039/C3RA44428G
12. Lee KY, Mooney DJ (2012) Alginate: properties and biomedical applications. Prog Polym
Sci 37(1):106–126. https://doi.org/10.1016/j.progpolymsci.2011.06.003
13. Donati I, Holtan S, Morch YA et al (2005) New hypothesis on the role of alternating sequences
in calcium-alginate gels. Biomacromolecules 6(2):1031–1040. https://doi.org/10.1021/
bm049306e
14. Andersen T, Auk-Emblem P, Dornish M (2015) 3D cell culture in alginate hydrogels.
Microarrays (Basel) 4(2):133–161. https://doi.org/10.3390/microarrays4020133
15. Haug A, Smidsrod O (1965) The effect of divalent metals on the properties of alginate solu-
tions. Acta Chem Scand 19:341–351
16. Leong J, Lam W, Ho K et al (2016) Advances in fabricating spherical alginate hydrogels with
controlled particle designs by ionotropic gelation as encapsulation systems. Particuology
24:44–60. https://doi.org/10.1016/j.partic.2015.09.004
17. Morch YA, Donati I, Strand BL et al (2006) Effect of Ca2+, Ba2+, and Sr2+ on alginate micro-
beads. Biomacromolecules 7(5):1471–1480. https://doi.org/10.1021/bm060010d
18. Jeong SI, Jeon O, Krebs MD et al (2012) Biodegradable photo-crosslinked alginate nanofibre
scaffolds with tuneable physical properties, cell adhesivity and growth factor release. Eur
Cell Mater 24:331–343
19. Zhao S, Cao M, Li H et al (2010) Synthesis and characterization of thermo-sensitive semi-
IPN hydrogels based on poly(ethylene glycol)-co-poly(epsilon-caprolactone) macromer,
N-isopropylacrylamide, and sodium alginate. Carbohydr Res 345(3):425–431. https://doi.
org/10.1016/j.carres.2009.11.014
20. Kingsley DM, Dias AD, Corr DT (2016) Microcapsules and 3D customizable shelled micro-
environments from laser direct-written microbeads. Biotechnol Bioeng 113(10):2264–2274.
https://doi.org/10.1002/bit.25987
21. Gasperini L, Mano JF, Reis RL (2014) Natural polymers for the microencapsulation of cells.
J R Soc Interface 11(100):20140817. https://doi.org/10.1098/rsif.2014.0817
92 A. Gonzalez-Pujana et al.
22. Orive G, Hernandez RM, Gascon AR et al (2003) Cell encapsulation: promise and progress.
Nat Med 9(1):104–107. https://doi.org/10.1038/nm0103-104
23. Qian D, Bai B, Yan G et al (2016) Construction of doxycycline-mediated BMP-2 trans-
gene combining with APA microcapsules for bone repair. Artif Cells Nanomed Biotechnol
44(1):270–276. https://doi.org/10.3109/21691401.2014.942458
24. Hu J, Li H, Chi G et al (2015) IL-1RA gene-transfected bone marrow-derived mesenchymal
stem cells in APA microcapsules could alleviate rheumatoid arthritis. Int J Clin Exp Med
8(1):706–713
25. Ibarra V, Appel AA, Anastasio MA et al (2016) This paper is a winner in the undergradu-
ate category for the SFB awards: evaluation of the tissue response to alginate encapsulated
islets in an omentum pouch model. J Biomed Mater Res 104(7):1581–1590. https://doi.
org/10.1002/jbm.a.35769
26. Pareta R, McQuilling JP, Sittadjody S et al (2014) Long-term function of islets encapsulated
in a re-designed alginate microcapsule construct in omentum pouches of immune-competent
diabetic rats. Pancreas 43(4):605–613. https://doi.org/10.1097/MPA.0000000000000107
27. Anderson JM, Rodriguez A, Chang DT (2008) Foreign body reaction to biomaterials. Semin
Immunol 20(2):86–100. https://doi.org/10.1016/j.smim.2007.11.004
28. Rokstad AM, Brekke OL, Steinkjer B et al (2013) The induction of cytokines by polycation
containing microspheres by a complement dependent mechanism. Biomaterials 34(3):621–
630. https://doi.org/10.1016/j.biomaterials.2012.10.012
29. Orning P, Hoem KS, Coron AE et al (2016) Alginate microsphere compositions dictate differ-
ent mechanisms of complement activation with consequences for cytokine release and leuko-
cyte activation. J Control Release 229:58–69. https://doi.org/10.1016/j.jconrel.2016.03.021
30. Rokstad AM, Brekke OL, Steinkjer B et al (2011) Alginate microbeads are complement com-
patible, in contrast to polycation containing microcapsules, as revealed in a human whole
blood model. Acta Biomater 7(6):2566–2578. https://doi.org/10.1016/j.actbio.2011.03.011
31. Tam SK, Dusseault J, Polizu S et al (2005) Physicochemical model of alginate-poly-L-lysine
microcapsules defined at the micrometric/nanometric scale using ATR-FTIR, XPS, and ToF-
SIMS. Biomaterials 26(34):6950–6961. https://doi.org/10.1016/j.biomaterials.2005.05.007
32. Tam SK, Bilodeau S, Dusseault J et al (2011) Biocompatibility and physicochemical char-
acteristics of alginate-polycation microcapsules. Acta Biomater 7(4):1683–1692. https://doi.
org/10.1016/j.actbio.2010.12.006
33. Juste S, Lessard M, Henley N et al (2005) Effect of poly-L-lysine coating on macrophage acti-
vation by alginate-based microcapsules: assessment using a new in vitro method. J Biomed
Mater Res A 72(4):389–398. https://doi.org/10.1002/jbm.a.30254
34. Santos E, Pedraz JL, Hernandez RM et al (2013) Therapeutic cell encapsulation: ten steps
towards clinical translation. J Control Release 170(1):1–14. https://doi.org/10.1016/j.
jconrel.2013.04.015
35. Demont A, Cole H, Marison IW (2016) An understanding of potential and limitations of algi-
nate/PLL microcapsules as a cell retention system for perfusion cultures. J Microencapsul
33(1):80–88. https://doi.org/10.3109/02652048.2015.1134686
36. Zhang W, Zhao S, Rao W et al (2013) A novel core-shell microcapsule for encapsulation and
3D culture of embryonic stem cells. J Mater Chem B Mater Biol Med 2013(7):1002–1009.
https://doi.org/10.1039/C2TB00058J
37. Zheng G, Liu X, Wang X et al (2014) Improving stability and biocompatibility of alginate/
chitosan microcapsule by fabricating bi-functional membrane. Macromol Biosci 14(5):655–
666. https://doi.org/10.1002/mabi.201300474
38. Yang HK, Ham DS, Park HS et al (2016) Long-term efficacy and biocompatibility of encap-
sulated islet transplantation with chitosan-coated alginate capsules in mice and canine models
of diabetes. Transplantation 100(2):334–343. https://doi.org/10.1097/TP.0000000000000927
39. Hillberg AL, Oudshoorn M, Lam JB et al (2015) Encapsulation of porcine pancreatic islets
within an immunoprotective capsule comprising methacrylated glycol chitosan and alginate.
J Biomed Mater Res B Appl Biomater 103(3):503–518. https://doi.org/10.1002/jbm.b.33185
3 Alginate Microcapsules for Drug Delivery 93
40. Orive G, Hernandez RM, Gascon AR et al (2003) Development and optimisation of alginate-
PMCG-alginate microcapsules for cell immobilisation. Int J Pharm 259(1–2):57–68. https://
doi.org/10.1016/S0378-5173(03)00201-1
41. Spasojevic M, Paredes-Juarez GA, Vorenkamp J et al (2014) Reduction of the inflamma-
tory responses against alginate-poly-L-lysine microcapsules by anti-biofouling surfaces of
PEG-b-PLL diblock copolymers. PLoS One 9(10):e109837. https://doi.org/10.1371/journal.
pone.0109837
42. Spasojevic M, Bhujbal S, Paredes G et al (2014) Considerations in binding diblock copoly-
mers on hydrophilic alginate beads for providing an immunoprotective membrane. J Biomed
Mater Res A 102(6):1887–1896. https://doi.org/10.1002/jbm.a.34863
43. Orive G, Santos E, Poncelet D et al (2015) Cell encapsulation: technical and clinical advances.
Trends Pharmacol Sci 36(8):537–546. https://doi.org/10.1016/j.tips.2015.05.003
44. Wang T, Adcock J, Kuhtreiber W et al (2008) Successful allotransplantation of encapsulated
islets in pancreatectomized canines for diabetic management without the use of immunosup-
pression. Transplantation 85(3):331–337. https://doi.org/10.1097/TP.0b013e3181629c25
45. Sun Y, Ma X, Zhou D et al (1996) Normalization of diabetes in spontaneously diabetic cyno-
mologus monkeys by xenografts of microencapsulated porcine islets without immunosup-
pression. J Clin Invest 98(6):1417–1422. https://doi.org/10.1172/JCI118929
46. Luca G, Arato I, Mancuso F et al (2016) Xenograft of microencapsulated Sertoli cells restores
glucose homeostasis in db/db mice with spontaneous diabetes mellitus. Xenotransplantation
23:429. https://doi.org/10.1111/xen.12274
47. Scharp DW, Marchetti P (2014) Encapsulated islets for diabetes therapy: history, current
progress, and critical issues requiring solution. Adv Drug Deliv Rev 67-68:35–73. https://doi.
org/10.1016/j.addr.2013.07.018
48. Qi M (2013) Transplantation of encapsulated pancreatic islets as a treatment for patients with
type 1 diabetes mellitus. Adv Med 2014:429710. https://doi.org/10.1155/2014/429710
49. Chen T, Yuan J, Duncanson S et al (2015) Alginate encapsulant incorporating CXCL12 sup-
ports long-term allo- and xenoislet transplantation without systemic immune suppression.
Am J Transplant 15(3):618–627. https://doi.org/10.1111/ajt.13049
50. Su J, Hu BH, Lowe WL Jr et al (2010) Anti-inflammatory peptide-functionalized hydro-
gels for insulin-secreting cell encapsulation. Biomaterials 31(2):308–314. https://doi.
org/10.1016/j.biomaterials.2009.09.045
51. Orive G, Gascon AR, Hernandez RM et al (2003) Cell microencapsulation technology for
biomedical purposes: novel insights and challenges. Trends Pharmacol Sci 24(5):207–210.
https://doi.org/10.1016/S0165-6147(03)00073-7
52. Olabisi RM (2015) Cell microencapsulation with synthetic polymers. J Biomed Mater Res A
103(2):846–859. https://doi.org/10.1002/jbm.a.35205
53. Paredes Juarez GA, Spasojevic M, Faas MM et al (2014) Immunological and technical
considerations in application of alginate-based microencapsulation systems. Front Bioeng
Biotechnol 2:26. https://doi.org/10.3389/fbioe.2014.00026
54. Marchioli G, van Gurp L, van Krieken PP et al (2015) Fabrication of three-dimensional
bioplotted hydrogel scaffolds for islets of Langerhans transplantation. Biofabrication
7(2):025009. https://doi.org/10.1088/1758-5090/7/2/025009
55. McHugh DJ (2003) A guide to the seaweed industry. Food and Agriculture Organization of
the United Nations, Rome
56. Hay ID, Ur Rehman Z, Moradali MF et al (2013) Microbial alginate produc-
tion, modification and its applications. Microb Biotechnol 6(6):637–650. https://doi.
org/10.1111/1751-7915.12076
57. Hernandez RM, Orive G, Murua A et al (2010) Microcapsules and microcarriers for in situ cell
delivery. Adv Drug Deliv Rev 62(7–8):711–730. https://doi.org/10.1016/j.addr.2010.02.004
58. Paredes-Juarez GA, de Haan BJ, Faas MM et al (2013) The role of pathogen-associated
molecular patterns in inflammatory responses against alginate based microcapsules. J Control
Release 172(3):983–992. https://doi.org/10.1016/j.jconrel.2013.09.009
94 A. Gonzalez-Pujana et al.
59. Dufrane D, Gianello P (2012) Macro- or microencapsulation of pig islets to cure type 1 dia-
betes. World J Gastroenterol 18(47):6885–6893. https://doi.org/10.3748/wjg.v18.i47.6885
60. Montanucci P, Terenzi S, Santi C et al (2015) Insights in behavior of variably formulated
alginate-based microcapsules for cell transplantation. Biomed Res Int 2015:965804. https://
doi.org/10.1155/2015/965804
61. Sondermeijer HP, Witkowski P, Woodland D et al (2016) Optimization of alginate purifica-
tion using polyvinylidene difluoride membrane filtration: Effects on immunogenicity and
biocompatibility of three-dimensional alginate scaffolds. J Biomater Appl 31(4):510–520.
https://doi.org/10.1177/0885328216645952
62. Calafiore R, Basta G (2014) Clinical application of microencapsulated islets: actual prospec-
tives on progress and challenges. Adv Drug Deliv Rev 67-68:84–92. https://doi.org/10.1016/j.
addr.2013.09.020
63. Kim AR, Hwang JH, Kim HM et al (2013) Reduction of inflammatory reaction in the use of
purified alginate microcapsules. J Biomater Sci Polym Ed 24(9):1084–1098. https://doi.org/
10.1080/09205063.2012.735100
64. Basta G, Calafiore R (2011) Immunoisolation of pancreatic islet grafts with no recipient’s
immunosuppression: actual and future perspectives. Curr Diab Rep 11(5):384–391. https://
doi.org/10.1007/s11892-011-0219-6
65. Dolgin E (2014) Encapsulate this. Nat Med 20(1):9–11. https://doi.org/10.1038/nm0114-9
66. Vegas AJ, Veiseh O, Gurtler M et al (2016) Long-term glycemic control using polymer-
encapsulated human stem cell-derived beta cells in immune-competent mice. Nat Med
22(3):306–311. https://doi.org/10.1038/nm.4030
67. Paredes-Juarez GA, de Haan BJ, Faas MM et al (2014) A technology platform to test the
efficacy of purification of alginate. Materials 7(3):2087–2103. https://doi.org/10.3390/
ma7032087
68. Langlois G, Dusseault J, Bilodeau S et al (2009) Direct effect of alginate purification on the
survival of islets immobilized in alginate-based microcapsules. Acta Biomater 5(9):3433–
3440. https://doi.org/10.1016/j.actbio.2009.05.029
69. Bhujbal SV, de Vos P, Niclou SP (2014) Drug and cell encapsulation: alternative delivery
options for the treatment of malignant brain tumors. Adv Drug Deliv Rev 67-68:142–153.
https://doi.org/10.1016/j.addr.2014.01.010
70. Swioklo S, Ding P, Pacek AW et al (2017) Process parameters for the high-scale production
of alginate-encapsulated stem cells for storage and distribution throughout the cell therapy
supply chain. Process Biochem 59:289. https://doi.org/10.1016/j.procbio.2016.06.005
71. Richardson T, Kumta PN, Banerjee I (2014) Alginate encapsulation of human embryonic
stem cells to enhance directed differentiation to pancreatic islet-like cells. Tissue Eng Part A
20(23–24):3198–3211. https://doi.org/10.1089/ten.TEA.2013.0659
72. Erro E, Bundy J, Massie I et al (2013) Bioengineering the liver: scale-up and cool chain
delivery of the liver cell biomass for clinical targeting in a bioartificial liver support system.
Biores Open Access 2(1):1–11. https://doi.org/10.1089/biores.2012.0286
73. Formo K, Cho CH, Vallier L et al (2015) Culture of hESC-derived pancreatic progenitors in
alginate-based scaffolds. J Biomed Mater Res A 103(12):3717–3726. https://doi.org/10.1002/
jbm.a.35507
74. de Vos P, Bucko M, Gemeiner P et al (2009) Multiscale requirements for bioencapsulation
in medicine and biotechnology. Biomaterials 30(13):2559–2570. https://doi.org/10.1016/j.
biomaterials.2009.01.014
75. Zhu J, Marchant RE (2011) Design properties of hydrogel tissue-engineering scaffolds.
Expert Rev Med Devices 8(5):607–626. https://doi.org/10.1586/erd.11.27
76. Gauvin R, Parenteau-Bareil R, Dokmeci MR et al (2012) Hydrogels and microtechnologies for
engineering the cellular microenvironment. Wiley Interdiscip Rev Nanomed Nanobiotechnol
4(3):235–246. https://doi.org/10.1002/wnan.171
77. Borg DJ, Bonifacio E (2011) The use of biomaterials in islet transplantation. Curr Diab Rep
11(5):434–444. https://doi.org/10.1007/s11892-011-0210-2
3 Alginate Microcapsules for Drug Delivery 95
78. Barkai U, Rotem A, de Vos P (2016) Survival of encapsulated islets: More than a membrane
story. World J Transplant 6(1):69–90. https://doi.org/10.5500/wjt.v6.i1.69
79. Jacobs-Tulleneers-Thevissen D, Chintinne M, Ling Z et al (2013) Sustained function of
alginate-encapsulated human islet cell implants in the peritoneal cavity of mice leading
to a pilot study in a type 1 diabetic patient. Diabetologia 56(7):1605–1614. https://doi.
org/10.1007/s00125-013-2906-0
80. Kerby A, Jones ES, Jones PM et al (2013) Co-transplantation of islets with mesenchymal
stem cells in microcapsules demonstrates graft outcome can be improved in an isolated-graft
model of islet transplantation in mice. Cytotherapy 15(2):192–200. https://doi.org/10.1016/j.
jcyt.2012.10.018
81. Figliuzzi M, Bonandrini B, Silvani S et al (2014) Mesenchymal stem cells help pancreatic
islet transplantation to control type 1 diabetes. World J Stem Cells 6(2):163–172. https://doi.
org/10.4252/wjsc.v6.i2.163
82. Iacovacci V, Ricotti L, Menciassi A et al (2016) The bioartificial pancreas (BAP): Biological,
chemical and engineering challenges. Biochem Pharmacol 100:12–27. https://doi.
org/10.1016/j.bcp.2015.08.107
83. Montanucci P, Pennoni I, Pescara T et al (2013) Treatment of diabetes mellitus with microen-
capsulated fetal human liver (FH-B-TPN) engineered cells. Biomaterials 34(16):4002–4012.
https://doi.org/10.1016/j.biomaterials.2013.02.026
84. Tomei AA, Villa C, Ricordi C (2015) Development of an encapsulated stem cell-based ther-
apy for diabetes. Expert Opin Biol Ther 15(9):1321–1336. https://doi.org/10.1517/14712598
.2015.1055242
85. Ngoc PK, Phuc PV, Nhung TH et al (2011) Improving the efficacy of type 1 diabetes therapy
by transplantation of immunoisolated insulin-producing cells. Hum Cell 24(2):86–95. https://
doi.org/10.1007/s13577-011-0018-z
86. Richardson T, Barner S, Candiello J et al (2016) Capsule stiffness regulates the efficiency
of pancreatic differentiation of human embryonic stem cells. Acta Biomater 35:153–165.
https://doi.org/10.1016/j.actbio.2016.02.025
87. Murua A, Orive G, Hernandez RM et al (2011) Emerging technologies in the delivery of
erythropoietin for therapeutics. Med Res Rev 31(2):284–309. https://doi.org/10.1002/
med.20184
88. Orive G, De Castro M, Kong HJ et al (2009) Bioactive cell-hydrogel microcapsules for
cell-based drug delivery. J Control Release 135(3):203–210. https://doi.org/10.1016/j.
jconrel.2009.01.005
89. Orive G, Santos E, Pedraz JL et al (2014) Application of cell encapsulation for con-
trolled delivery of biological therapeutics. Adv Drug Deliv Rev 67-68:3–14. https://doi.
org/10.1016/j.addr.2013.07.009
90. Acarregui A, Pedraz JL, Blanco FJ et al (2013) Hydrogel-based scaffolds for enclosing
encapsulated therapeutic cells. Biomacromolecules 14(2):322–330. https://doi.org/10.1021/
bm301690a
91. Hashemi M, Kalalinia F (2015) Application of encapsulation technology in stem cell therapy.
Life Sci 143:139–146. https://doi.org/10.1016/j.lfs.2015.11.007
92. Klinge PM, Harmening K, Miller MC et al (2011) Encapsulated native and glucagon-like pep-
tide-1 transfected human mesenchymal stem cells in a transgenic mouse model of Alzheimer’s
disease. Neurosci Lett 497(1):6–10. https://doi.org/10.1016/j.neulet.2011.03.092
93. Knippenberg S, Thau N, Dengler R et al (2012) Intracerebroventricular injection of encap-
sulated human mesenchymal cells producing glucagon-like peptide 1 prolongs survival in a
mouse model of ALS. PLoS One 7(6):e36857. https://doi.org/10.1371/journal.pone.0036857
94. Wright EJ, Farrell KA, Malik N et al (2012) Encapsulated glucagon-like peptide-1-producing
mesenchymal stem cells have a beneficial effect on failing pig hearts. Stem Cells Transl Med
1(10):759–769. https://doi.org/10.5966/sctm.2012-0064
95. Heile AMB, Wallrapp C, Klinge PM et al (2009) Cerebral transplantation of encapsulated
mesenchymal stem cells improves cellular pathology after experimental traumatic brain
injury. Neurosci Lett 463(3):176–181. https://doi.org/10.1016/j.neulet.2009.07.071
96 A. Gonzalez-Pujana et al.
96. Skinner SJM, Geaney MS, Lin H et al (2009) Encapsulated living choroid plexus cells:
potential long-term treatments for central nervous system disease and trauma. J Neural Eng
6:065001. https://doi.org/10.1088/1741-2560/6/6/065001
97. Huang SL, Wang J, He XJ et al (2014) Secretion of BDNF and GDNF from free and encap-
sulated choroid plexus epithelial cells. Neurosci Lett 566:42–45. https://doi.org/10.1016/j.
neulet.2014.02.017
98. Aliaghaei A, Digaleh H, Khodagholi F et al (2015) Encapsulated choroid plexus epithelial
cells actively protect against intrahippocampal abeta-induced long-term memory dysfunc-
tion; upregulation of effective neurogenesis with the abrogated apoptosis and neuroinflam-
mation. J Mol Neurosci 56(3):708–721. https://doi.org/10.1007/s12031-015-0492-y
99. Pettingill LN, Wise AK, Geaney MS et al (2011) Enhanced auditory neuron survival follow-
ing cell-based BDNF treatment in the deaf guinea pig. PLoS One 6(4):e18733. https://doi.
org/10.1371/journal.pone.0018733
100. Gillespie LN, Zanin MP, Shepherd RK (2015) Cell-based neurotrophin treatment supports
long-term auditory neuron survival in the deaf guinea pig. J Control Release 198:26–34.
https://doi.org/10.1016/j.jconrel.2014.11.026
101. Lindvall O, Wahlberg LU (2008) Encapsulated cell biodelivery of GDNF: a novel clini-
cal strategy for neuroprotection and neuroregeneration in Parkinson’s disease? Exp Neurol
209(1):82–88. https://doi.org/10.1016/j.expneurol.2007.08.019
102. Grandoso L, Ponce S, Manuel I et al (2007) Long-term survival of encapsulated GDNF
secreting cells implanted within the striatum of parkinsonized rats. Int J Pharm 343(1–2):69–
78. https://doi.org/10.1016/j.ijpharm.2007.05.027
103. Date I, Ohmoto T, Imaoka T et al (1996) Cografting with polymer-encapsulated human nerve
growth factor-secreting cells and chromaffin cell survival and behavioral recovery in hemipar-
kinsonian rats. J Neurosurg 84(6):1006–1012. https://doi.org/10.3171/jns.1996.84.6.1006
104. Date I, Ohmoto T, Imaoka T et al (1996) Chromaffin cell survival from both young and
old donors is enhanced by co-grafts of polymer-encapsulated human NGF-secreting cells.
Neuroreport 7(11):1813–1818
105. Emerich DF, Orive G, Thanos C et al (2014) Encapsulated cell therapy for neurodegenera-
tive diseases: from promise to product. Adv Drug Deliv Rev 67-68:131–141. https://doi.
org/10.1016/j.addr.2013.07.008
106. Spuch C, Antequera D, Portero A et al (2010) The effect of encapsulated VEGF-secreting
cells on brain amyloid load and behavioral impairment in a mouse model of Alzheimer’s
disease. Biomaterials 31(21):5608–5618. https://doi.org/10.1016/j.biomaterials.2010.03.042
107. Antequera D, Portero A, Bolos M et al (2012) Encapsulated VEGF-secreting cells enhance
proliferation of neuronal progenitors in the hippocampus of AbetaPP/Ps1 mice. J Alzheimers
Dis 29(1):187–200. https://doi.org/10.3233/JAD-2011-111646
108. Shen Y, Qiao H, Fan Q et al (2015) Potentiated osteoinductivity via cotransfection with
BMP-2 and VEGF genes in microencapsulated C2C12 Cells. Biomed Res Int 2015:435253.
https://doi.org/10.1155/2015/435253
109. Han YF, Han YQ, Pan YG et al (2010) Transplantation of microencapsulated cells express-
ing VEGF improves angiogenesis in implanted xenogeneic acellular dermis on wound.
Transplant Proc 42(5):1935–1943. https://doi.org/10.1016/j.transproceed.2009.12.070
110. Chen W, Yang D, Wang P et al (2011) Microencapsulated myoblasts transduced by the vas-
cular endothelial growth factor (VEGF) gene for the ischemic skin flap. Aesthet Plast Surg
35(3):326–332. https://doi.org/10.1007/s00266-010-9610-y
111. Kim C, Chung S, Yuchun L et al (2012) In vitro angiogenesis assay for the study of cell-
encapsulation therapy. Lab Chip 12(16):2942–2950. https://doi.org/10.1039/c2lc40182g
112. Selimoglu SM, Elibol M (2010) Alginate as an immobilization material for MAb produc-
tion via encapsulated hybridoma cells. Crit Rev Biotechnol 30(2):145–159. https://doi.
org/10.3109/07388550903451652
113. Dubrot J, Portero A, Orive G et al (2010) Delivery of immunostimulatory monoclonal anti-
bodies by encapsulated hybridoma cells. Cancer Immunol Immunother 59(11):1621–1631.
https://doi.org/10.1007/s00262-010-0888-z
3 Alginate Microcapsules for Drug Delivery 97
114. Saenz del Burgo L, Compte M, Aceves M et al (2015) Microencapsulation of therapeutic
bispecific antibodies producing cells: immunotherapeutic organoids for cancer management.
J Drug Target 23(2):170–179. https://doi.org/10.3109/1061186X.2014.971327
115. Read TA, Sorensen DR, Mahesparan R et al (2001) Local endostatin treatment of gliomas
administered by microencapsulated producer cells. Nat Biotechnol 19(1):29–34. https://doi.
org/10.1038/83471
116. Kleinschmidt K, Klinge PM, Stopa E et al (2011) Alginate encapsulated human mesenchymal
stem cells suppress syngeneic glioma growth in the immunocompetent rat. J Microencapsul
28(7):621–627. https://doi.org/10.3109/02652048.2011.599441
117. Goren A, Dahan N, Goren E et al (2010) Encapsulated human mesenchymal stem cells: a
unique hypoimmunogenic platform for long-term cellular therapy. FASEB J 24(1):22–31.
https://doi.org/10.1096/fj.09-131888
118. Shah K (2013) Encapsulated stem cells for cancer therapy. Biomatter 3(1.):Epub 2013 Jan 1).
https://doi.org/10.4161/biom.24278
119. Soon-Shiong P, Heintz RE, Merideth N et al (1994) Insulin independence in a type 1 diabetic
patient after encapsulated islet transplantation. Lancet 343(8903):950–951
120. Calafiore R, Basta G, Luca G et al (2006) Microencapsulated pancreatic islet allografts
into nonimmunosuppressed patients with type 1 diabetes: first two cases. Diabetes Care
29(1):137–138. https://doi.org/10.2337/diacare.29.01.06.dc05-1270
121. Basta G, Montanucci P, Luca G et al (2011) Long-term metabolic and immunological fol-
low-up of nonimmunosuppressed patients with type 1 diabetes treated with microencapsu-
lated islet allografts: four cases. Diabetes Care 34(11):2406–2409. https://doi.org/10.2337/
dc11-0731
122. Tuch BE, Keogh GW, Williams LJ et al (2009) Safety and viability of microencapsulated
human islets transplanted into diabetic humans. Diabetes Care 32(10):1887–1889. https://
doi.org/10.2337/dc09-0744
123. A phase I/II study of the safety and efficacy of Sernova’s Cell PouchTM for therapeutic islet
transplantation. Available via https://clinicaltrials.gov/ct2/show/NCT01652911. Accessed
Nov 2016
124. Elliott RB, Escobar L, Tan PL et al (2007) Live encapsulated porcine islets from a type 1
diabetic patient 9.5 yr after xenotransplantation. Xenotransplantation 14(2):157–161. https://
doi.org/10.1111/j.1399-3089.2007.00384.x
125. Living cell technologies: DIABECELL®. Available via http://www.lctglobal.com/products/
diabecell/development-to-date. Accessed Nov 2016
126. Open-label investigation of the safety and effectiveness of DIABECEL® in patients with type
I diabetes mellitus. Available via https://clinicaltrials.gov/ct2/show/NCT00940173. Accessed
Nov 2016
127. Open-label investigation of the safety and effectiveness of DIABECELL® in patients with
type 1 diabetes mellitus. Available via https://clinicaltrials.gov/ct2/show/NCT01739829.
Accessed Nov 2016
128. Open-label investigation if the safety and efficacy of DIABECELL® in patients with type 1
diabetes mellitus. Available via https://clinicaltrials.gov/ct2/show/NCT01736228. Accessed
Nov 2016
129. GLP-1 CellBeads® for the treatment of stroke patients with space-occupying intrecerebral
hemorrhage. Available via https://clinicaltrials.gov/ct2/show/NCT01298830. Accessed Nov
2016
130. Heile A, Brinker T (2011) Clinical translation of stem cell therapy in traumatic brain injury:
the potential of encapsulated mesenchymal cell biodelivery of glucagon-like peptide-1.
Dialogues Clin Neurosci 13(3):279–286
131. Luo XM, Lin H, Wang W et al (2013) Recovery of neurological functions in non-human
primate model of Parkinson’s disease by transplantation of encapsulated neonatal porcine
choroid plexus cells. J Parkinsons Dis 3(3):275–291. https://doi.org/10.3233/JPD-130214
98 A. Gonzalez-Pujana et al.
132. Open-label investigation of the safety and clinical effects of NTCELL® in patients with
Parkinson’s disease. Available via https://clinicaltrials.gov/ct2/show/NCT01734733.
Accessed Nov 2016
133. Investigation of the safety and efficacy of NTCELL® [Immunoprotected (Alginate-
Encapsulated) porcine choroid plexus cells for xenotransplantation] in patients with
Parkinson’s disease. Available via https://clinicaltrials.gov/ct2/show/NCT02683629.
Accessed Nov 2016
134. Living cell technoogies: NTCELL®. Available via http://www.lctglobal.com/products/ntcell/
development-to-date. Accessed Nov 2016
135. Santos E, Orive G, Hernandez RM et al (2011) Cell-biomaterial interaction: strategies to
mimic the extracellular matrix. In: Pramatarova L (ed) On biomimetics. InTech, Rijeka
136. Cruz-Acuna R, Garcia AJ (2016) Synthetic hydrogels mimicking basement membrane matri-
ces to promote cell-matrix interactions. Matrix Biol 57–58:324. https://doi.org/10.1016/j.
matbio.2016.06.002
137. Dalheim MO, Vanacker J, Najmi MA et al (2016) Efficient functionalization of alginate bio-
materials. Biomaterials 80:146–156. https://doi.org/10.1016/j.biomaterials.2015.11.043
138. Marturano JE, Schiele NR, Schiller ZA et al (2016) Embryonically inspired scaffolds regulate
tenogenically differentiating cells. J Biomech 49(14):3281–3288. https://doi.org/10.1016/j.
jbiomech.2016.08.011
139. Garate A, Ciriza J, Casado JG et al (2015) Assessment of the behavior of mesenchymal stem
cells immobilized in biomimetic alginate microcapsules. Mol Pharm 12(11):3953–3962.
https://doi.org/10.1021/acs.molpharmaceut.5b00419
140. Garate A, Santos E, Pedraz JL et al (2015) Evaluation of different RGD ligand densities in
the development of cell-based drug delivery systems. J Drug Target 23(9):806–812. https://
doi.org/10.3109/1061186X.2015.1020428
141. Santos E, Garate A, Pedraz JL et al (2014) The synergistic effects of the RGD density and the
microenvironment on the behavior of encapsulated cells: in vitro and in vivo direct compara-
tive study. J Biomed Mater Res A 102(11):3965–3972. https://doi.org/10.1002/jbm.a.35073
142. Llacua A, de Haan BJ, Smink SA et al (2016) Extracellular matrix components supporting
human islet function in alginate-based immunoprotective microcapsules for treatment of dia-
betes. J Biomed Mater Res A 104(7):1788–1796. https://doi.org/10.1002/jbm.a.35706
143. Mazzitelli S, Luca G, Mancuso F et al (2011) Production and characterization of engineered
alginate-based microparticles containing ECM powder for cell/tissue engineering applica-
tions. Acta Biomater 7(3):1050–1062. https://doi.org/10.1016/j.actbio.2010.10.005
144. Lou R, Xie H, Zheng H et al (2016) Alginate-based microcapsules with galactosylated chito-
san internal for primary hepatocyte applications. Int J Biol Macromol 93(Pt A):1133–1140.
https://doi.org/10.1016/j.ijbiomac.2016.09.078
145. Jeon O, Powell C, Solorio LD et al (2011) Affinity-based growth factor delivery using bio-
degradable, photocrosslinked heparin-alginate hydrogels. J Control Release 154(3):258–266.
https://doi.org/10.1016/j.jconrel.2011.06.027
146. Dang TT, Thai AV, Cohen J et al (2013) Enhanced function of immuno-isolated islets in diabe-
tes therapy by co-encapsulation with an anti-inflammatory drug. Biomaterials 34(23):5792–
5801. https://doi.org/10.1016/j.biomaterials.2013.04.016
147. Murua A, Herran E, Orive G et al (2011) Design of a composite drug delivery system to
prolong functionality of cell-based scaffolds. Int J Pharm 407(1–2):142–150. https://doi.
org/10.1016/j.ijpharm.2010.11.022
148. Acarregui A, Herran E, Igartua M et al (2014) Multifunctional hydrogel-based scaffold for
improving the functionality of encapsulated therapeutic cells and reducing inflammatory
response. Acta Biomater 10(10):4206–4216. https://doi.org/10.1016/j.actbio.2014.06.038
149. Azadi SA, Vasheghani-Farahani E, Hashemi-Najafbabadi S et al (2016) Co-encapsulation of
pancreatic islets and pentoxifylline in alginate-based microcapsules with enhanced immuno-
suppressive effects. Prog Biomater 5:101–109. https://doi.org/10.1007/s40204-016-0049-3
3 Alginate Microcapsules for Drug Delivery 99
150. Jo EH, Hwang YH, Lee DY (2015) Encapsulation of pancreatic islet with HMGB1 fragment
for attenuating inflammation. Biomater Res. 19:21-015-0042-2. eCollection 2015. https://
doi.org/10.1186/s40824-015-0042-2
151. Arlov O, Skjak-Braek G, Rokstad AM (2016) Sulfated alginate microspheres associate
with factor H and dampen the inflammatory cytokine response. Acta Biomater 42:180–188.
https://doi.org/10.1016/j.actbio.2016.06.015
152. Zanotti L, Sarukhan A, Dander E et al (2013) Encapsulated mesenchymal stem cells for
in vivo immunomodulation. Leukemia 27(2):500–503. https://doi.org/10.1038/leu.2012.202
153. Alunno A, Montanucci P, Bistoni O et al (2015) In vitro immunomodulatory effects of
microencapsulated umbilical cord Wharton jelly-derived mesenchymal stem cells in pri-
mary Sjogren’s syndrome. Rheumatology (Oxford) 54(1):163–168. https://doi.org/10.1093/
rheumatology/keu292
154. English K (2013) Mechanisms of mesenchymal stromal cell immunomodulation. Immunol
Cell Biol 91(1):19–26. https://doi.org/10.1038/icb.2012.56
155. Vaithilingam V, Evans MD, Rowe A et al (2016) Coencapsulation of target effector cells with
mesenchymal stem cells reduces pericapsular fibrosis and improves graft survival in a xeno-
transplanted animal model. Cell Transplant 25(7):1299–1317. https://doi.org/10.3727/09636
8915X688975
156. Stucky EC, Schloss RS, Yarmush ML et al (2015) Alginate micro-encapsulation of mesen-
chymal stromal cells enhances modulation of the neuro-inflammatory response. Cytotherapy
17(10):1353–1364. https://doi.org/10.1016/j.jcyt.2015.05.002
157. Luca G, Mancuso F, Calvitti M et al (2015) Long-term stability, functional competence,
and safety of microencapsulated specific pathogen-free neonatal porcine Sertoli cells: a
potential product for cell transplant therapy. Xenotransplantation 22(4):273–283. https://doi.
org/10.1111/xen.12175
158. Luca G, Bellezza I, Arato I et al (2016) Terapeutic potential of microencapsulated sertoli
cells in Huntington disease. CNS Neurosci Ther 22(8):686–690. https://doi.org/10.1111/
cns.12569
159. Arifin DR, Manek S, Call E et al (2012) Microcapsules with intrinsic barium radiopacity for
immunoprotection and X-ray/CT imaging of pancreatic islet cells. Biomaterials 33(18):4681–
4689. https://doi.org/10.1016/j.biomaterials.2012.03.008
160. Qie F, Astolfo A, Wickramaratna M et al (2015) Self-assembled gold coating enhances X-ray
imaging of alginate microcapsules. Nanoscale 7(6):2480–2488. https://doi.org/10.1039/
c4nr06692h
161. Yang F, Zhang X, Maiseyeu A et al (2012) The prolonged survival of fibroblasts with
forced lipid catabolism in visceral fat following encapsulation in alginate-poly-L-lysine.
Biomaterials 33(22):5638–5649. https://doi.org/10.1016/j.biomaterials.2012.04.035
162. Barnett BP, Arepally A, Stuber M et al (2011) Synthesis of magnetic resonance-, X-ray- and
ultrasound-visible alginate microcapsules for immunoisolation and noninvasive imaging of
cellular therapeutics. Nat Protoc 6(8):1142–1151. https://doi.org/10.1038/nprot.2011.352
163. Catena R, Santos E, Orive G et al (2010) Improvement of the monitoring and biosafety
of encapsulated cells using the SFGNESTGL triple reporter system. J Control Release
146(1):93–98. https://doi.org/10.1016/j.jconrel.2010.05.018
164. Allen AB, Gazit Z, Su S et al (2014) In vivo bioluminescent tracking of mesenchymal stem
cells within large hydrogel constructs. Tissue Eng Part C Methods 20(10):806–816. https://
doi.org/10.1089/ten.TEC.2013.0587
165. Santos E, Larzabal L, Calvo A et al (2013) Inactivation of encapsulated cells and their
therapeutic effects by means of TGL triple-fusion reporter/biosafety gene. Biomaterials
34(4):1442–1451. https://doi.org/10.1016/j.biomaterials.2012.10.076
166. Chan KW, Liu G, van Zijl PC et al (2014) Magnetization transfer contrast MRI for non-
invasive assessment of innate and adaptive immune responses against alginate-encapsulated
cells. Biomaterials 35(27):7811–7818. https://doi.org/10.1016/j.biomaterials.2014.05.057
100 A. Gonzalez-Pujana et al.
167. Krishnan R, Arora RP, Alexander M et al (2014) Noninvasive evaluation of the vascular
response to transplantation of alginate encapsulated islets using the dorsal skin-fold model.
Biomaterials 35(3):891–898. https://doi.org/10.1016/j.biomaterials.2013.10.012
168. Chang HK, Kim PH, Cho HM et al (2016) Inducible HGF-secreting human umbilical cord
blood-derived MSCs produced via TALEN-mediated genome editing promoted angiogen-
esis. Mol Ther 24(9):1644–1654. https://doi.org/10.1038/mt.2016.120
169. Wang P, Song Y, Weir MD et al (2016) A self-setting iPSMSC-alginate-calcium phosphate
paste for bone tissue engineering. Dent Mater 32(2):252–263. https://doi.org/10.1016/j.
dental.2015.11.019
170. Zhao L, Weir MD, HH X (2010) An injectable calcium phosphate-alginate hydrogel-umbilical
cord mesenchymal stem cell paste for bone tissue engineering. Biomaterials 31(25):6502–
6510. https://doi.org/10.1016/j.biomaterials.2010.05.017
171. Barkai U, Weir GC, Colton CK et al (2013) Enhanced oxygen supply improves islet viability
in a new bioartificial pancreas. Cell Transplant 22(8):1463–1476. https://doi.org/10.3727/09
6368912X657341
172. Neufeld T, Ludwig B, Barkai U et al (2013) The efficacy of an immunoisolating mem-
brane system for islet xenotransplantation in minipigs. PLoS One 8(8):e70150. https://doi.
org/10.1371/journal.pone.0070150
173. Veiseh O, Doloff JC, Ma M et al (2015) Size- and shape-dependent foreign body immune
response to materials implanted in rodents and non-human primates. Nat Mater 14(6):643–
651. https://doi.org/10.1038/nmat4290
Chapter 4
Alginate Processing Routes to Fabricate
Bioinspired Platforms for Tissue Engineering
and Drug Delivery
4.1 Introduction
Alginates are natural polysaccharides typically obtained from brown seaweed which
exhibit excellent biocompatibility and biodegradability that can be useful for many
applications in the field of biomedicine. They mainly work as ionic polymers
derived by the presence of divalent cations such as Ca2+, which confer them
Traditional methods for producing porous biopolymer scaffolds include gas foam-
ing, phase separation (freeze-drying), solvent casting, and particulate leaching [19].
In general, a polymer is first dissolved in a suitable solvent and subsequently placed
in a mold that will be rapidly cooled until the solvent freezes. The solvent is then
removed by freeze-drying, and pores will be left behind in the polymer. Different
types of phase separation techniques are available, including thermally induced,
solid-liquid, and liquid-liquid separation mechanisms [20–22].
Porous alginate-based scaffolds, foams, and sponges with interconnected porous
structures and predictable shapes can be easily manufactured by a regular thermally
induced phase separation (Fig. 4.1) [18]. When the temperature is low enough to
allow freezing the solution, the phase separation mechanism induces the solid-
liquid demixing, therefore forming frozen solvent and concentrated polymer phases.
By adjusting the polymer concentration or varying the cooling rate, phase separa-
tion could occur via different mechanisms, resulting in the formation of scaffolds
with various morphologies. “Freeze-drying” is one of the most extensively used
methods that produce matrices with porosity greater than 90%. The pore sizes
depend on the growth rate of ice crystals during the freeze-drying process. After the
Fig. 4.1 Alginate porous foams fabricated via freeze-drying technique and different bioactivation
procedures for improving cell recognition
4 Alginate Processing Routes to Fabricate Bioinspired Platforms for Tissue… 105
removal of the liquid or frozen solvent contained in the demixed solution, the space
originally occupied by the solvent would become pores in the prepared scaffolds.
Obviously, in the stage of solvent removal, the porous structure contained in the
solution needs to be carefully retained. Without freeze-drying, a rise in temperature
during the drying stage could result in remixing of the phase-separated solution or
remelting of the frozen solution, leading to destruction of the porous structure. In
the case of hydrogels, the freeze-dry processing does not require additional chemi-
cals, relying on the water already present in hydrogels to form ice crystals that can
be sublimated from the polymer, creating a particular micro-architecture. Because
the direction of growth and the size of ice crystals are a function of the temperature
gradient, linear, radial, and/or random pore directions and sizes can be produced
with this methodology [23]. The mechanical properties and biodegradation rate of
freeze-dried scaffolds can be simply modulated by changing the relative parameters
of the polymers. Porous scaffolds formed by pure alginate are unable to provide
enough bioactive properties to support cell metabolism due to the lack of cellular
interaction in the molecular structures [8, 18, 24, 25]. Therefore, alginate has been
blended with collagen and/or gelatin to enhance cell ligand-specific binding proper-
ties to fabricate hybrid scaffolds, which showed better properties for supporting
cells [18, 26, 27]. Recently, other efforts were made to enhance the biological prop-
erties of alginate porous scaffolds. In this way, alginate was irradiated and oxidized
to modify its degradation and covalently grafted with growth factors, lectins, and
peptides containing a RGD (arginylglycylaspartic acid) sequence to promote cell
adhesion and proliferation [18, 28–30]. As we have seen, the porosity and pore sizes
of the scaffolds fabricated through this method are largely dependent on the param-
eters such as the ratio of water to polymer solution and viscosity of the emulsion.
This technique also does not necessitate an extra leaching step, but the addition of
organic solvents such inhibits the incorporation of bioactive molecules or cells dur-
ing scaffold fabrication. In addition, the small pore sizes obtained are another limit-
ing factor of scaffolds fabricated by phase separation [22].
4.2.2 Atomization
e ntrapping matrix to fulfil the specific requirements. Alginate is by far the most
studied material for cell encapsulation, and it has been adopted for many biomedical
applications. This material has historically been used as a protective barrier to
enhance cell therapies, for immunoprotection of pancreatic islets, treatment of brain
tumors, treatment of anemia, and cryopreservation [47–49]. The main benefit of
using water-absorbable polymers for the preparation of these systems is not only to
develop a confined barrier to entrap living xenogeneic or allogeneic cells to be
transplanted but also to prohibit the entry of the host antibodies and immune cells.
Besides, encapsulated cells generally show limited interactions, and the physical
barrier concurs to mask them from the immune surveillance at a local level, thereby
assuring a reduction of the inflammatory response after transplantation [50].
Microcapsule and micro-carrier systems provide a larger surface area for cellular
attachment, provide cell protection against excessive mechanical stresses, and
simulate an in vivo environment [51, 52]. Recent improvements in fabrication
technologies allow generating matrices with different conformations – i.e., core
shell with liquid or hollow core and/or multicomponent coatings – which may
optimally be adapted to co-culture of single/multiple cell types for molecular guided
tissue engineering and “bio-organs” manufacturing [52, 53].
108 V. Guarino et al.
4.3.1 Emulsion
Alginate has many possible applications in the area of drug delivery due to their low
cost, low toxicity, biocompatibility, and biodegradability [61, 62]. Alginate micro-
spheres have been widely used as carriers for the controlled release of active agents
4 Alginate Processing Routes to Fabricate Bioinspired Platforms for Tissue… 109
due to their low immunogenicity and their muco-adhesive properties [63]. In addi-
tion, alginate matrices have the ability to encapsulate protein and DNA while main-
taining their biological activities and have shown very strong bioadhesive abilities
making alginate a promising candidate for site-specific mucosal delivery [64–66].
Moreover, the alginate polymer also finds application as a surfactant stabilizer in
oil-water emulsions in microbead preparation. Alginate surfactants serve to lower
the interfacial energy between the phases, thereby increasing the stability and life-
time of the emulsion. In fact, an emulsion is a thermodynamically unstable system
consisting of at least two immiscible liquid phases, one of which is dispersed as
globules in the other liquid phase, stabilized by the presence of an emulsifying
agent, as can be observed in Fig. 4.3. Emulsions fall into a greater class of two-phase
systems known as colloids, with the special characteristic that both the dispersed
and continuous phases are liquid. Depending on the volume fraction of the phases,
both water-in-oil and oil-in-water emulsions can be formed. There is a rule which
governs the emulsion formation, known as the Bancroft rule: emulsifiers and emul-
sifying tend to promote dispersion of the phase in which they do not dissolve well;
for example, proteins dissolve better in water than in oil and therefore tend to form
oil-in-water emulsions, promoting the dispersion of oil droplets throughout a con-
tinuous phase of water. Emulsions can be prepared through various methods of agi-
tation, as the two phases are immiscible and droplets will not form spontaneously,
such as sonication, which can produce droplets of 100–400nm. Alginate microbeads
for drug delivery can be easily prepared by emulsion-gelation technique. As reported
by Putta S.K. et al. [67], an example of alginate microbead preparation consists of
sodium alginate dissolution in purified water to form a homogeneous polymer solu-
tion. The drug is added to the solution and mixed thoroughly with a stirrer to form a
viscous dispersion. The dispersion is then added in an oil phase such as heavy liquid
paraffin while stirring at sufficiently high speed to emulsify the added dispersion as
fine droplets. Then appropriate amount of calcium chloride solution (15% w/v) is
transferred into the emulsion while stirring to complete the gelation reaction and to
Alginate beads
A. Uncoated alginate microspheres crosslinked in
calcium, barium
Polycation
bath (e.g. PLL, PLO)
B. Alginate coated by a polycation layer allows the fomation
of alginate beads
with immuno-
protective barrier
Fig. 4.4 Multilayered alginate microbeads consist of a perm-selective membrane around the algi-
nate core (a–b) followed by an outer layer alginate (c). The outer layer alginate allows for encap-
sulation of proteins and growth factors while also maintaining biocompatibility
4.4 A
pplications in Regenerative Medicine and Industrial
Pharmacology
tides in 3D versus 2D culture. The cells actively reorganized on the nanoscale the
adhesion ligands presented from the gels. Alginate gels have also been used to
examine how a 3D culture microenvironment influences cancer cell signaling and
tumor vascularization [53, 89–92]. Alginate has also been combined with inorganic
materials to enhance bone tissue formation. Alginate/hydroxyapatite (HAP) com-
posite scaffolds with interconnected porous structures were prepared by several
methods to enhance the adhesion of bone cells. Also cell-encapsulating alginate gel
beads were introduced into calcium phosphate cement and demonstrated potential
for bone tissue engineering under moderate stress-bearing conditions [53, 93–96].
In this context, alginate has demonstrated great potential for the fabrication of
micro-sized devices for many applications in drug delivery, tissue engineering, and
cancer therapy. Most attractive features of alginate in biomedical field include bio-
compatibility, mild gelation conditions, and simple modifications to prepare alginate
derivatives with new properties. Alginate has a track record of safe clinical uses as a
wound healing dressing material and pharmaceutical component; it has been safely
implanted in a variety of applications, including islet cell transplantation [50, 53]. As
one looks to the future, the alginate-based materials used in medicine are likely to
evolve considerably. In wound healing, and more generally drug delivery applica-
tions, precise control over the delivery of single vs. multiple drugs or sustained vs.
sequential release in response to external environmental changes is highly desirable.
Dynamical control over delivery can potentially improve the safety and effectiveness
of drugs, providing new therapies. On-demand drug release from alginate gels in
response to external cues such as mechanical signals and magnetic fields can be used
to design active depots of many drugs, including therapeutic cells. The introduction
of appropriate cell-interactive features to alginate will also be crucial in many tissue
engineering applications. The type of adhesion ligands and their spatial organization
in gels are key variables, as they can regulate cell phenotype and the resultant func-
tion of regenerated tissues. Further, current understating of fundamental properties
of alginate and developing of new types of cell and tissue-interactive alginate gels
may enable future advances in biomedical science and engineering [53].
tural properties and novel functions have already been prepared by this approach
and explored for biomedical applications [99]. The ability to chemically or physi-
cally engineer new classes of alginates with precisely controlled functionalities,
unlike the limited repertoire available from natural sources, designed for a specific
application, paves toward an enormous revolution for the future use of alginate in
biomedical field.
Acknowledgments This study was financially supported by the Ministero dell’Universita’ e della
Ricerca through the funds of POLIFARMA (PON02 3203241) and the National Operative Program
REPAIR (PON01-02342).
References
1. Gombotz WR, Wee SF (1998) Protein release from alginate matrices. Adv Drug Deliv Rev
31:267–285
2. Rinaudo M (2008) Main properties and current applications of some polysaccharides as bio-
materials. Polym Int 57:397–430
3. Fischer FG, Dörfel H (1955) Die polyuronsauren der braunalgen–(kohlenhydrate der algen–I).
Z Physiol Chem 302:186–203
4. George M, Abraham TE (2006) Polyionic hydrocolloids for the intestinal delivery of protein
drugs. J Control Release 114:1–14
5. SE S–E, Hubbell JA (2001) Functional biomaterials: design of novel biomaterials. Ann Rev
Mater Res 31:183–201
6. Augst AD, Kong HJ, Mooney DJ (2006) Alginate hydrogels as biomaterials. Macromol Biosci
6:623–633
7. Kuo CK, Ma PX (2001) Ionically crosslinked alginate hydrogels as scaffolds for tissue engi-
neering: part 1. Structure, gelation rate and mechanical properties. Biomaterials 22:511–521
8. Florczyk SJ, Kim DJ, Wood DL et al (2011) Influence of processing parameters on pore struc-
ture of 3D porous chitosan–alginate polyelectrolyte complex scaffolds. J Biomed Mater Res A
98:614–620
9. Kong HJ, Smith MK, Mooney DJ (2003) Designing alginate hydrogels to maintain viability of
immobilized cells. Biomaterials 24:4023–4029
10. Hay ID, Rehman ZU, Ghafoor A et al (2010) Bacterial biosynthesis of alginates. J Chem
Technol Biotechnol 85:752–759
11. Varghese S, Elisseeff JH (2006) Hydrogels for musculoskeletal tissue engineering. Adv Polym
Sci 203:95–144
12. Guarino V, D’Albore M, Altobelli R (2016) Polymer bioprocessing to fabricate 3D scaffolds
for tissue engineering. Int Polym Process. https://doi.org/10.3139/217.3239
13. Hollister SJ (2009) Scaffold design and manufacturing: from concept to clinic. Adv Mater
21:3330–3342
14. Manferdini C, Guarino V, Zini N (2010) Mineralization occurs faster on a new biomimetic
hyaluronic acid–based scaffold. Biomaterials 14:3986–3996
15. Chen CY, Ke CJ, Yen KC et al (2015) 3D porous calcium–alginate scaffolds cell culture sys-
tem improved human osteoblast cell clusters for cell therapy. Theranostics 5:643–655
16. Rodríguez–Vázquez M, Vega–Ruiz B, Ramos–Zúñiga R et al (2015) Chitosan and its potential
use as a scaffold for tissue engineering in regenerative medicine. Biomed Res Int 2015:821279
17. Guarino V, Lewandowska M, Bil M, Polak B, Ambrosio L (2010) Morphology and degra-
dation properties of pcl/hyaff11-based composite scaffolds with multiscale degradation rate.
Comp Sc Tech 70(2010):1826–1837
4 Alginate Processing Routes to Fabricate Bioinspired Platforms for Tissue… 117
18. Novotna K, Havelka P, Sopuch T et al (2013) Cellulose–based materials as scaffolds for tissue
engineering. Cellulose 20:2263–2278
19. Su J, Tan H (2013) Alginate–based biomaterials for regenerative medicine applications.
Materials 6:1285–1309
20. Nam YS, Guarino V, Causa F, Salerno A, Ambrosio L, Netti PA (2008) Design and manufac-
ture of microporous polymeric materials with hierarchal complex structure for biomedical
application. Mat Sci Tech 24(9):1111–1117
21. Schugens C, Maquet V, Grandfils C et al (1996) Polylactide macroporous biodegradable
implants for cell transplantation. II. Preparation of polylactide foams by liquid-liquid phase
separation. J Biomed Mater Res A 30:449–461
22. Loh QL, Choong C (2013) Three–dimensional scaffolds for tissue engineering applications:
role of porosity and pore size. Tissue Eng Part B Rev 19:485–502
23. Ikada Y (2011) Tissue engineering: fundamentals and applications Vol. 8 in Interface Science
and Technology. Academic Press, Cambridge, Elsevier, Uk
24. Yuan NY, Lin YA, Ho MH et al (2009) Effects of the cooling mode on the structure and
strength of porous scaffolds made of chitosan, alginate, and carboxymethyl cellulose by the
freeze–gelation method. Carbohydr Polym 78:349–356
25. Sapir Y, Kryukov O, Cohen S (2011) Integration of multiple cell–matrix interactions into algi-
nate scaffolds for promoting cardiac tissue regeneration. Biomaterials 3:1838–1847
26. Kim G, Ahn S, Kim Y et al (2011) Coaxial structured collagen–alginate scaffolds: fabrication,
physical properties, and biomedical application for skin tissue regeneration. J Mater Chem
21:6165–6172
27. Petrenko YA, Ivanov RV, Petrenko AY et al (2011) Coupling of gelatin to inner surfaces of pore
walls in spongy alginate–based scaffolds facilitates the adhesion, growth and differentiation of
human bone marrow mesenchymal stromal cells. J Mater Sci Mater Med 22:1529–1540
28. Han J, Zhou Z, Yin R et al (2010) Alginate–chitosan/hydroxyapatite polyelectrolyte complex
porous scaffolds: preparation and characterization. Int J Biol Macromol 46:199–205
29. Shachar M, Tsur–Gang O, Dvir T et al (2011) The effect of immobilized RGD peptide in algi-
nate scaffolds on cardiac tissue engineering. Acta Biomater 7:152–162
30. Zhou H, HH X (2011) The fast release of stem cells from alginate–fibrin microbeads in inject-
able scaffolds for bone tissue engineering. Biomaterials 32:7503–7513
31. Guarino V, Cirillo V, Ambrosio L (2016) Bicomponent electrospun scaffolds to design ECM
tissue analogues. Exp Rev Med Dev 13:83–102
32. Guarino V, Cirillo V, Altobelli R et al (2015) Polymer based platforms by electric field assisted
techniques for tissue engineering and cancer therapy. Exp Rev Med Dev 12:113–129
33. Oliveira MB, Mano JF (2011) Polymer-based microparticles in tissue engineering and regen-
erative medicine. Biotechnol Prog 27:897–912
34. Luciani A, Guarino V, Ambrosio L, Netti A Solvent and melting induced microsphere sinter-
ing techniques: a comparative study of morphology and mechanical properties. J Mater Sci
Mater Med 22:2019–2028
35. Altobelli R, Guarino V, Ambrosio L (2016) Micro- and nanocarriers by electrofludodynamic
technologies for cell and molecular therapies. Process Biochem https://doi.org/10.1016/j.
procbio.2016.09.002
36. Tabata Y, Horiguchi I, Lutolf MP et al (2014) Development of bioactive hydrogels capsules for
the 3D expansion of pluripotent stem cells in bioreactors. Biomater Sci 2:176–183
37. De Koker S, Richard H, de Geest BG (2012) Polymeric multilayer capsules for drug delivery.
Chem Soc Rev 41:2867–2884
38. Guarino V, Gloria A, Raucci MG et al (2012) Hydrogel–based platforms for the regeneration
of osteochondral tissue and intervertebral disc. Polymers 4:1590–1612
39. Guarino V, Galizia M, al APMA (2015) Improving surface and transport properties of macro-
porous hydrogels for bone regeneration. J Biomed Mat Res A 103:1095–1105
118 V. Guarino et al.
40. Leferink A, Schipper D, Arts E et al (2014) Engineered micro-objects as scaffolding ele-
ments in cellular building blocks for bottom-up tissue engineering approaches. Adv Mat
26:2592–2599
41. Kelm JM, Djonov V, Ittner LM et al (2006) Design of custom-shaped vascularized tissues
using microtissue spheroids as minimal building units. Tissue Eng 12:2151–2160
42. Rivron NC, Rouwkema J, Truckenmüller R et al (2009) Tissue assembly and organization:
developmental mechanisms in microfabricated tissues. Biomaterials 30:4851–4858
43. Causa F, Netti PA, Ambrosio L (2007) A multi–functional scaffold for tissue regeneration: the
need to engineer a tissue analogue. Biomaterials 28:5093–5099
44. Fukui Y, Maruyama T, Iwamatsu Y et al (2010) Preparation of monodispersed polyelectrolyte
microcapsules with high encapsulation efficiency by an electrospray technique. Colloids Surf
A Physicochem Eng Asp 370:28–34
45. Pankongadisak P, Ruktanonchai UR, Supaphol P et al (2014) Preparation and characterization
of silver nanoparticles–loaded calcium alginate beads embedded in gelatin scaffolds. AAPS
PharmSciTech 15:1105–1115
46. Kang AR, Park JS, Ju J et al (2014) Cell encapsulation via microtechnologies. Biomaterials
35:2651–2663
47. Gasperini L, Mano JF, Reis RL (2014) Natural polymers for the microencapsulation of cells.
J R Soc Interface 11:20140817
48. Hunt NC, Grover LM (2010) Cell encapsulation using biopolymer gels for regenerative medi-
cine. Biotechnol Lett 32:733–742
49. Nicodemus GD, Bryant SJ (2008) Cell encapsulation in biodegradable hydrogels for tissue
engineering applications. Tissue Eng Part B Rev 14:149–165
50. Leung A, Nielsen LK, Trau M et al (2010) Tissue transplantation by stealth coherent alginate
microcapsules for immunoisolation. Biochem Eng J 48:337–347
51. Malda J, Frondoza CG (2006) Microcarriers in the engineering of cartilage and bone. Trends
Biotechnol 24:299–304
52. Naqvi SM, Vedicherla S, Gansau J et al (2016) Living cell factories-electrosprayed microcap-
sules and microcarriers for minimally invasive delivery. Adv Mater 28:5662–5671
53. Lee KY, Mooney DJ (2012) Alginate: properties and biomedical applications. Prog Polym Sci
37:106–126
54. Wu H, Liao C, Jiao Q, Wang Z, Cheng W, Wan Y (2012) Fabrication of core-shell micro-
spheres using alginate and chitosan-polycaprolactone for controlled release of vascular endo-
thelial growth factor. React Funct Polym 72:427–437
55. Vasir JK, Tambwekar K, Garg S (2003) Bioadhesive microspheres as a controlled drug deliv-
ery systems. Int J Pharm 255:13–32
56. Tanaka H, Matsumura M, Veliky IA (1984) Diffusion characteristics of substrates in
Ca-alginate gel beads. Biotechnol Bioeng 26(1):53–58. PubMed PMID:18551586
57. Smith TJ (1994) Calcium algiante hydrogel as a matrix for enteric delivery of nucleic acids.
Biopharm 4:54–55
58. Mestecky J (1987) The common mucosal immune system and current strategies for induction
of immune responses in external secretions. J Clin Immunol 7:265–276
59. Putta S, Kumar A, Kumar A (2010) Formulation and in-vitro evaluation of mucoadhesive
microcapsules of glipizide with gum kondagogu. J Chem Pharm Res 2(5):356–364
60. Sailaja R, Amareshwar P, Chakravarty P (2010) Chitosan nanoparticles as a drug delivery
systems. J Pharm Biol Chem Sci Res 1:476
61. Jaiswal D, Bhattacharya A, Yadav I, Singh H, Chandra D, Jain D (2009) Formulation and
evaluation of oil entrapped floating alginate beads of ranitidine hydrochloride. Int J Pharm
Pharm Sci 1:129–141
62. Singhal P, Kumar K, Pandey M, Shubhini A (2010) Evaluation of acyclovir loaded oil
entrapped calcium alginate beads prepared by ionotropic gelation method. Int J ChemTech
Res 2:2076–2085
4 Alginate Processing Routes to Fabricate Bioinspired Platforms for Tissue… 119
84. Huebsch N, Kearney CJ, Zhao X et al (2014) Ultrasound–triggered disruption and self–healing
of reversibly cross–linked hydrogels for drug delivery and enhanced chemotherapy. Proc Natl
Acad Sci U S A 111:9762–9767
85. Wang Y, Zhou J, Qiu L et al (2014) Cisplatin–alginate conjugate liposomes for targeted deliv-
ery to EGFR–positive ovarian cancer cells. Biomaterials 35:4297–4309
86. Boekhoven J, Zha RH, Tantakitti F et al (2015) Alginate–peptide amphiphile core–shell mic-
roparticles as a targeted drug delivery system. RSC Adv 5:8753–8756
87. Wu JL, Wang CQ, Zhuo RX (2014) Multi–drug delivery system based on alginate/calcium
carbonate hybrid nanoparticles for combination chemotherapy. Colloid Surf B 123:498–505
88. Lucinda–Silva RM, Salgado HRN, Evangelista RC (2010) Alginate–chitosan systems: in vitro
controlled release of triamcinolone and in vivo gastrointestinal transit. Carbohydr Polym
81:260–268
89. Jeon O, Powell C, Ahmed SM et al (2010) Biodegradable, photocrosslinked alginate hydro-
gels with independently tailorable physical properties and cell adhesivity. Tissue Eng Part A
16:2915–2925
90. Degala S, Zipfel WR, Bonassar LJ (2011) Chondrocyte calcium signaling in response to
fluid flow is regulated by matrix adhesion in 3–D alginate scaffolds. Arch Biochem Biophys
505:112–117
91. Lee JW, Park YJ, Lee SJ et al (2010) The effect of spacer arm length of an adhesion ligand
coupled to an alginate gel on the control of fibroblast phenotype. Biomaterials 31:5545–5551
92. Huebsch N, Arany PR, Mao AS et al (2010) Harnessing traction–mediated manipulation of the
cell/matrix interface to control stem–cell fate. Nat Mater 9:518–526
93. Jin HH, Kim DH, Kim TW et al (2012) In vitro evaluation of porous hydroxyapatite/chitosan–
alginate composite scaffolds for bone tissue engineering. Int J Biol Macromol 51:1079–1085
94. Rubert M, Monjo M, Lyngstadaas SP et al (2012) Effect of alginate hydrogel containing poly-
proline–rich peptides on osteoblast differentiation. Biomed Mater 7:055003
95. Florczyk SJ, Leung M, Jana S et al (2012) Enhanced bone tissue formation by alginate gel–
assisted cell seeding in porous ceramic scaffolds and sustained release of growth factor.
J Biomed Mater Res A 100:3408–3415
96. Tang M, Chen W, Weir MD et al (2012) Human embryonic stem cell encapsulation in algi-
nate microbeads in macroporous calcium phosphate cement for bone tissue engineering. Acta
Biomater 8:3436–3445
97. Lee KY, Peters MC, Anderson KW et al (2000) Controlled growth factor release from syn-
thetic extracellular matrices. Nature 408:998–1000
98. Zhao XH, Kim J, Cezar CA et al (2011) Active scaffolds for on–demand drug and cell delivery.
Proc Natl Acad Sci U S A 108:67–72
99. MacKay JA, Chen MN, McDaniel JR et al (2009) Self–assembling chimeric polypeptide–
doxorubicin conjugate nanoparticles that abolish tumours after a single injection. Nat Mater
8:993–999
Chapter 5
Alginate Utilization in Tissue Engineering
and Cell Therapy
Bapi Sarker and Aldo R. Boccaccini
B. Sarker
Department of Mechanical Engineering & Materials Science, Washington University
in St. Louis, St. Louis, MO, USA
A.R. Boccaccini (*)
Institute of Biomaterials, Department of Materials Science and Engineering,
University of Erlangen-Nuremberg, Erlangen, Germany
e-mail: aldo.boccaccini@ww.uni-erlangen.de
5.2 A
lginate in Tissue Engineering: Opportunities
and Limitations
Though alginate has a lot of advantages, it possesses some major limitations for the
applications in tissue engineering, which will be discussed in this section. Alginate
hydrogels do not degrade but rather disintegrate when the coordinated divalent
cations are replaced by monovalent cations present in the surrounding fluids. The
interactions between the alginate chains and the monovalent cations lead to the dissolu-
tion of the gels. The unbounded polymer chains cannot be degraded by the biologi-
cal activity of the mammalian hosts. Even if the gel dissolves in the physiological
environment of mammals, the molecules of alginate cannot be completely removed
from the body since the average molecular weights of commercially available algi-
nates are higher than the renal clearance threshold of the kidneys [4]. However,
alginates can be degraded by the specific enzymes, alginate lyases also known as
5 Alginate Utilization in Tissue Engineering and Cell Therapy 123
alginases, which are not present in mammals [5]. In addition, another drawback of
alginate is its lack of cell adhesion motifs and the resulting failure of cell attachment
leading to very poor cell-material interactions both in 2D and 3D environments [6,
7]. Moreover, alginate hydrogels promote minimal protein adsorption due to its
hydrophilic nature. Consequently, mammalian cells are unable to interact with the
hydrogel through serum proteins [8]. Cell anchorage or cell-material interaction is
the key factor for cell survival in 2D and 3D cultures and orchestrates most of the
cellular functions including migration, proliferation, differentiation, and apoptosis
[6, 9]. Moreover, a high mechanical stress is exerted on the cells embedded in algi-
nate hydrogel that impedes elongation and migration of active viable cells.
Therefore, since alginate hydrogels do not promote cell adhesion and migration, the
cells cultured on the surface (2D) or inside (3D) of alginate hydrogel form multicel-
lular aggregates [10]. The drawbacks of alginate as an emerging biopolymer for
biomedical applications can be overcome by some approaches, such as by enhanc-
ing its degradation through chemical modification, like partial oxidation [2, 11] or
gamma irradiation [12], and by introducing cell-binding motifs, like RGD (Arg-
Gly-Asp)-containing peptides or proteins, which can be conjugated with alginate to
promote cell adhesion [6]. However, designing an alginate-based hydrogel system
which can support cell anchorage and promote typical cellular functions including
elongation, migration, proliferation, and differentiation in 3D environments suitable
for tissue engineering applications remains challenging.
These limitations of alginate can be overcome by incorporation of gelatin through
covalently crosslinking with alginate dialdehyde (ADA) [13, 14]. ADA is a partially
oxidized product of alginate which facilitates the covalent crosslinking with gelatin
through the Schiff’s base formation due to the reaction of free amino groups of lysine
or hydroxylysine amino acid residues of gelatin and available aldehyde groups of
ADA [13, 15]. The partial oxidation cleaves the carbon-carbon bond of the cis-diol
group in the uronate residue of alginate and alters the chair conformation to an
open-chain adduct, which facilitates degradation of the alginate [16]. Moreover, the
biodegradability of the covalently crosslinked hydrogel can be tuned by using ADA
of different degrees of oxidation which can control the hydrolysis property of
alginate [16, 17] and also by changing the ratio of ADA and gelatin [18, 19].
culture research, where 3D cell culture systems are replacing 2D cell cultures. In
vitro 3D cell culture models span the gap between conventional in vitro 2D cell
culture models and in vivo animal models [20, 21]. Alginate-based 3D cell cultures
are performed in multiple ways. Among them, (1) encapsulation of cells in alginate
beads, (2) fabrication of scaffolds by plotting cell-loaded hydrogel precursors with
subsequent gelation of the precursors, and (3) seeding cells in prefabricated porous
scaffolds are the most used techniques. Alginate has been used in advanced research
in various tissue engineering application fields, e.g., bone, cartilage, dental, cardiac,
muscle, adipose, vascular, neural, and retinal tissue engineering.
Since the last decades, porous biomaterials-based constructs with a high degree of
porosity and an interconnected pore structure are considered for tissue engineering
scaffolds. Several techniques including gas foaming, solvent casting and salt leaching,
freeze-drying, electrospinning, and rapid prototyping (bioprinting) are applied to
fabricate 3D hydrogel-based polymeric scaffolds [22]. Among the conventional
techniques, freeze-drying is widely used for fabrication of tissue engineering
scaffolds from hydrogel-based materials, like alginate. Porous solid free-form
scaffolds can be fabricated from an alginate solution by simple two steps: freezing
followed by freeze-drying, as shown in Fig. 5.1. The method is based on rapid cool-
ing of the material under freezing temperature to generate thermal instabilities
within the system which cause phase separation and sublimation of the solvent
under vacuum resulting in voids in the space it previously occupied [23]. Among the
various relevant processing parameters, freezing temperature, freezing rate, and
freezing process before the lyophilization have a major impact on the pore structure
(overall porosity, pore size, and pore morphology) of the resulting scaffolds
[23–25]. Generally, the porosity and pore size of freeze-dried scaffolds increase
with increasing freezing temperature. This phenomenon can be attributed to the
formation of ice crystals which are larger in size and lower in number at higher
freezing temperatures than at lower freezing temperatures. During the freeze-drying
Fig. 5.1 Schematic diagram showing the fabrication technique of porous alginate-based scaffolds
using the freeze-drying method (Reproduced with permission from Ref. [27]. Copyright 2013,
MDPI publishing)
5 Alginate Utilization in Tissue Engineering and Cell Therapy 125
process, the larger ice crystals push and expand the biopolymer to a greater extent
leading to the formation of highly porous scaffolds with large pore sizes [26].
Scaffolds with open-pore structures are generated at a high freezing temperature
(between −20 °C and −80 °C), whereas scaffolds with parallel sheet-like morphol-
ogy are obtained at a very low freezing temperature (−196 °C in liquid N2) [23].
This technique is generally used to fabricate porous scaffolds with intercon-
nected pores from ionically crosslinked alginate hydrogels [24, 27, 28]. However, it
is well known that the scaffolds from pure alginate do not provide enough biocom-
patibility to support cell adhesion and cell metabolism due to lack of cell adhesion
motifs in the alginate network [27, 29]. Therefore, in order to promote cell adhesion
and cell metabolism, different polymers, proteins, or peptides are combined with
alginate, such as chitosan [30, 31], gelatin [32], especially peptides having sequences
like G4SPPRRARVTY, G4SPPLLALVTY, G4RGDY, etc., which are known to pro-
mote cell adhesion [29]. Alginate hydrogel-based porous freeze-dried scaffolds
have been mostly used for bone and cartilage tissue engineering research, which
will be discussed in the next paragraphs.
Sustained release of BFP-1 from the scaffolds was observed that promoted the
osteogenic differentiation of osteoblasts. Among the three concentrations of BFP-1,
optimal results regarding cell viability and osteogenic activity were obtained for the
hydrogel composing 10 μg/mL peptide. Furthermore, the incorporated peptide sig-
nificantly enhanced osteo-regeneration in vivo when the scaffolds were implanted
into Beagle calvarial defects. Osteoid, the organic portion of the bone matrix that is
composed of collagen-I and chondroitin sulfate, deposited in the calvarial defects
where BFP-1-containing alginate scaffolds were implanted. In another study, car-
bodiimide chemistry was used to crosslink gelatin with alginate, in which stable
amide bonds formed between alginate and gelatin in the presence of EDC and NHS
through the three steps of reaction [34]. Microwave-vacuum drying technique was
used to fabricate porous scaffolds from low and high crosslinked alginate-gelatin
hydrogels. In both conditions, in vitro and in vivo, the scaffolds exhibited excellent
cytocompatibility, biocompatibility, and bioresorbability. It is important to note that
a strong bond was observed between implanted alginate-gelatin scaffolds with
upper subcutaneous tissue and lower muscle tissue at the implanted site of mice.
Moreover, angiogenesis and neovascularization were observed and that generally
help to supply nutrient to the implanted cells, which enhance tissue regeneration
process. However, the degree of osteogenic, adipogenic, and chondrogenic differen-
tiations of mesenchymal stem cells (MSCs) was observed to be lower for subcuta-
neously implanted scaffolds in mice compared to in vitro culture, which shows that
the in vitro differentiation potential of MSCs does not always correlate with in vivo
tissue regeneration. Not only proteins or peptides but also other polymers are used
with alginate to design BTE scaffolds. Among the naturally derived polymers,
chitosan is the most studied material with alginate because these two polysaccha-
rides are ionically opposite in nature and therefore polyelectrolyte complex forms in
their mixture that can minimize the drawbacks of both materials. Since alginate is a
negatively charged polymer, no protein adsorption usually takes place on it. Due to
the inclusion of positively charged polymer, chitosan, protein can be adsorbed on
the polymer composite and that can improve cell adhesion [30]. Li et al. showed
alginate-chitosan hybrid freeze-dried porous scaffolds supported growth and prolif-
eration of MG63 osteoblast-like cells in vitro [31]. Moreover, calcium and
phosphate-rich mineral deposition was observed on osteoblasts grown on the hybrid
scaffolds, and this type of deposition was not observed on pristine chitosan scaf-
folds. Mineral deposition in large areas was also confirmed on the bone marrow-
infused hybrid scaffolds in vivo at 4 and 8 weeks after implantation into the muscle
of rats. In another study, undifferentiated rat bone marrow-derived MSCs were used
to investigate comparative osteogenesis in the defect sites of calvaria of Sprague
Dawley rats, which were treated separately with freeze-dried chitosan-alginate
(CA) scaffolds and the scaffolds with seeded MSCs, bone marrow (BM) aspirate,
and bone morphogenetic protein-2 (BMP2) growth factor [35]. Cells and growth
factor were mixed with 0.5% (w/v) alginate solution prior to loading into the porous
scaffolds. The cell- or growth factor-immobilized alginate solution was ionically
crosslinked after loading. Among the four groups of constructs, the BMP2-loaded
scaffolds exhibited the greatest defect closure (71.6 ± 19.7%) after 16 weeks of
implantation. Mature lamellar bone and the greatest amount of mineralization with
5 Alginate Utilization in Tissue Engineering and Cell Therapy 127
blasts for the application in BTE. It has also been proven that the TCP-incorporated
oxidized alginate-gelatin hydrogels can be used as a potential drug delivery carrier.
Moreover, the encapsulated osteoblasts within the TCP-incorporated oxidized
alginate-gelatin hydrogels exhibited high ALP activity, proving that the encapsu-
lated cells maintained their osteoblastic nature. Recently, Sarker et al. designed
hydrogel-based freeze-dried scaffolds composed of oxidized alginate, gelatin, and
BG for bone tissue engineering applications as shown in Fig. 5.2 [41]. The presence
of BG (45S5) enhanced the crosslinking kinetic and crosslinking degree of the oxi-
dized alginate-gelatin (ADA-GEL)-based hydrogel. The enhanced crosslinked
structure of hydrogels ultimately contributed to the mechanical properties of the
hydrogel-derived scaffolds. High BG content enhanced the crosslinking degree of
the hydrogel significantly due to the alkalinity of the reaction medium, which facili-
tated Schiff’s base bond formation between free amino groups of gelatin and avail-
able aldehyde groups of oxidized alginate. Highly crosslinked hydrogel networks
enhanced mechanical properties of the scaffolds. Degradation property of the scaf-
folds was also successfully tailored by changing the BG content in the ADA-GEL
hydrogel. Moreover, the scaffolds composed of high BG exhibited low protein
release profile since the high BG-containing ADA-GEL hydrogel possessed high
crosslinked structure that inhibited protein release and hydrolytic degradation. Bone
marrow-derived stromal cells were cultured in the scaffolds, and the cell growth was
found to be promoted in ADA-GEL scaffolds and 1% BG-containing ADA-GEL
scaffolds compared to pure alginate scaffolds and 5% BG-containing ADA-GEL
scaffolds. Low cell viability in 5% BG-containing ADA-GEL scaffolds could be
due to the possible cytotoxic effect of a high concentration of released ions from BG
particles. Scaffolds with optimum BG content (1%), adequate protein, and con-
trolled degradation supported better cell growth, proliferation, migration, and osteo-
genic differentiation. Moreover, the presence of BG facilitated HAp deposition onto
the scaffolds that made the constructs osteoconductive.
with required degradation properties. The same strategy was also applied to design
the scaffolds with optimal shape-memory properties that enable in vivo scaffold
implantation via a minimally invasive manner. Moreover, the molar mass and partial
oxidation of alginate are the tuning tools of mechanical strength of scaffolds, which
controls the mechanotransduction of cell that controls myoblast functionality such
as adhesion, proliferation, and differentiation. Partially oxidized alginate has also
been used to design suitable exogenous matrix for liver tissue engineering applica-
tions [49]. In this study, the partially oxidized alginate was covalently crosslinked
with galactosylated chitosan via Schiff’s base reaction since hepatocyte surface
receptors, asialoglycoprotein (ASGP), recognize galactose moieties, which pro-
motes hepatocytes adhesion. Due to the presence of galactose, the water solubility
of chitosan increases, which is a major limitation of using galactosylated chitosan
alone in liver tissue engineering applications. High water solubility of galactosyl-
ated chitosan was reduced and controlled by crosslinking with partially oxidized
alginate, a biocompatible material. Moreover, the mechanical strength and degrada-
tion properties of the covalently crosslinked freeze-dried scaffolds were easily tuned
by changing the ratio of oxidized alginate and galactosylated chitosan, which
changed the degree of crosslinking of the synthesized hydrogel. The covalently
crosslinked composite scaffolds supported hepatocytes adhesion and growth.
Moreover, hepatocytes exhibited typical spheroidal morphology and formed multi-
cellular aggregates, which prove the alginate-based composite matrix is a favorable
niche for the growth of hepatocytes.
One of the major challenges in tissue engineering is insufficient nutrient supply
to the distant cells in the tissue engineering constructs due to the lack of vascularity.
Freeze-dried porous alginate scaffolds have been employed to induce vasculariza-
tion via physical stimulation to endothelial cells using magnetic nanoparticles [50].
The aortic endothelial cells in magnetite-impregnated alginate scaffolds were stim-
ulated with an external magnetic field during the first 7 days of culture, which
exhibited high metabolic activity. Most importantly, a capillary-like organization of
endothelial cells was observed in the magnetite-impregnated alginate scaffolds that
were exposed to an alternating magnetic field. It is believed that the mechanical
stimulation with magnetic field enhances the organization of endothelial cells and
eventually the capillary-like structures form.
the native biological tissue. The hydrogel must be biocompatible and non-immuno-
genic and support appropriate cellular activities, such as migration, proliferation,
and differentiation of embedded cells. The hydrogels with suitable mechanical
properties and swelling and degradation characteristics are desirable properties for
printing tissue engineering scaffolds to promote cell proliferation, migration, and
other important cellular functions. Among the various naturally derived hydrogel-
forming materials, alginate is the most used material for printing tissue engineering
scaffolds because of its biocompatibility, non-immunogenic property, and most
importantly excellent ionic gelation property.
exhibited good processability for plotting scaffolds with tailored architecture using
a bioplotter. MBG with a pore size of 5 nm was mixed with alginate at various con-
centrations and added into polyvinyl alcohol (PVA) solution to prepare the ink for
plotting scaffolds. The scaffolds were plotted in layer-by-layer deposition of the
prepared ink and the models of the scaffolds were designed by CAD. The composite
ink was extruded through a nozzle by a specific dosing air pressure at a constant
plotting speed at room temperature. The strut width and pore size of the scaffolds
can be tailored by changing the inner diameter of nozzle, dosing pressure, and plot-
ting speed. The pore structure and pore size can also be tailored by optimizing the
plotting pattern. In the study of Luo et al., two different plotting patterns, named XY
and XXYY, were employed to design the scaffolds with different pore sizes and
pore structures. For XY pattern, the first layer was plotted in the X direction, and the
second layer was plotted in the Y direction, and this plotting fashion was repeated
until the whole scaffold was completed. For XXYY pattern, the ink was first plotted
in the X direction for two layers, and the next two layers were plotted in the Y direc-
tion, then repeated until the scaffold was completed. The scaffolds that were plotted
in XXYY pattern exhibited significantly improved pore interconnectivity compared
to the scaffolds plotted in XY direction. Mechanical properties of the scaffolds were
modulated by changing the amount of the MBG content in the plotting ink. Due to
incorporation of 30% and 50% MBG into the alginate ink, mechanical strength and
modulus of the plotted scaffolds have increased significantly. Apatite mineralization
was observed on the scaffolds that contain MBG after soaking in SBF, which confirms
the bioactivity of the scaffolds. It is well established that apatite mineralization
enhances the cell-material and tissue-implant interactions and improves osteoblas-
tic activity, which was also observed in the study of Luo et al. Improved cell-mate-
rial interaction with high ALP activity of hBMSCs was found on the mineralized
MBG/alginate scaffolds compared to the pristine alginate scaffolds. Due to miner-
alization, protein binding was improved and bioactive silicate ions released from
MBG, which might be the possible reason of high cell-material interactions and
improved osteoblastic activity of hBMSCs. The designed MBG/alginate scaffolds
were also used to deliver drug, where Dexamethasone was used as the model drug,
which is generally used for treatment of rheumatoid arthritis. Much controlled and
sustained release of the drug was found when the drug was loaded to MBG particles
prior to fabrication of MBG/alginate scaffolds compared to the pure alginate scaf-
folds. The release kinetic of the drug can be further controlled by changing MBG
content in the plotting ink, which is very important for biomedical applications. In
another study, the effect of BG on growth and mineralization of human osteogenic
sarcoma cells, SaOS-2 cells, immobilized into a printable alginate-gelatin hydrogel,
was investigated [54]. Furthermore, the hydrogel was supplemented either with
polyphosphate (as polyP·Ca2+ complex) or silica, or as biosilica that was enzymati-
cally prepared from ortho-silicate by silicatein. The cell-embedded bio-ink was
loaded into the printing cartridge, connected with a needle, and mounted in the
preheated printing head of a 3D bioplotter. The scaffolds were plotted in a meander-
like pattern by changing the directions of the consecutive layers. The strut width of
the plotted scaffolds was 300 μm, and the struts were arranged in a perpendicular
orientation. In the presence of BG nanoparticles with a size of 55 nm and a molar
134 B. Sarker and A.R. Boccaccini
which was developed by Sarker et al. [18], was used to fabricate scaffolds using a 3D
bioplotting technique [56]. In this study oxidized alginate-gelatin covalently cross-
linked hydrogel was used in which osteoblast-like MG63 cells were immobilized
prior to plotting. Over the incubation period, an increasing trend in metabolic activity
of embedded MG63 cells in the plotted scaffolds was observed. Moreover, the release
of vascular endothelial growth factor (VEGF) was observed, and the release trend
was found to be increased over the incubation period. The upregulation of VEGF
expression of embedded osteoblast-like cells in ADA-GEL matrix proved the ongo-
ing angiogenesis, which is very important for all types of tissue engineering includ-
ing bone tissue engineering. The embedded cells were observed to grow out of the
ADA-GEL hydrogel strut and covered the whole scaffold. This result confirms the
migration of the embedded cells in the hydrogel matrix, which occurred due to the
high matrix porosity, low stiffness, and high degradability of the ADA-GEL hydrogel
as stated elsewhere [57–59]. Moreover, cells exhibited spreading morphology with
high cell-material interaction, which has never been observed for alginate-based bio-
plotted scaffolds. The outcomes demonstrated that ADA-GEL hydrogel is a superior
material in the context of bioplotting compared to pure alginate hydrogel. However,
designing scaffolds with tailored height in Z direction was found to be very challeng-
ing to achieve with the ADA-GEL hydrogel, which is the bottleneck of this matrix
system for biofabrication in the field of tissue engineering. In another study, stron-
tium-doped bioactive glass nanoparticles (BGNPs) were incorporated into ADA-
GEL hydrogel to fabricate bioplotted scaffolds with improved bioactivity for bone
tissue engineering applications [60]. Growth of bone-like apatite layer on the surface
of the plotted scaffolds occurred when the constructs were immersed in SBF. The
embedded MG63 cells exhibited high viability, and no difference was found in cell
viability between pristine ADA-GEL scaffolds and the BGNP-incorporated ADA-
GEL constructs, proving that the addition of BGNPs did not affect cell viability.
Alginate-based hydrogels are widely used as therapeutic materials for cartilage tis-
sue regeneration [61]. However, the major challenge of using alginate hydrogels in
biofabrication technique to design 3D scaffold is their inability to maintain a uni-
form 3D structure. To overcome this problem, alginate hydrogel was integrated with
a synthetic polymer, polycaprolactone (PCL) [62]. In this study, additive manufac-
turing (AM) technique with a multihead deposition system (MHDS) was used to
fabricate 3D scaffolds by plotting primary nasal septal cartilage chondrocytes and/
or TGF-β-embedded alginate hydrogel in between the plotted PCL struts using a
layer-by-layer plotting approach. Using the MHDS, multiple biomaterial inks can
be dispensed to plot complex 3D scaffolds, which can closely mimic our tissue
structure. In vitro cell studies showed that cell viability was reduced to about 85%
when multiple PCL layers were plotted compared to the single-layer constructs,
where cell viability was found to be 95–97%. Though the cell viability reduced due
to plotting more PCL matrix in the constructs, the reduction was not significantly
136 B. Sarker and A.R. Boccaccini
low, which suggested that the shear stress due to dispensing cell-containing alginate
matrix did not have a significant impact on cell viability. In vitro biochemical assay
indicated that more glycosaminoglycan (GAG) and total collagen were expressed
by the cells in the constructs having TGF-β-containing alginate. Moreover, more
cartilaginous tissue formation was observed for the constructs having 4% alginate
gels compared to the 6% alginate gels. For in vivo study, three different types of
scaffolds were printed and that were composed of PCL+ alginate gel (no cells)
named as group I;, PCL+ alginate gel (chondrocytes), group II; and PCL+ alginate
gel (chondrocytes + TGF-β), group III, which were implanted into the dorsal subcu-
taneous spaces of a 7-week-old female nude mice. After 4 weeks, accumulation of
more GAG and formation of cartilaginous tissue and type II collagen were observed
in the constructs of group III compared to that in the constructs of other two groups.
Type II collagen exhibits a fibrillar structure of healthy cartilage tissue that main-
tains the mesh structure of cartilage and takes water-retentive proteoglycan into its
pores. In order to recapitulate the nanofibrous matrix constitution of the native mus-
culoskeletal soft tissue, a fibrous bio-ink composed of alginate and polylactic acid
(PLA) nanofibers was used to print tissue engineering constructs [63]. Human
adipose-derived stem cells (hADSCs) containing alginate-PLA nanofibers bio-ink
was plotted to fabricate the construct that can mimic the structure of human medial
knee meniscus, which was digitally modeled using magnetic resonance imaging.
Viability study showed that the PLA nanofibers-containing alginate constructs pro-
moted higher cell viability and proliferation compared to the constructs having algi-
nate only. At day 7, metabolic activity of the hADSCs was found to be 28.5% higher
in the nanofiber-containing constructs compared to the constructs without nanofi-
ber. Most importantly, collagen and proteoglycans were found to be prominent in
the areas surrounding the hADSCs in the nanofiber-containing constructs which
confirmed the ability of bioprinted hMSC to differentiate down the chondrogenic
pathway. In another study, nanofibrillated cellulose (NFC) was mixed with alginate
to design a printable bio-ink by taking the advantage of shear-thinning properties of
NFC [64]. The low zero-shear viscosity of alginate makes the pristine alginate ink a
poor shape fidelity when printing. To improve the shape fidelity of the alginate bio-
ink, NFC was used with alginate in the bio-ink formulation, which combines the
high rheological properties of NFC and the good ionic gelation ability of alginate.
Human nasoseptal chondrocytes (hNC) were used for 3D bioprinting of gridded
constructs using the NFC-alginate bio-ink for cartilage tissue engineering. The
cytotoxicity and live/dead assays showed no potential detrimental effect of the bio-
ink on embedded cell viability. These preliminary results demonstrate that NFC-
alginate bio-ink is a biocompatible hydrogel well suited for 3D bioprinting for
designing cartilage tissue engineering constructs.
Highly organized alginate-based 3D scaffolds were developed using a microfluidic
technique [65]. The microfluidic device that has been used is a two-channel fluid
jacket microencapsulator for bubble formation equipped with a micropipette (inner
diameter 45 μm and outer diameter 95 μm). A 1.5% alginate solution with 1%
Pluronic® F127 (surfactant) was fed into the outer channel, and nitrogen gas was fed
into the inner channel as shown in Fig. 5.3a. Hollow alginate microcapsules were
Fig. 5.3 (a) The microfluidic device and the scheme of the dispenser tip, in which the surfactant-
containing alginate solution and nitrogen gas were fed separately to generate (b1) bubble-like
alginate microcapsules, which formed (b2) a honeycomb structure after gelation. (b3 and b4) A 3D
ordered array structure with high and precise organization is shown in the confocal microscopy
images. (b5) SEM image of the 3D alginate scaffold with a highly interconnected porous structure,
which was achieved after vacuum degassing (Reproduced with permission from Ref. [65].
Copyright 2011, Elsevier)
138 B. Sarker and A.R. Boccaccini
generated at a specific flow rate and gas pressure. The microcapsules were kept in a
gelation bath with 2% calcium chloride solution to facilitate ionic gelation. The micro-
capsules formed a honeycomb structure (Fig. 5.3b2) when the array was adequate
with the high organization. After proper gelation, the structure was put in the vacuum
system overnight for removing air bubbles and fabricating the highly organized porous
scaffold with high interconnected porosity as shown in Fig. 5.3b5. The pore size of the
scaffolds is influenced by the viscosity of alginate solution, nitrogen gas pressure, and
the inner diameter of the micropipette. Inconsistence in pore size of the scaffolds fab-
ricated by other conventional methods (e.g., freeze-drying) is a major drawback,
which is overcome by this technique. The seeded porcine chondrocytes in the scaf-
folds showed excellent viability and proliferation, revealing that the highly porous and
organized scaffold has good biocompatibility. GAG content was found to be increased
over the culture time. In addition, the expression of collagen type II and aggrecan
increased. On the other side, a decreasing trend for collagen type I and X was observed,
which proved that the chondrocytes maintained their normal phenotype without dif-
ferentiating toward osteogenesis in the alginate scaffolds.
Apart from the specific tissue engineering purpose, various biofabrication
strategies have been utilized to design biocompatible alginate-based scaffolds,
which show excellent viability of the embedded cells. Park et al. used the combi-
nation of low molecular weight (LMW) alginate and high molecular weight
(HMW) alginate at various ratios as the bio-ink to investigate the effect of the
composition of alginate bio-ink on the printability of the scaffolds and on the
embedded cell viability. As shown in Fig. 5.4, alginate bio-ink composed of LMW
and HMW alginate in the ratio of 1:2 with a concentration of 3 w% showed the
best performance, including good printability and a suitable environment for cell
growth and proliferation. In another study, a cell-laden alginate hydrogel-based
3D constructs were designed that contain hollow calcium alginate filaments,
which were fabricated by using a coaxial nozzle [67]. Using this bioprinting
method, the hollow calcium alginate filament-based scaffolds were fabricated by
controlling the crosslinking time to allow fusion of adjacent hollow filament as
shown in Fig. 5.5. Porous 3D hydrogel constructs with various sizes and shapes
and with tunable mechanical properties can be fabricated using this approach. The
viability of L929 mouse fibroblasts in the constructs with built-in microchannels
was found to be higher than that in alginate structures without microchannels,
which proves that the microchannels in the constructs can facilitate transporting
of nutrients and oxygen to the embedded cells.
Fig. 5.4 (continued) with concentration of 3%), Ink 5 (LMW alginate: HMW alginate 1:2 with
concentration of 3%), and Ink 8 (HMW alginate with concentration of 3%) before and after ionic
crosslinking with calcium ion, and thick 3D porous constructs are fabricated using the three inks.
(c) Fluorescence live/dead images of cells printed in the alginate bio-inks (Ink 2, Ink 5, and Ink 8)
after 3 h and 7 days of culture, where living and dead cells are appeared in green and red, respec-
tively (scale bar: 500 μm) (Reproduced with permission from Ref. [66]. Copyright 2017, Elsevier)
Fig. 5.4 (a) Schematic presentation of 3D bioprinting process using the bio-inks with vari-
ous ratios of LMW alginate and HMW alginate. The fibroblasts-embedded scaffolds were
plotted in a layer-by-layer printing sequence, followed by subsequent gelation with calcium ions.
(b) Photographs of the 3D porous scaffolds printed with Ink 2 (LMW alginate: HMW alginate 1:1
140 B. Sarker and A.R. Boccaccini
Fig. 5.5 (a) Schematic showing the fabrication process of a 3D alginate structure with built-in
microchannels that was fabricated using a coaxial nozzle-assisted 3D bioprinting system. (b)
Printed 3D alginate porous structures with built-in microchannels: (a) hollow cylinder, (b) grid, (c)
cuboid, and (d) hemispheroid. (c) SEM image of the alginate hollow filaments in a printed cuboid
structure consisting of six layers of hollow filament (Reproduced with permission from Ref. [67].
Copyright 2015, Elsevier)
5.3.3 Microspheres
release system of growth factors and drugs [68]. In this section, applications of cell
encapsulation will be highlighted. Cell encapsulation technique aims to embed via-
ble and functional cells within a biocompatible matrix in order to provide the
embedded cells with a 3D tissue-like environment, mimicking in vivo condition. In
addition to biocompatibility, a suitable matrix for cell encapsulation must possess
semipermeability that allows the exchange of oxygen, nutrients, and metabolites
and simultaneously protects the encapsulated cells from the toxic foreign bodies [2,
69]. Due to their ability to protect the encapsulated cells from antibodies and the
host immune system, this technique has drawn attention in clinical applications as a
delivery system enabling the transport of specific cells to the target site in vivo.
Designing matrix for encapsulation of cell or tissue is the major challenge in this
field. Initially, this approach was proposed as a mean to protect the encapsulated
cells from the external environment, thereby preventing the rejection by the host
immune system [70]. For the applications in tissue engineering, this property is not
only the major requirement. The permeability of smaller molecules like oxygen,
nutrients, growth factors, and metabolites are equally important for cell survival
inside the matrix, and these properties must be possessed by ideal microsphere used
for tissue engineering applications, especially for cell encapsulation. Furthermore,
equilibrated mass transfer through the whole microcapsule is very important. If the
dimension of the pores and the porosity of matrix are not sufficient for the diffusion
through the matrix, the cells that are far from the surface of the microsphere may
receive a lower amount of nutrients at a given time [2]. Moreover, the matrix of the
microspheres must be biodegradable to provide enough space for migration and
proliferation of cells, leading to build the ECM [71]. Another important factor is the
mechanical stiffness of the matrix that should be optimized to protect the encapsu-
lated cells from external stress. However, the matrix should not exert high mechani-
cal stress to the embedded cells, which can inhibit normal cellular behavior [72].
Hydrogels are commonly used as matrices for cell encapsulation since they pro-
vide a number of features which are advantageous for the biocompatibility. Among
the various synthetic and naturally derived hydrogels, alginate is widely used for the
encapsulation of cells and biomolecules due to its excellent ionic gelation properties
with divalent cations which occurs under mild, nontoxic conditions for the encapsu-
lated cells. Microencapsulation of cells using alginate has been originally done by
Lim and Sun [73]. Cell encapsulation strategy has been used mostly for the applica-
tions in bone, cartilage, vascular, muscle, adipose, neural, and other soft tissue engi-
neering, which will be discussed in the next paragraphs.
modification. In one of the studies, murine-derived adipose tissue stromal cells have
been encapsulated to investigate growth and osteogenic differentiation in compari-
son to the conventional 2D culture [74]. Cells were found to be grown in clusters
with rounded and cuboidal morphology in alginate microspheres during osteoin-
duction whether cells exhibited elongated fibroblastic morphology in 2D culture.
Compared to 2D culture, encapsulated cells exhibited significant osteogenic activ-
ity, which was revealed by a high expression of ALP and osteocalcin mRNA, and
prove that alginate hydrogel provides a niche for osteogenic differentiation of stem
cells.
Local microenvironment, like the presence of cytokines, immune cells, and
growth factors, controls the stem cell-mediated bone tissue regeneration. Among
the various growth factors, bone morphogenetic protein-2 (BMP2) is often used to
induce osteogenic differentiation of stem cells [75]. As an alternative to using
recombinant BMP2, an anti-BMP2 monoclonal antibody (mAb) was used in RGD-
conjugated alginate microspheres to design an appropriate microenvironment for
osteogenic differentiation of human bone marrow mesenchymal stem cells (hBM-
SCs). Anti-BMP2 mAb can trap endogenous BMP ligands, which activate BMP
receptors of hBMSCs and thus induce osteogenic differentiation of the stem cells
[75, 76]. An in vivo study was conducted using this approach, in which hBMMSCs-
encapsulated anti-BMP2 mAb-preloaded RGD-alginate microspheres were
implanted in a calvarial defect of immunocompromised mice. A significant amount
of bone repair in the presence of hBMSCs and anti-BMP2 mAb was confirmed by
micro-CT and histological analyses compared to the groups with hBMSCs or anti-
BMP2 mAb alone or even recombinant human BMP2 (rhBMP2) (Fig. 5.6). In addi-
tion, in vitro model was used to investigate the molecular mechanism governing the
osteogenic differentiation of hBMSCs, which also confirmed that the osteogenic
differentiation is modulated by the BMP signaling pathway through capturing
BMP2 ligands by the incorporated anti-BMP2 mAb. The use of RGD-conjugated
alginate as an exogenous ECM matrix makes the system simpler with it and can be
easily modified to provide a perfect 3D niche to the encapsulated hBMSCs.
Efficient mass transfer throughout the microsphere is very challenging in conven-
tional static in vitro culture. To facilitate nutrient and osteogenic supplement supply
throughout the microsphere for enhancing viability, proliferation, and osteogenic
differentiation of embryonic stem cells (ESCs), a study was performed using an
efficient rotary cell culture microgravity bioreactor integrated with a cell encapsula-
tion unit [77]. High viability and osteogenic differentiation of ESCs and mineraliza-
tion were observed throughout the microspheres. In bone tissue engineering, mineral
deposition is specifically controlled by osteoblasts or differentiated osteogenic cells
rather than nonspecific mineral precipitation. This phenomenon can be confirmed
by the presence of collagen type I and osteocalcin within the mineralized constructs
since type I collagen is the most abundant protein and osteocalcin is the most abun-
dant non-collagenous protein in bone. The presence of collagen type I and osteocal-
cin in the cell-encapsulated microspheres was confirmed by immunocytochemistry.
Moreover, the composition of the deposited minerals throughout the constructs is
very important to understand the nature of the mineralization, which is generally
5 Alginate Utilization in Tissue Engineering and Cell Therapy 143
Fig. 5.6 Contribution of anti-BMP2 mAb to bone regeneration in a critical size mouse calvarial
defect, where hBMSC-encapsulated RGD-alginate microspheres loaded with or without anti-
BMP2 mAb or Iso mAb were implanted. (a) Micro-CT results of bone repair in mouse calvarial
defects, where regenerated bone is denoted with pseudo-colored red. (b) Characterization of the
hBMSCs after transplantation: cells and alginate positive for BMP2 epitopes are stained brown
(middle panel), and cells expressing the bone-associated transcription factor Runx2 are stained
brown (black arrows) shown in the bottom panel. (c) Semiquantitative analysis of bone formation
based on micro-CT images. (d) Histomorphometric analysis of calvarial defects showing the rela-
tive amount of bone formation in the critical size calvarial defects. *p < 0.05, **p < 0.01
(Reproduced with permission from Ref. [75]. Copyright 2013, Elsevier)
Fig. 5.7 Viability and morphology of encapsulated osteoblast-like MG63 cells in the microbeads
of various alginate-based hydrogels. (a) Bright field (top row) and confocal microscopy (bottom
row) images of cell-loaded microbeads made of pure alginate (ALG), physically blended alginate-
gelatin (ALG-GEL-b), and chemically crosslinked oxidized alginate-gelatin (ADA-GEL-x) hydro-
gels after 28 days of incubation. Cells were stained for the nucleus (green) and F-actin (red). (b)
The mitochondrial activity of the encapsulated cells over the incubation period showing higher cell
activity in the chemically crosslinked composition compared to pure alginate and blended alginate-
gelatin hydrogel. Asterisks denote significant difference, *p < 0.05, **p < 0.01, and ***p < 0.001
(Bonferroni’s post hoc test was used) (Reproduced with permission from Ref. [59]. Copyright
2015, Elsevier)
146 B. Sarker and A.R. Boccaccini
hydrogel [57]. Apart from gelatin, other proteins, e.g., keratin [83], silk fibroin [84],
fibrin [85], have been used with alginate hydrogel to encapsulate cells, which proves
that alginate is a suitable material that can be modified chemically or with different
proteins and peptides to promote cellular functions in 3D. Utilizing degradation
characteristic of oxidized alginate, a fast cell release study was designed by encap-
sulating human umbilical cord mesenchymal stem cells (hUCMSCs) into oxidized
alginate-fibrin microbeads [85]. A fast degradation of oxidized alginate- fibrin
microbeads was observed, and that facilitated the fast release of hUCMSCs, which
showed osteogenic differentiation with elevated bone marker gene expressions of
ALP, osteocalcin (OC), collagen type I, and Runx2. Moreover, an increasing trend
of hUCMSC-synthesized bone mineral formation over the culturing period was
confirmed by Alizarin Red S (ARS) staining.
As an inorganic phase, BG was introduced into alginate to generate composite
hydrogel microspheres as cell career for bone tissue regeneration [86]. As stated
earlier due to the presence of BG, apatite phase deposition was induced on the sur-
face of the microspheres after being soaked in SBF. Higher proliferation and stimu-
lated osteogenic differentiation of encapsulated preosteoblastic cell line, MC3T3-E1,
were observed in BG-incorporated alginate microspheres compared to the pristine
alginate microspheres. Some studies revealed that the released ions from BG, espe-
cially silicon and calcium, change the local environment and stimulate osteogenesis
and angiogenesis in vitro and in vivo [87, 88].
Cell encapsulation is a very useful strategy whereby living cells are entrapped
within biocompatible, biomimetic hydrogel-based matrix, such as alginate. Cell
encapsulation within alginate microbeads has been extensively used for cartilage
regeneration using various cell types, e.g., chondrocytes, stem cells, and progenitor
cells. In order to effectively differentiate stem cells and progenitor cells into chon-
drocytes, an appropriate microenvironment and signaling molecules are required in
alginate-based hydrogel systems. It is well known that growth factors such as TGF-
β1, BMP-4, and FGF-2 are often used to induce chondrogenesis. Moshaverinia
et al. [89] developed alginate hydrogel-based co-delivery system, in which dental
MSCs such as periodontal ligament stem cells (PDLSCs) and gingival mesenchy-
mal stem cells (GMSCs) were encapsulated within TGF-β1-loaded RGD-conjugated
alginate hydrogel. Comparative chondrogenic differential potentiality of the two
dental MSCs was investigated over bone marrow MSCs (BMMSCs) in this hydro-
gel system. Successful chondrogenesis of encapsulated PDLSCs and GMSCs was
observed in vitro and in vivo, confirmed by histochemical analysis and immuno-
fluorescence staining. Moreover, the two major matrices of articular cartilage, pro-
teoglycan and type II collagen, were produced for the groups having PDLSCs and
GMSCs, and the synthesis of the two matrices was found to be induced in the pres-
ence of TGF-β1. More chondrogenesis was observed for PDLSCs than BMMSCs.
The results showed that the presence of TGF-β1 and RGD peptides in alginate
5 Alginate Utilization in Tissue Engineering and Cell Therapy 147
Though the microencapsulation of cells within alginate hydrogel is mostly used for
regeneration of bone and cartilage, a significant research has been carried out for
regeneration of soft tissues, like muscle, nerve, ovarian follicle, cardiac, and vascu-
lar tissues [82, 96–99]. One of the major challenges in the tissue regeneration using
stem cells is the design of a suitable microenvironment, which can support cell
attachment and promote proliferation and differentiation of stem cells into the
required lineage. Differentiation of stem cells is generally controlled by the ECM
microenvironment, including the presence of cell adhesion ligands, growth factors,
cytokines, etc. [96]. Moreover, mechanical cues of ECM also direct stem cell dif-
ferentiation lineage [100]. Ansari et al. used RGD-conjugated alginate hydrogel to
design niche for growth and differentiation of encapsulated GMSCs [96]. This
microsphere system has been used to deliver multiple growth factors, (Forskolin
(FSK), 6-Bromo-1-methylindirubin-3′-oxime (MeBIO), and basic fibroblast growth
factor (bFGF)) to the encapsulated GMSCs to differentiate to myogenic lineage
in vitro and in vivo, where GMSC-encapsulated RGD-coupled alginate micro-
spheres were transplanted subcutaneously into immunocompromised mice. The
presence of the myogenic growth factors doubled the expression of myogenic-
specific genes (MyoG, MyoD, and Myf5) compared to non-encapsulated GMSCs.
The porous microstructure of alginate hydrogel synergistically facilitates cell
growth and differentiation due to the availability of nutrients and growth factors. It
has also been revealed that the mechanical properties of RGD-conjugated alginate
hydrogel control the fate of encapsulated GMSCs. Highest myogenic differentiation
of GMSCs was observed within the hydrogel with an intermediate modulus of elas-
ticity (10–16 kPa) in comparison to softer (<5 kPa) or stiffer (>20 kPa) hydrogels.
It is interesting to note that greater myogenic differentiation was observed for
GMSCs compared to the positive control, hBMMSCs. Moreover, the implanted
GMSCs encapsulated in RGD-conjugated alginate microspheres loaded with mul-
tiple myogenic growth factors showed increased neovascularization and local
angiogenesis compared to the microspheres containing hBMMSCs. Taking into
account the outcomes of the study, it can be concluded that GMSCs have better
muscle tissue regeneration capacity compared to hBMMSCs in an appropriate
5 Alginate Utilization in Tissue Engineering and Cell Therapy 149
environment with suitable inducing signals. GMSCs and other dental-derived stem
cells, PDLSCs, have also been used for tendon tissue regeneration, where a co-
delivery system based on TGF-β3-loaded RGD-conjugated alginate microspheres
was developed for encapsulation of the two stem cells along with hBMMSCs as a
positive control [101]. Since collagen bundle structure in the periodontal ligament
is similar to that of tendon tissue, it has been hypothesized that PDLSCs could be a
suitable stem cell for tendon regeneration. All three stem cells encapsulated in
RGD-coupled alginate exhibited high levels of mRNA expression for gene markers,
Scx, DCn, Tnmd, and Bgy, which are related to tendon regeneration. However, it is
interesting to note that expression levels of the gene markers were found to be
higher for PDLSCs compared to the other two cell types. In the in vivo study, where
the MSCs-encapsulated TGF-β3-loaded RGD-coupled alginate microspheres
were transplanted subcutaneously in mice, histological and immunohistochemical
staining confirmed the regeneration of ectopic neo-tendon tissue in the implanted
constructs. Similar to the in vitro outcomes, significantly higher tendon tissue
regeneration was observed for PDLSC-encapsulated microspheres compared to the
microspheres containing GMSCs or hBMMSCs in the in vivo study. The unique
porous structural property of alginate and the presence of conjugated RGD tripep-
tide and suitable signal molecule (TGF-β3) facilitated cell-matrix interactions,
leading to enhanced MSC adhesion, proliferation, and differentiation to a specific
lineage. The porous structural properties, ease of conjugation or loading of peptides,
proteins, and growth factors, and non-cytotoxicity make alginate hydrogel a highly
suitable biomaterial for tissue engineering. Cell encapsulation strategy was also
used to grow and expand neural stem cells (NSCs), where alginate-gelatin micro-
beads were designed with tuned internal pore size and structure, which provided a
suitable 3D environment supporting NSC proliferation [82].
5.4 Conclusions
References
1. Lee KY, Mooney DJ (2001) Hydrogels for tissue engineering. Chem Rev 101:1869–1880.
https://doi.org/10.1021/cr000108x
2. Gasperini L, Mano JF, Reis RL (2014) Natural polymers for the microencapsulation of cells.
J R Soc Interface 11:20140817. https://doi.org/10.1098/rsif.2014.0817
3. Cukierman E, Pankov R, Yamada KM (2002) Cell interactions with three-dimensional matri-
ces. Curr Opin Cell Biol 14:633–640. https://doi.org/10.1016/S0955-0674(02)00364-2
4. Al-shamkhani A, Duncan R (1995) Radioiodination of alginate via covalently-bound tyros-
inamide allows monitoring of its fate in vivo. J Bioact Compat Polym 10:4–13. https://doi.
org/10.1177/088391159501000102
5. Wong TY, Preston LA, Schiller NL (2000) Alginate lyase: review of major sources and
enzyme characteristics, structure-function analysis, biological roles, and applications. Annu
Rev Microbiol 54:289–340. https://doi.org/10.1146/annurev.micro.54.1.289
6. Rowley JA, Madlambayan G, Mooney DJ (1999) Alginate hydrogels as synthetic extracel-
lular matrix materials. Biomaterials 20:45–53
7. Boontheekul T, Kong H-J, Mooney DJ (2005) Controlling alginate gel degradation utilizing
partial oxidation and bimodal molecular weight distribution. Biomaterials 26:2455–2465.
https://doi.org/10.1016/j.biomaterials.2004.06.044
8. Smetana K (1993) Cell biology of hydrogels. Biomaterials 14:1046–1050. https://doi.
org/10.1016/0142-9612(93)90203-E
9. Price LS (1997) Morphological control of cell growth and viability. Bioessays 19:941–943.
https://doi.org/10.1002/bies.950191102
10. Kim D, Monaco E, Maki A et al (2010) Morphologic and transcriptomic comparison of adi-
pose- and bone-marrow-derived porcine stem cells cultured in alginate hydrogels. Cell Tissue
Res 341:359–370. https://doi.org/10.1007/s00441-010-1015-3
11. Lee KY, Mooney DJ (2012) Alginate: properties and biomedical applications. Prog Polym
Sci 37:106–126. https://doi.org/10.1016/j.progpolymsci.2011.06.003
5 Alginate Utilization in Tissue Engineering and Cell Therapy 151
12. Kong HJ, Kaigler D, Kim K, Mooney DJ (2004) Controlling rigidity and degradation of algi-
nate hydrogels via molecular weight distribution. Biomacromolecules 5:1720–1727. https://
doi.org/10.1021/bm049879r
13. Balakrishnan B, Jayakrishnan A (2005) Self-cross-linking biopolymers as injectable in situ
forming biodegradable scaffolds. Biomaterials 26:3941–3951. https://doi.org/10.1016/j.
biomaterials.2004.10.005
14. Liao H, Zhang H, Chen W (2009) Differential physical, rheological, and biological properties
of rapid in situ gelable hydrogels composed of oxidized alginate and gelatin derived from
marine or porcine sources. J Mater Sci Mater Med 20:1263–1271. https://doi.org/10.1007/
s10856-009-3694-4
15. Boanini E, Rubini K, Panzavolta S, Bigi A (2010) Chemico-physical characterization of gela-
tin films modified with oxidized alginate. Acta Biomater 6:383–388. https://doi.org/10.1016/j.
actbio.2009.06.015
16. Bouhadir KH, Lee KY, Alsberg E et al (2001) Degradation of partially oxidized alginate
and its potential application for tissue engineering. Biotechnol Prog 17:945–950. https://doi.
org/10.1021/bp010070p
17. Balakrishnan B, Mohanty M, Umashankar PR, Jayakrishnan a (2005) Evaluation of an in
situ forming hydrogel wound dressing based on oxidized alginate and gelatin. Biomaterials
26:6335–6342. https://doi.org/10.1016/j.biomaterials.2005.04.012
18. Sarker B, Papageorgiou DG, Silva R et al (2014) Fabrication of alginate–gelatin crosslinked
hydrogel microcapsules and evaluation of the microstructure and physico-chemical proper-
ties. J Mater Chem B 2:1470–1482. https://doi.org/10.1039/c3tb21509a
19. Sarker B, Singh R, Silva R et al (2014) Evaluation of fibroblasts adhesion and proliferation on
alginate-gelatin crosslinked hydrogel. PLoS One 9:e107952. https://doi.org/10.1371/journal.
pone.0107952
20. Rangarajan A, Hong SJ, Gifford A, Weinberg RA (2004) Species- and cell type-specific
requirements for cellular transformation. Cancer Cell 6:171–183. https://doi.org/10.1016/j.
ccr.2004.07.009
21. Griffith LG, M a S (2006) Capturing complex 3D tissue physiology in vitro. Nat Rev Mol
Cell Biol 7:211–224. https://doi.org/10.1038/nrm1858
22. Liu X, Ma X (2004) Polymeric scaffolds for bone tissue engineering. Ann Biomed Eng
32:477–486
23. Annabi N, Nichol JW, Zhong X et al (2010) Controlling the porosity and microarchitecture of
hydrogels for tissue engineering. Tissue Eng Part B Rev 16:371–383. https://doi.org/10.1089/
ten.teb.2009.0639
24. Zmora S, Glicklis R, Cohen S (2002) Tailoring the pore architecture in 3-D alginate scaffolds
by controlling the freezing regime during fabrication. Biomaterials 23:4087–4094
25. Sarker B, Hum J, Nazhat SN, Boccaccini AR (2015) Combining collagen and bioactive
glasses for bone tissue engineering: a review. Adv Healthc Mater 4:176–194. https://doi.
org/10.1002/adhm.201400302
26. Wang Y, Yang C, Chen X, Zhao N (2006) Development and characterization of novel biomi-
metic composite scaffolds based on bioglass-collagen-hyaluronic acid-phosphatidylserine for
tissue engineering applications. Macromol Mater Eng 291:254–262. https://doi.org/10.1002/
mame.200500381
27. Sun J, Tan H (2013) Alginate-based biomaterials for regenerative medicine applications.
Materials (Basel) 6:1285–1309. https://doi.org/10.3390/ma6041285
28. Shapiro L, Cohen S (1997) Novel alginate sponges for cell culture and transplantation.
Biomaterials 18:583–590. https://doi.org/10.1016/S0142-9612(96)00181-0
29. Sapir Y, Kryukov O, Cohen S (2011) Integration of multiple cell-matrix interactions into algi-
nate scaffolds for promoting cardiac tissue regeneration. Biomaterials 32:1838–1847. https://
doi.org/10.1016/j.biomaterials.2010.11.008
30. Florczyk SJ, Kim DJ, Wood DL, Zhang M (2011) Influence of processing parameters on pore
structure of 3D porous chitosan-alginate polyelectrolyte complex scaffolds. J Biomed Mater
Res – Part A 98(A):614–620. https://doi.org/10.1002/jbm.a.33153
152 B. Sarker and A.R. Boccaccini
31. Li Z, Ramay HR, Hauch KD et al (2005) Chitosan-alginate hybrid scaffolds for bone tissue
engineering. Biomaterials 26:3919–3928. https://doi.org/10.1016/j.biomaterials.2004.09.062
32. Petrenko YA, Ivanov RV, Petrenko a Y, Lozinsky VI (2011) Coupling of gelatin to inner sur-
faces of pore walls in spongy alginate-based scaffolds facilitates the adhesion, growth and
differentiation of human bone marrow mesenchymal stromal cells. J Mater Sci Mater Med
22:1529–1540. https://doi.org/10.1007/s10856-011-4323-6
33. Luo Z, Yang Y, Deng Y et al (2016) Peptide-incorporated 3D porous alginate scaffolds with
enhanced osteogenesis for bone tissue engineering. Colloids Surf B Biointerfaces 143:243–
251. https://doi.org/10.1016/j.colsurfb.2016.03.047
34. Yang C, Frei H, Rossi FM, Burt HM (2009) The differential in vitro and in vivo responses
of bone marrow stromal cells on novel porous gelatin-alginate scaffolds. J Tissue Eng Regen
Med 3:601–614. https://doi.org/10.1002/term.201
35. Florczyk SJ, Leung M, Li Z et al (2013) Evaluation of three-dimensional porous chitosan-
alginate scaffolds in rat calvarial defects for bone regeneration applications. J Biomed Mater
Res Part A 101:2974–2983. https://doi.org/10.1002/jbm.a.34593
36. Lin H-R, Yeh Y-J (2004) Porous alginate/hydroxyapatite composite scaffolds for bone tissue
engineering: Preparation, characterization, and in vitro studies. J Biomed Mater Res 71B:52–
65. https://doi.org/10.1002/jbm.b.30065
37. Rajesh R, Ravichandran D (2015) Development of a new carbon nanotube – alginate –
hydroxyapatite tricomponent composite scaffold for application in bone tissue engineering.
Int J Nanomedicine 10:7–15
38. Sowjanya JA, Singh J, Mohita T et al (2013) Biocomposite scaffolds containing chitosan/
alginate/nano-silica for bone tissue engineering. Colloids Surfs B Biointerfaces 109:294–
300. https://doi.org/10.1016/j.colsurfb.2013.04.006
39. Mishra R, Basu B, Kumar A (2009) Physical and cytocompatibility properties of bioactive
glass-polyvinyl alcohol-sodium alginate biocomposite foams prepared via sol-gel process-
ing for trabecular bone regeneration. J Mater Sci Mater Med 20:2493–2500. https://doi.
org/10.1007/s10856-009-3814-1
40. Suárez-González D, Barnhart K, Saito E et al (2010) Controlled nucleation of hydroxyapatite
on alginate scaffolds for stem cell-based bone tissue engineering. J Biomed Mater Res Part A
95A:222–234. https://doi.org/10.1002/jbm.a.32833
41. Sarker B, Li W, Zheng K et al (2016) Designing porous bone tissue engineering scaffolds
with enhanced mechanical properties from composite hydrogels composed of modified algi-
nate, gelatin, and bioactive glass. ACS Biomater Sci Eng. acsbiomaterials.6b00470. https://
doi.org/10.1021/acsbiomaterials.6b00470
42. Cai K, Zhang J, Deng L et al (2007) Physical and biological properties of a novel hydrogel
composite based on oxidized alginate, gelatin and tricalcium phosphate for bone tissue engi-
neering. Adv Eng Mater 9:1082–1088. https://doi.org/10.1002/adem.200700222
43. Li Z, Zhang M (2005) Chitosan-alginate as scaffolding material for cartilage tissue engineer-
ing. J Biomed Mater Res Part A 75:485–493. https://doi.org/10.1002/jbm.a.30449
44. Benya P (1982) Dedifferentiated chondrocytes reexpress the differentiated collagen phenotype
when cultured in agarose gels. Cell 30:215–224. https://doi.org/10.1016/0092-8674(82)90027-7
45. Homicz MR, Chia SH, Schumacher BL et al (2003) Human septal chondrocyte redifferentia-
tion in alginate, polyglycolic acid scaffold, and monolayer culture. Laryngoscope 113:25–32.
https://doi.org/10.1097/00005537-200301000-00005
46. Hsu SH, Shu WW, Hsieh SC et al (2004) Evaluation of chitosan-alginate-hyaluronate com-
plexes modified by an RGD-containing protein as tissue-engineering scaffolds for cartilage
regeneration. Artif Organs 28:693–703. https://doi.org/10.1111/j.1525-1594.2004.00046.x
47. Re’em T, Tsur-Gang O, Cohen S (2010) The effect of immobilized RGD peptide in macro-
porous alginate scaffolds on TGFβ1-induced chondrogenesis of human mesenchymal stem
cells. Biomaterials 31:6746–6755. https://doi.org/10.1016/j.biomaterials.2010.05.025
5 Alginate Utilization in Tissue Engineering and Cell Therapy 153
48. Wang L, Shansky J, Borselli C et al (2012) Design and fabrication of a biodegradable, cova-
lently crosslinked shape-memory alginate scaffold for cell and growth factor delivery. Tissue
Eng Part A 18:2000–2007. https://doi.org/10.1089/ten.tea.2011.0663
49. Chen F, Tian M, Zhang D et al (2012) Preparation and characterization of oxidized alginate
covalently cross-linked galactosylated chitosan scaffold for liver tissue engineering. Mater
Sci Eng C 32:310–320. https://doi.org/10.1016/j.msec.2011.10.034
50. Sapir Y, Cohen S, Friedman G, Polyak B (2012) The promotion of in vitro vessel-like orga-
nization of endothelial cells in magnetically responsive alginate scaffolds. Biomaterials
33:4100–4109. https://doi.org/10.1016/j.biomaterials.2012.02.037
51. Murphy SV, Skardal A, Atala A (2013) Evaluation of hydrogels for bio-printing applications.
J Biomed Mater Res A 101:272–284. https://doi.org/10.1002/jbm.a.34326
52. Luo Y, Lode A, Wu C et al (2015) Alginate/nanohydroxyapatite scaffolds with designed core/
shell structures fabricated by 3D plotting and in situ mineralization for bone tissue engineer-
ing. ACS Appl Mater Interfaces 7:6541–6549. https://doi.org/10.1021/am508469h
53. Luo Y, Wu C, Lode A, Gelinsky M (2013) Hierarchical mesoporous bioactive glass/alginate
composite scaffolds fabricated by three-dimensional plotting for bone tissue engineering.
Biofabrication 5:15005. https://doi.org/10.1088/1758-5082/5/1/015005
54. Wang X, Tolba E, Der HCS et al (2014) Effect of bioglass on growth and biomineralization
of saos-2 cells in hydrogel after 3d cell bioprinting. PLoS One 9:1–7. https://doi.org/10.1371/
journal.pone.0112497
55. Lee H, Ahn S-H, Kim GH (2012) Three-dimensional collagen/alginate hybrid scaffolds
functionalized with a drug delivery system (DDS) for bone tissue regeneration. Chem Mater
24:881–891. https://doi.org/10.1021/cm200733s
56. Zehnder T, Sarker B, Boccaccini AR, Detsch R (2015) Evaluation of an alginate–
gelatine crosslinked hydrogel for bioplotting. Biofabrication 7:1–12. https://doi.
org/10.1088/1758-5090/7/2/025001
57. Grigore A, Sarker B, Fabry B et al (2014) Behavior of encapsulated MG-63 cells in RGD
and gelatine-modified alginate hydrogels. Tissue Eng Part A 20:2140–2150. https://doi.
org/10.1089/ten.tea.2013.0416
58. Detsch R, Sarker B, Zehnder T et al (2015) Advanced alginate-based hydrogels. Mater Today
18:590–591. https://doi.org/10.1016/j.mattod.2015.10.013
59. Sarker B, Rompf J, Silva R et al (2015) Alginate-based hydrogels with improved adhesive
properties for cell encapsulation. Int J Biol Macromol 78:72–78. https://doi.org/10.1016/j.
ijbiomac.2015.03.061
60. Leite ÁJ, Sarker B, Zehnder T et al (2016) Bioplotting of a bioactive alginate dialdehyde-
gelatin composite hydrogel containing bioactive glass nanoparticles. Biofabrication 8:35005.
https://doi.org/10.1088/1758-5090/8/3/035005
61. Marijnissen WJC, van Osch GJV, Aigner J et al (2002) Alginate as a chondrocyte-delivery
substance in combination with a non-woven scaffold for cartilage tissue engineering.
Biomaterials 23:1511–1517. https://doi.org/10.1016/S0142-9612(01)00281-2
62. Kundu J, Shim J-H, Jang J et al (2015) An additive manufacturing-based PCL-alginate-
chondrocyte bioprinted scaffold for cartilage tissue engineering. J Tissue Eng Regen Med
9:1286–1297. https://doi.org/10.1002/term.1682
63. Narayanan LK, Huebner P, Fisher MB et al (2016) 3D-bioprinting of polylactic acid (PLA)
nanofiber–alginate hydrogel bioink containing human adipose-derived stem cells. ACS
Biomater Sci Eng 2:1732–1742. https://doi.org/10.1021/acsbiomaterials.6b00196
64. Markstedt K, Mantas A, Tournier I et al (2015) 3D bioprinting human chondro-
cytes with nanocellulose- alginate bioink for cartilage tissue engineering applications.
Biomacromolecules 16:1489–1496. https://doi.org/10.1021/acs.biomac.5b00188
65. Wang CC, Yang KC, Lin KH et al (2011) A highly organized three-dimensional alginate
scaffold for cartilage tissue engineering prepared by microfluidic technology. Biomaterials
32:7118–7126. https://doi.org/10.1016/j.biomaterials.2011.06.018
154 B. Sarker and A.R. Boccaccini
66. Park J, Lee SJ, Chung S et al (2017) Cell-laden 3D bioprinting hydrogel matrix depending on
different compositions for soft tissue engineering: characterization and evaluation. Mater Sci
Eng C 71:678–684. https://doi.org/10.1016/j.msec.2016.10.069
67. Gao Q, He Y, J zhong F et al (2015) Coaxial nozzle-assisted 3D bioprinting with built-in
microchannels for nutrients delivery. Biomaterials 61:203–215. https://doi.org/10.1016/j.
biomaterials.2015.05.031
68. Man Y, Wang P, Guo Y et al (2012) Angiogenic and osteogenic potential of platelet-rich
plasma and adipose-derived stem cell laden alginate microspheres. Biomaterials 33:8802–
8811. https://doi.org/10.1016/j.biomaterials.2012.08.054
69. Ausländer S, Wieland M, Fussenegger M (2012) Smart medication through combina-
tion of synthetic biology and cell microencapsulation. Metab Eng 14:252–260. https://doi.
org/10.1016/j.ymben.2011.06.003
70. Li H-B, Jiang H, Wang C-Y et al (2006) Comparison of two types of alginate microcapsules
on stability and biocompatibility in vitro and in vivo. Biomed Mater 1:42–47. https://doi.
org/10.1088/1748-6041/1/1/007
71. Schacht K, Jüngst T, Schweinlin M et al (2015) Biofabrication of cell-loaded 3D spider silk
constructs. Angew Chemie Int Ed 54:2816–2820. https://doi.org/10.1002/anie.201409846
72. Dhote V, Skaalure S, Akalp U et al (2013) On the role of hydrogel structure and degradation
in controlling the transport of cell-secreted matrix molecules for engineered cartilage. J Mech
Behav Biomed Mater 19:61–74. https://doi.org/10.1016/j.jmbbm.2012.10.016
73. Lim F, Sun A (1980) Microencapsulated islets as bioartificial endocrine pancreas. Science
210:908–910. https://doi.org/10.1126/science.6776628
74. Abbah SA, WW L, Chan D et al (2006) In vitro evaluation of alginate encapsulated adipose-
tissue stromal cells for use as injectable bone graft substitute. Biochem Biophys Res Commun
347:185–191. https://doi.org/10.1016/j.bbrc.2006.06.072
75. Moshaverinia A, Ansari S, Chen C et al (2013) Co-encapsulation of anti-BMP2 monoclonal
antibody and mesenchymal stem cells in alginate microspheres for bone tissue engineering.
Biomaterials 34:6572–6579. https://doi.org/10.1016/j.biomaterials.2013.05.048
76. Freire MO, You H-K, Kook J-K et al (2011) Antibody-mediated osseous regeneration: a novel
strategy for bioengineering bone by immobilized anti–bone morphogenetic protein-2 anti-
bodies. Tissue Eng Part A 17:2911–2918. https://doi.org/10.1089/ten.tea.2010.0584
77. Hwang Y-S, Cho J, Tay F et al (2009) The use of murine embryonic stem cells, alginate
encapsulation, and rotary microgravity bioreactor in bone tissue engineering. Biomaterials
30:499–507. https://doi.org/10.1016/j.biomaterials.2008.07.028
78. Wang H, Lee J-K, Moursi A, Lannutti JJ (2003) Ca/P ratio effects on the degradation of
hydroxyapatite in vitro. J Biomed Mater Res A 67:599–608. https://doi.org/10.1002/
jbm.a.10538
79. Arpornmaeklong P, Kochel M, Depprich R et al (2004) Influence of platelet-rich plasma
(PRP) on osteogenic differentiation of rat bone marrow stromal cells. An in vitro study. Int
J Oral Maxillofac Surg 33:60–70. https://doi.org/10.1054/ijom.2003.0492
80. Kanno T, Takahashi T, Tsujisawa T et al (2005) Platelet-rich plasma enhances human
osteoblast-like cell proliferation and differentiation. J Oral Maxillofac Surg 63:362–369.
https://doi.org/10.1016/j.joms.2004.07.016
81. Rottensteiner U, Sarker B, Heusinger D et al (2014) In vitro and in vivo biocompatibility of
alginate dialdehyde/gelatin hydrogels with and without nanoscaled bioactive glass for bone
tissue engineering applications. Materials (Basel) 7:1957–1974. https://doi.org/10.3390/
ma7031957
82. Song K, Yang Y, Li S et al (2014) In vitro culture and oxygen consumption of NSCs in size-
controlled neurospheres of Ca-alginate/gelatin microbead. Mater Sci Eng C Mater Biol Appl
40:197–203. https://doi.org/10.1016/j.msec.2014.03.028
83. Silva R, Singh R, Sarker B et al (2014) Hybrid hydrogels based on keratin and alginate for
tissue engineering. J Mater Chem B 2:5441–5451. https://doi.org/10.1039/c4tb00776j
5 Alginate Utilization in Tissue Engineering and Cell Therapy 155
84. Silva R, Singh R, Sarker B et al (2016) Soft-matrices based on silk fibroin and alginate
for tissue engineering. Int J Biol Macromol 2:5441–5451. https://doi.org/10.1016/j.
ijbiomac.2016.04.045
85. Zhou H, HHK X (2011) The fast release of stem cells from alginate-fibrin microbeads in
injectable scaffolds for bone tissue engineering. Biomaterials 32:7503–7513. https://doi.
org/10.1016/j.biomaterials.2011.06.045
86. Zeng Q, Han Y, Li H, Chang J (2014) Bioglass/alginate composite hydrogel beads as cell
carriers for bone regeneration. J Biomed Mater Res Part B Appl Biomater 102:42–51. https://
doi.org/10.1002/jbm.b.32978
87. Gorustovich A a, Roether JA, Boccaccini AR (2010) Effect of bioactive glasses on angiogen-
esis: a review of in vitro and in vivo evidences. Tissue Eng Part B Rev 16:199–207. https://
doi.org/10.1089/ten.TEB.2009.0416
88. Hoppe A, Güldal NS, Boccaccini AR (2011) A review of the biological response to ionic dis-
solution products from bioactive glasses and glass-ceramics. Biomaterials 32:2757–2774
89. Moshaverinia A, Xu X, Chen C et al (2013) Dental mesenchymal stem cells encapsulated in
an alginate hydrogel co-delivery microencapsulation system for cartilage regeneration. Acta
Biomater 9:9343–9350. https://doi.org/10.1016/j.actbio.2013.07.023
90. Guo J, Jourdian GW, Maccallum DK (1989) Culture and growth characteristics of chon-
drocytes encapsulated in alginate beads. Connect Tissue Res 19:277–297. https://doi.
org/10.3109/03008208909043901
91. Ma H-L, Hung S-C, Lin S-Y et al (2003) Chondrogenesis of human mesenchymal stem cells
encapsulated in alginate beads. J Biomed Mater Res 64A:273–281. https://doi.org/10.1002/
jbm.a.10370
92. Endres M, Wenda N, Woehlecke H et al (2010) Microencapsulation and chondrogenic
differentiation of human mesenchymal progenitor cells from subchondral bone mar-
row in Ca-alginate for cell injection. Acta Biomater 6:436–444. https://doi.org/10.1016/j.
actbio.2009.07.022
93. Chang J-C, Hsu S-H, Chen DC (2009) The promotion of chondrogenesis in adipose-derived
adult stem cells by an RGD-chimeric protein in 3D alginate culture. Biomaterials 30:6265–
6275. https://doi.org/10.1016/j.biomaterials.2009.07.064
94. Focaroli S, Teti G, Salvatore V et al (2016) Calcium/cobalt alginate beads as func-
tional scaffolds for cartilage tissue engineering. Stem Cells Int 2016:1–12. https://doi.
org/10.1155/2016/2030478
95. Gaetani P, Torre ML, Klinger M et al (2008) Adipose-derived stem cell therapy for interver-
tebral disc regeneration: an in vitro reconstructed tissue in alginate capsules. Tissue Eng Part
A 14:1415–1423. https://doi.org/10.1089/ten.tea.2007.0330
96. Ansari S, Chen C, Xu X et al (2016) Muscle tissue engineering using gingival mesenchy-
mal stem cells encapsulated in alginate hydrogels containing multiple growth factors. Ann
Biomed Eng 44:1908–1920. https://doi.org/10.1007/s10439-016-1594-6
97. Kreeger PK, Deck JW, Woodruff TK, Shea LD (2006) The in vitro regulation of ovarian fol-
licle development using alginate-extracellular matrix gels. Biomaterials 27:714–723. https://
doi.org/10.1016/j.biomaterials.2005.06.016
98. Manju S, Muraleedharan CV, Rajeev A et al (2011) Evaluation of alginate dialdehyde cross-
linked gelatin hydrogel as a biodegradable sealant for polyester vascular graft. J Biomed
Mater Res B Appl Biomater 98:139–149. https://doi.org/10.1002/jbm.b.31843
99. Yu J, KT D, Fang Q et al (2010) The use of human mesenchymal stem cells encapsulated
in RGD modified alginate microspheres in the repair of myocardial infarction in the rat.
Biomaterials 31:7012–7020. https://doi.org/10.1016/j.biomaterials.2010.05.078
100. Engler AJ, Sen S, Sweeney HL, Discher DE (2006) Matrix elasticity directs stem cell lineage
specification. Cell 126:677–689. https://doi.org/10.1016/j.cell.2006.06.044
101. Moshaverinia A, Xu X, Chen C et al (2014) Application of stem cells derived from the
periodontal ligament orgingival tissue sources for tendon tissue regeneration. Biomaterials
35:2642–2650. https://doi.org/10.1016/j.biomaterials.2013.12.053
Chapter 6
Alginate-Based Three-Dimensional In Vitro
Tumor Models: A Better Alternative
to Current Two-Dimensional Cell Culture
Models
Amit Khurana and Chandraiah Godugu
The advancements in the biological sciences resulted in the origin of cell culture-
based experimental models to study various biological phenomena. Most of the ini-
tial animal-based experiments are performed on in vitro cell culture models.
Currently, in vitro cell culture-based models play a crucial role in new drug discov-
ery and developmental process. The cell culture models have important role in can-
cer research. The initial anticancer effects of new chemical entities and novel
anticancer formulations are performed on in vitro cancer cell lines [1]. Although
these cell culture-based models offer numerous advantages in various stages of drug
discovery and developmental process, they also suffer from several shortcomings.
Most of the cell culture studies are based on the traditional two-dimensional (2D)
models where the cells are grown on a polystyrene plastic surface. In these models,
the plastic cell culture surfaces are coated with biocompatible matrix which supports
the cell attachment and growth [2]. Though 2D models offer the advantage of ease of
subculture and cell isolation for further molecular studies, they often produce exces-
sive positive response in such a way that these effects are observed only in in vitro
cell culture models and may never translate those effects while performing the
in vivo animal models for confirmation of observed in vitro effects. When 2D-based
models are used, generally there is found to be poor in vitro and in vivo correlation.
The in vitro and in vivo correlation was one of the biggest challenges while develop-
ing the new drugs and formulations. Experimental outcomes when performed on 2D
cell culture models are in general found to produce increased sensitive responses,
and questions on relevance of conventional 2D grown cancer cells have been raised
[3–5]. That means 2D models suffer from several disadvantages including (1) lack of
cell–cell interactions, (2) lack of cell–extracellular matrix (ECM) interactions, (3)
lack of permeation barriers to hinder the diffusion of drugs or agents into cells, (4)
lack of attainment of original shape, etc. [6, 7]. Owing to multiple shortcomings
associated with 2D cell culture-based models, scientists felt the need to develop new
in vitro models with better correlation with in vivo condition. As a result of the con-
quest for finding better alternatives, a multiple three-dimensional (3D) cell culture-
based models were proposed with reproducibility and integrity across a wide array
of variables including the matrix used and the type of culture system employed
(adherent and suspension) [8, 9]. The 3D cell culture-based systems/matrices, also
6 Alginate-Based Three-Dimensional In Vitro Tumor Models: A Better Alternative… 159
discrepancies associated with 2D cell culture and provides a better platform for
oncology drug discovery and development process at a wide array of closets includ-
ing the preclinical cytotoxicity screening, screening of pharmacological activity,
and toxicological evaluation. It is expected that the development of 3D tumor mod-
els will reduce animal testing, yield more predictive data, improve cell culture effi-
ciency, reduce cost and time to identify new drug candidate, and reduce development
time to market [17, 18].
6.2 P
roblem with In Vitro and In Vivo Correlation Based
on 2D Cell Culture Models
As discussed in the introduction, there are multiple areas where 2D and 3D cell
culture systems differ including the cellular interactions, morphology, and cellular
dynamics. Another important area of relevance is how much we can correlate these
systems with in vivo conditions be it the 2D or the 3D systems. In general, owing to
the multiple advantages associated with 3D system, one can easily judge 3D sys-
tems to be much superior to the 2D systems. Moreover, in terms of clinical and
in vivo correlation, 3D systems stand way ahead of 2D systems. For in vitro to
in vivo translation of results for successful cancer chemotherapy, it is essential to
mimic the in vivo environment in the in vitro models of screening. The rationale
behind using a 3D system in vitro is to mimic the 3D tumorlike microenvironment
which gives better correlation with in vivo results. For example, multiple research
findings report that 3D cell culture systems are much more resistant to anticancer
drugs compared to 2D systems [5, 19]. The difference in sensitivity may be attrib-
uted to the poor diffusion of drug across the tumor. Another explanation may be
that, owing to 3D microenvironment, the tumor mass may become hypoxic which
may upregulate some of the important genes responsible for drug resistance [20].
Imamura et al. showed that multicellular spheroids of breast cancer cells become
hypoxic causing an increase in the cells in G0 phase and/or suppression of caspase-
3, and the results were in line with the in vivo patient-derived tumors, in contrast to
the results obtained with 2D monolayer of the same cells [21]. It is almost impos-
sible to study tumor fibrosis in case of 2D culture, whereas the same can be decently
mimicked with 3D multicellular tumor spheroids (MCTS). Hypoxia and tissue
necrosis are certain issues associated with molecular complexities of the 3D micro-
environment and are observed in vivo as well. Study of these events in 2D mono-
layer cell culture poorly correlates with in vivo microenvironment. In contrast,
study of hypoxia is very much possible with 3D MCTS and better correlate with
in vivo tumors. Thus, these observations serve as a warning on the relevance of 2D
systems for anticancer drug screening and warrant due consideration for the use of
pragmatic 3D in vitro models for anticancer drug research.
6 Alginate-Based Three-Dimensional In Vitro Tumor Models: A Better Alternative… 161
Bridging the gap between 2D and 3D cell culture requires a base matrix which can
allow the cells to sufficiently grow in an environment which mimics the in vivo
conditions and at the same time protects the cellular physiology from extreme
hypoxia owing to the limited fluid turbulence during 3D cell culture. Wide arrays of
matrices of natural origin have been explored for 3D cell culture including collagen
and alginate as the most widely studied scaffold materials. Among the natural scaf-
folds, alginate has received wide attention by virtue of its high biocompatibility, low
toxicity, and low cost. It is an anionic polymer isolated from brown seaweeds pos-
sessing excellent gelation properties which can be controlled for generating wide
variety of scaffolds with various sizes and matrix properties [22]. AlgiMatrix™ is
an innovative marketed product from Invitrogen (Thermo Fisher Scientific) for
development of tumor spheroids. It is a biologically inert alginate sponge-based
(with pore size in the range of 50–200 μm) bioscaffold for 3D cell culture. The
product is available in 6-well, 24-well, and 96-well format for growing spheroids
with remarkable similarities with the in vivo phenotype of cells. AlgiMatrix™ 3D
culture system can be applied to a wide number of cell-based assays, including
multicellular tumor spheroid (MCTS) assays, hepatocyte and cardiomyocyte organ-
ogenesis studies, coculture studies, high-throughput drug screening, and embryonic
stem cell 3D differentiation studies. These scaffolds have a unique feature of dis-
solution of matrix by using dissolving buffer, which is a nonenzymatic solution and
solubilizes the matrix in few minutes, and the grown spheroids can be isolated for
performing various assays like staining, histology, immunohistochemistry, Western
blotting, RT-PCR, etc. It has been observed that the 3D alginate scaffold microenvi-
ronment promotes the development of better cellular phenotype compared to tradi-
tional sandwich culture as evident from long-term viability, toxicity assays,
molecular expressions, and ultrastructural features [17].
Our group has standardized the tumor spheroids from H460, A549, and H1650
non-small cell lung cancer (NSCLC) cell lines on this unique 3D scaffold and com-
pared the results with conventional monolayer 2D culture. The tumor spheroids
were cultured in 6-well and 96-well formats. We could develop highly reproducible
3D tumor spheroids in the size range of 100–250 μm with optimized cell density.
We limited the size to below 250 μm. The nutrient distribution and air circulation
are compromised with larger-sized tumor spheroids, and the risk of necrotic lesions
arises. The IC50 values showed significant differences in 2D and 3D formats for the
conventional anticancer drugs (cisplatin, docetaxel, gemcitabine, 5-fluorouracil,
and camptothecin) which correlate with earlier reports. There was a significant
reduction in the expression of caspase-3 in the 3D culture format (2.09- and 2.47-
folds, respectively, for 5-fluorouracil and camptothecin) compared to the 2D format
in H460 spheroids. In addition, we could study the cancer stem cell population
behavior in H1650 stem cells and parental cells where the number of spheroids was
found to be higher for stem cells compared to parental population. The stem cell
162 A. Khurana and C. Godugu
Fig. 6.1 Drug and nanoparticles uptake by spheroids. (a) Representative images of doxorubicin
uptake in H1650 parental cell spheroids grown in 3D alginate scaffold system uptake images were
taken at 0.5, 1, and 2 h time points. (b) Representative images of the nanoparticle uptake by H1650
parental cell spheroids. 1,1′-Dioctadecyl-3,3,3′,3′-tetramethylindodicarbocyanine perchlorate
(DIO) oil was encapsulated in nanolipid carrier (NLC) and incubated with the spheroids for 2 h.
(a) Fluorescent image, (b) DIC image, and (c) merged image. The fluorescent images clearly indi-
cate the nanoparticles uptake into spheroids. (c) Relative fluorescence intensities of free doxorubi-
cin uptake and DIO oil loaded NLC nanoparticles into 3D spheroids. Each data point is represented
as mean±sem (n = 3). **P<0.01 vs doxorubicin group (Reproduced from Ref. [17])
spheroids exhibited higher IC50 values, and the results correlated with previous find-
ings [23]. The tumor penetration/uptake results also depicted the extended role of
matrix proteins as evident from poor penetration of doxorubicin (10.52%) and even
nanoformulations (3.41%) as shown in Fig. 6.1. Moreover, we investigated the
molecular features of the spheroids by immunohistochemistry and RT-PCR as well
and concluded that the tumor spheroids in 3D culture show higher chemoresistance
compared to 2D monolayer culture [17].
Canine mammary tumors (CMTs) are the most common type of cancer in female
dogs. Most CMT have epithelial origin and are known as simple adenoma/simple
carcinoma (SC), some consist of both epithelial and myoepithelial tissues and are
called complex carcinoma (CC). Cardoso et al. developed 3D tumor spheroids of
6 Alginate-Based Three-Dimensional In Vitro Tumor Models: A Better Alternative… 163
canine complex carcinoma (CC) and canine simple carcinoma (SC) cell lines
derived from canine tumors in 6-well AlgiMatrix™ plates and reported spheroids in
size range 50–125 μm for CC and 175–200 μm for SC. The spheroids were grown
for 2 weeks and thereafter evaluated the expression of various molecular markers.
3D tumor spheroids obtained from both the cell lines showed expression of epider-
mal growth factor receptor (EGFR). However, there was differential expression of
the same in spheroids from both the cell lines. Microarray analysis showed upregu-
lation of crucial extracellular matrix proteins (MMP1, MMP3, MMP9, and MMP13)
and downregulation of cadherin-1. The results obtained with 3D tumor spheroids
were in concordance with the in vivo canine tumors showing the higher correlation
of 3D tumor models with in vivo tumors. This study demonstrated the utility of
AlgiMatrix™ 3D tumor models for veterinary oncology and opens new avenue for
animal 3D tumor-derived cells for future research [24].
Thus, 3D AlgiMatrix can fill the gap in the area of preclinical high-throughput
screening (HTS) of anticancer drugs owing to the teething discrepancies in 2D and
3D in vitro cancer models and may serve as a better method of cytotoxicity evalua-
tion with higher correlation to in vivo results.
6.4 A
lginate Scaffold as a Matrix for 3D Cell Culture-Based
Tumor Model
A wide variety of matrices have been explored for developing 3D tumor models.
Some of the most commonly used matrices are collagen-based scaffolds, hyaluronic
acid-based scaffolds, alginate-based scaffolds, peptide-based scaffolds, hydrogels,
etc. Alginate-based scaffolds have received significant attention for 3D cell culture
and especially 3D tumor models for anticancer studies. Some of the advantages of
alginate scaffold are animal-free product, stable at room temperature, and flexible
porosity. The animal-derived scaffolds like collagen, fibrin, or hyaluronic acid pos-
sess bioactivity. In contrast, alginate is bioinert and does not contain intrinsic bioac-
tivity; thus, its bioactive properties can be improved by modifications with peptides
or by using composite hydrogels, such as alginate–collagen [25–27]. In case of
cancer research, a desirable need is the formation of multicellular spheroids which
can mimic the in vivo tumor microenvironment without compromising the respira-
tory features of the tumor owing to the fact that continuous culture for a long dura-
tion may lead to the development of hypoxia-induced necrotic regions which may
lead to discrepancies in the results obtained. By virtue of its highly porous nature,
alginate scaffolds allow the formation of multicellular tumor spheroids. Unique fea-
ture of alginate scaffold spurred substantial efforts toward the development of 3D
tumor models, and a significant number of reports are a proof of the commercial
viability of the products based on alginate matrix.
164 A. Khurana and C. Godugu
similar to in vivo conditions that promotes the conversion of cultured cancer cells to
a more malignant in vivo-like phenotype. The 3D scaffolds were prepared by lyoph-
ilization and subsequent cross-linking of chitosan and alginate. The formed scaf-
folds were highly porous to allow the influx of cells throughout the scaffold and
provide a large surface area for cell attachment and proliferation, ideal for modeling
the tumor microenvironment. The tumor model was established by seeding U-87
MG and U-118 MG human glioma cells on the scaffolds and allowing the tumor
cells to proliferate in vitro for 10 days. The developed 3D matrix was evaluated for
human (U-87 MG and U-118 MG) and rat C6 glioma cells (cancer stem cell-like
cells). The in vitro expression of VEGF, MMP-2, fibronectin, and laminin was
assessed to understand the malignancy and angiogenesis potential. The electron
microscopy results showed biocompatibility of the CA scaffold as the studied cell
lines could grow inside the scaffold. The results were compared to cells grown in 2D
(24-well plates) and in 3D Matrigel matrix. The cells growing in the CA matrix were
found to grow at a slower pace. Similarity to the in vivo microenvironment condi-
tions was speculated to be the reason behind the observed retarded growth. The
scaffolds precultured with U-87 MG and C6 cells for 10 days were then implanted
into nude mice to evaluate tumor proliferation and angiogenesis compared to the
standard 2D cell culture and 3D Matrigel matrix xenograph controls. The study
presented interesting findings as the malignancy potential of the human glioma cell
lines considerably increased, whereas the results were similar for the 2D cell culture
and Matrigel-implanted C6 glioma cells when compared to the CA scaffold in vivo.
The results of 3D scaffolds outweighed the 2D monolayer-based results. The immu-
nohistochemistry for CD31, an angiogenesis marker, showed notably higher expres-
sion in the CA scaffold-grown tumors suggesting the close resemblance of cancer
microenvironment inside the CA scaffolds [30]. The CA scaffolds provided a 3D
microenvironment for the cancer cells and were successfully grown in vivo, indicat-
ing the potential for further preclinical exploration for developing better 3D systems
to reduce the use of animal experiments in anticancer drug discovery.
The same group explored the utility of this versatile CA scaffold to understand
the ex vivo tumor cell physiological interaction between prostate cancer cells
(LNCaP, C4-2, and C4-2B) and peripheral blood lymphocytes (PBLs) to aid in
understanding the immunotherapeutic behavior of castration-resistant prostate can-
cer. The results were found to be interesting. The growth of tumor spheroids was
supported well by CA scaffolds and Matrigel. The growth could be monitored up to
15 days, whereas the in situ imaging of spheroids could be possible even up to
55 days. Live cell imaging is an excellent tool to observe and measure real-time
events and provides a useful means to monitor the external surface of tumor spher-
oids and their interaction with surrounding matrix and other cells. Live cell imaging
by using tracker dyes revealed the infiltration of immune cells within the 3D micro-
environment of tumor spheroids. Scanning electron microscopy (SEM) images
showed dynamic interaction of PBLs with tumor spheroids in CA scaffolds as well
as Matrigel. The interaction was found to be significant on the second day.
Heterogeneity in the PBL population was observed, and these immune cells were
distinguished from tumor cells by their morphology and size. In addition, the
166 A. Khurana and C. Godugu
localization of specific PBLs on the cancer cells seeded scaffolds was assessed via
immunohistochemical analysis of CD45R staining (for B cells and activated T
cells). CD45R+ cells were observed bound to the in vitro tumors in CA scaffolds
after 2 days, which is in confirmation with the SEM results. The cells could be har-
vested for flow cytometry-based analysis and showed promising results with CA
scaffolds in terms of future therapeutic utility [31]. This study opened up a new
innovation 3D culture where understanding the interaction of microenvironment
cells with tumor spheroids could be possible and lays foundation to head on and
develop matrices where tumor along with multiple other TME cells can be grown
and simulate the in vivo tumors within the boundaries of 3D culture. The use of 3D
biomaterial scaffolds for immunological or immunotherapy studies provides a more
relevant model that should accelerate the successful translation of new immuno-
therapies into the clinic.
Gene therapy for cancer has shown promising results preclinically and offers
molecular treatment of cancer. Different approaches have been used to ferry the
payload to tumors. In a subsequent study with CA scaffolds, the relevance of this
novel CA scaffold was investigated to mimic 3D microenvironment for prostate
cancer and its utility as tumor model for nanoparticle-mediated gene therapy. The
results were compared with 2D-cultured cells and were found distinguishably mod-
ulated in 3D-cultured cells. The selected TRAMP-C2 (TC-2) prostate cancer cells
could form nice tumor spheroids when cultured in CA scaffolds. The harvested
spheroids showed upregulation of ECM and epithelial to mesenchymal transition
(EMT) gene markers as compared to standard 2D culture, indicating better mimicry
of in vivo conditions. Targeted nanoparticle (NP)-mediated delivery of red fluores-
cent protein (RFP) plasmid DNA into TC-2 tumor spheroids was achieved with
chlorotoxin (CTX) in cells cultured in CA scaffolds. However, in case of 2D culture,
the delivery could not be achieved. This finding questions the utility of the 3D cul-
ture for testing cationic-targeted NPs with CA scaffolds as the standard 2D culture
did not show the targeting effect of CTX. The 3D culture results could be repro-
duced in vivo model, thus proving the viability of this platform for wide-scale usage
[32]. This scaffold may be of utility in drug delivery programs as good correlation
in vitro and in vivo results could be generated as compared to standard 2D format.
Ginzberg et al. explored the use of porous 3D scaffolds of alginate for the entrap-
ment of retroviral vector-producing cells (VPC) for producing a local and sustained
release of viral particles for the treatment of murine colon cancer (MC38 tumors)
in vivo. The incorporated gene codes for the prodrug-activating enzyme, which pro-
duces local cytotoxicity. In this study, PA317/STK cells were used as VPC, which
carries the gene for herpes simplex virus thymidine kinase (HSV-tk) and constantly
produces a local release. HSV-tk monophosphorylates the prodrugs ganciclovir and
N-methanocarbathymidine (N-MCT), which are further converted to triphosphate
form by kinases. The triphosphate form produces DNA adducts and leads to cancer
cell death. By using the static seeding method, VPC were seeded into porous 3D
alginate scaffolds at two densities 0.5×106 and 1×106. The cell-seeded 3D system
was thoroughly characterized for cell number, cell leakage, and spheroid
morphology. The VPC were found to form spheroids after 1 day and high seeding
6 Alginate-Based Three-Dimensional In Vitro Tumor Models: A Better Alternative… 167
Fig. 6.2 (a) Chitosan/collagen/alginate (CCA) fibers under scanning electron microscope (SEM),
(b–d) the cell mass 3 days after cell seeding. (c) and (d) are enlarged partial views of (b)
(Reproduced with permission from Ref. [34])
tumor spheroids (MCTS) may replicate several parameters of the tumor microenvi-
ronment, including oxygen and nutrient gradients as well as the development of
dormant tumor regions. Wenzel et al. reported the setup of a 3D MCTS assay on
384-well microtiter plates compatible for high-content screening system. The plat-
form could exhibit the characteristics of real-time tumor with the study of inner core
of tumor, which shows distinct respiration profile. The developed system was used
to identify nine substances from two commercially available drug libraries that spe-
cifically target cells in inner MCTS core regions, while cells in outer MCTS regions
or in 2D cell culture remain unaffected. They identified all hits as being inhibitors
of the respiratory chain and further characterized their mode of action in MCTS as
shown in Fig. 6.4. The cumulative data suggested that targeting cytostatic-resistant
tumor cells in dormant tumor regions with respiratory chain inhibitors could be a
therapeutic option that could enhance the effectiveness of cytostatic-based chemo-
therapy, and 3D MCTS may be a boon to screen such agents [35]. These findings
could facilitate the establishment of secondary assays in more extensive screening
programs for early hit classification for respiratory chain inhibition.
Microarray is a high-end technique where the genetic profile can be screened on
high-throughput scale. Meli et al. developed an alginate-based 3D cellular microarray
6 Alginate-Based Three-Dimensional In Vitro Tumor Models: A Better Alternative… 169
platform for the high-throughput analysis of growth, cytotoxicity, and protein expres-
sion profile of a liver cancer cell line, HepG2. The results obtained were compared to
2D and 3D environments at the preliminary developmental scale. The antiprolifera-
tive effects of four anticancer drugs, tamoxifen, 5-fluorouracil, doxorubicin, and ami-
triptyline, were studied as a function of seeding density in the selected three platforms.
The 3D-cultured grown liver cancer cells showed substantial resistance at high cell
number seeding, whereas no seeding density dependence was observed in the IC50
values obtained in the 3D microarray culture platform. These results could be
explained by higher confluency which limits cell growth by virtue of restricted cell–
cell contact and nutrient supply. To explore the versatility of this platform, additional
studies on biocompatible chips was carried out. The in-cell immunofluorescence (IF)
assay provided quantitative data on the levels of specific target proteins involved in
proliferation, adhesion, angiogenesis, and drug metabolism and was used to compare
expression profiles between 2D and 3D environments. The upregulation of several
CYP450 enzymes, β1-integrin, and vascular endothelial growth factor (VEGF) in the
3D microarray cultures suggested that this platform provides a more in vivo-like
environment [36]. This approach thereby facilitates to better understand the variables
that affect drug resistance such as hypoxia, cell quiescence, and cell adhesion.
Moreover, the use of this platform for in-cell IF assays (on-chip) may aid in explora-
tion of critical factors playing role in 3D-dependent cell behavior and signaling at a
high-throughput scale.
As a new dimension of 3D culture, the role of 3D alginate gel beads was studied
for the investigation of metastasis. In a study, the metastatic potential of two
170 A. Khurana and C. Godugu
Fig. 6.4 Multicellular tumor spheroids (MCTS) mimic several parameters of the tumor
microenvironment. (a) T47D breast cancer MCTS: 2D projection of a 3D image stack of 142
z-planes with 2.58 μm spacing showing the organization of T47D spheroids. Cell nuclei are stained
with Hoechst (red) and dead cells are labeled with Sytox green (green). Scale bar, 100 lm. (b)
RT-qPCR shows upregulation of hypoxia and low nutrition-responsive genes in MCTS compared
to standard 2D cell culture conditions. Geometric mean of reference genes ACTB and RPLP13A
was used for normalization of all RT-qPCR results. Bars show average of 3 biological replicates
(+SD), p < 0.01. For full gene names, see material and methods. (c) Cryosections of an untreated
spheroid cultured for 5 days, stained with Hoechst (blue) and incubated 18 h with 5-ethynyl-2′-
deoxyuridine (EdU) probe (red). EdU, as a thymidine analogue, is incorporated into DNA in
S-phase and indicates proliferating cells. Mainly outer cell layers of the MCTS show EdU incor-
poration (see histogram of EdU signal in MCTS cross section). Scale bar, 100 lm. (d) Cytostatics,
paclitaxel (100 nM) and cisplatin (100 μM), mainly affect the outer proliferative layer in MCTS,
which could be removed after 3 days of treatment and 3 days of recovery by pipetting. An inner
cytostatic-resistant viable core could be isolated (arrows). The staurosporine (10 lm) cytotoxic
control leads to complete disruption of the MCTS. Scale bar, 100 lm. (e) Sytox green staining
shows that isolated cytostatic-resistant cores (see D) are viable. Scale bar, 100 μm (Reproduced
with permission from Ref. [35])
6 Alginate-Based Three-Dimensional In Vitro Tumor Models: A Better Alternative… 171
6.4.2 A
lginate-Based Scaffolds for Studying Cancer Stem Cell
(CSC) Biology
Cancer stem cells (CSCs) have emerged as an important factor responsible for clinical
failure of various chemotherapeutic agents. A large number of reports indicate the
unprecedented role of CSCs behind the emergence of chemoresistance and associated
complications with anticancer therapy in humans. There is an advent need of systems
where CSCs can be isolated and cultured and concentrated to study the molecular
pharmacodynamics. Study of CSCs is another dynamics of 3D cell culture and offers
unique features of concentration of CSCs from a cancer cell population by inducing
selection pressure. Therefore, simple, reliable, reproducible, and authentic 3D culture
platforms are needed to accelerate the CSC research.
Various studies indicate the successful utility of alginate-based 3D culture for
development of CSCs spheroids. In one report, human osteosarcoma cells were
isolated from patients, made into single-cell suspension, and were grown on algi-
nate gel for producing 3D microenvironment. Epirubicin was used to enrich the
CSCs by killing the general cancer cell population. The study revealed that the
enrichment of CSC under selection pressure and single-cell cloning spheres was
obtained with the features of CSCs after 7–10 days. Confirmation of molecular fea-
tures of CSCs was done by carrying immunofluorescent staining for Oct3/4 and
Nanog, markers for self-renewal and multipotential differentiation of CSCs. The
spheres were found to contain positive cells for both the markers indicating the
development of enriched CSCs with 3D culture. Moreover, the expression was
found to be more in the core compared to the periphery indicating the formation of
CSCs niche. The in vivo tumorigenicity study also showed similar pattern with suc-
cessful development of tumors 1 week after the inoculation. Similar characteristics
of CSCs were obtained upon immunohistochemical analysis of Oct3/4 and Nanog.
This study shows the importance of 3D culture for selective growth of CSCs with
in vivo reproducibility [43]. Therefore, study of CSCs by 3D culture provides better
correlation in vivo and needs to be further explored in this area.
The conventional methods of CSC enrichment take about 7–10 days with fre-
quent change in media and growth factors. Cellular encapsulation inside biocompat-
ible microcapsules is an attractive option for 3D cell culture of CSCs. Maintenance
of stemness is one of the important criteria for fruitful results with 3D cell culture.
Miniaturized 3D liquid core of core-shell microcapsules (CSMCs) with a hydrogel
shell has been shown to significantly better maintain the pluripotency of the pluripo-
tent stem cells compared to traditional culture [44]. In an attempt to reduce the time
taken for CSC enrichment, Rao et al. developed a miniature 3D liquid core of
microcapsules using alginate hydrogel, a strategy for concentration of CSCs from
PC-3 prostate carcinoma cells. The results were compared with ultralow attachment
plate (ULAP)-based enrichment method for CSCs. There was a significant differ-
ence in the number of prostaspheres formed by CSMCs compared to ULAP culture.
Encapsulation of 40 cells per microcapsules gave the highest number of prosta-
spheres with larger sizes compared to ULAP. Evaluation of CSCs surface markers
174 A. Khurana and C. Godugu
(CD44 and CD133) showed the superiority of prostaspheres obtained with CSMCs
with better results with prostaspheres obtained on day 2. The gene expression stud-
ies also revealed significant differences in the expression of Oct 4, Nanog, and Klf4
with day 2 prostaspheres obtained with CSMCs. In vivo xenograft studies also
showed superiority of day 2 prostaspheres obtained by CSMCs compared to ULAP
and parent PC3 cell inoculation as the tumors obtained by inoculation of 3000 pros-
taspheres from CSMCs were significantly compared to both the groups. In conclu-
sion, this study revealed better maintenance of stemness with the proposed
CSMC-based method of CSCs enrichment (due to autocrine signaling within the
microcapsule) and offers unique advantages over the conventional method of CSCs
enrichment by ULAP as the former is time-consuming (7–10 days) and very costly
(almost ten times the cost of normal plates). Thus, the miniaturized system based on
this approach may aid in CSC biology studies and may help in the elimination of
CSCs, the major culprit behind chemoresistance, recurrence, and metastasis [45]. In
an independent study, Xu et al. developed alginate gel bead-based 3D cell culture
platform for the enrichment of CSCs from human liver (HCCLM3) and human head
and neck squamous cell carcinoma cells (TCA8113). The results were compared
with 2D monolayer and showed some interesting findings. The CSCs obtained with
3D alginate beads were found to possess higher chemoresistance and tumorigenic-
ity compared to 2D culture. Both the tumor cell line-derived CSCs could thrive well
in alginate beads. The findings suggested suitability of alginate beads for expanding
and enriching CSCs of different histological origin and was claimed to be superior
to other inventions where different cell lines of same histological origin were stud-
ied. The markers for CSCs characteristics (Oct3/4, Nanog, Notch-1, Smoothened,
CD44, EpCAM, CD133, and ABCG2) were significantly modulated in both types
of 3D-cultured CSCs compared to respective 2D monolayer system indicating the
viability of alginate bead-based system for enrichment of CSCs and superior plat-
form for the investigation of growth and biophysical factors like matrix stiffness
(Fig. 6.5). Higher chemoresistance, metastatic potential, and tumorigenicity were
observed in both the types of 3D cultured CSCs. Immunostaining for von Willebrand
factor (vWF) showed the presence of this endothelial surface marker in the xeno-
grafts as shown in Fig. 6.6. Furthermore, significantly increased levels of hypoxia-
inducing factors (HIF1A and HIF2A) were observed in CSCs obtained in
3D-cultured cells compared to respective 2D-cultured CSCs (grown under normal
and low oxygen conditions) which might be the reason for the higher levels of CSCs
marker genes Nanog and Oct3/4 in 3D-cultured CSCs [25]. Collectively, these
results indicate that alginate hydrogel beads might be a better matrix compared to
synthetic (polyethyleneglycoldiacrylate) and natural polymers (agarose, collagen,
hyaluronic acid, etc.)-derived hydrogels by virtue of inertness, ease of dissolution
and reconstitution, and high mass transfer rate which aids in better cell viability.
However, despite the advantages, more research is warranted to reliably use these
scaffolds for understanding CSCs biology.
Kievit et al. reported the use of CA scaffold for proliferation and enrichment of
CD133+ glioblastoma stem cells (U-87 MG and U-118 MG cells) for understanding
CSC biology. The SEM imaging and immunostaining confirmed the selective
6 Alginate-Based Three-Dimensional In Vitro Tumor Models: A Better Alternative… 175
Fig. 6.5 Gene expression of CSC-related markers in HCCLM3 cells (a, c) and TCA8113 cells (b,
d). There are three groups of genes: Oct3/4, Nanog as self-renew-related genes; Notch-1, SMO as
stemness-related genes; CD44, EpCAM, CD133, and ABCG2 as CSC marker genes. β-Actin
mRNA was utilized as an internal control. *P < 0.05 and **P < 0.001, compared with 2D cells
(Reproduced with permission from Ref. [25])
Fig. 6.6 Morphology and von Willebrand factor (vWF) protein expression of xenografts formed by
2D- and 3D-cultured cells. HE-stained histology slides of HCCLM3 cells (a) and TCA8113 cells
(b) forming xenografts were morphologically similar to the HCC and HNSCC in vivo. Xenograft
sections formed by HCCLM3 (c) and TCA8113 (d) cells in both 2D monolayer and 3D ALG beads
were stained with anti-vWF antibody (green). Nuclei were stained with Hoechst 33342 (blue).
Scale bar 50 μm (Reproduced with permission from Ref. [25])
bioprinted scaffolds. The CSC marker Nestin was also found to be upregulated along
with enhanced levels of VEGF, a crucial angiogenesis promoter. The 3D-printed
CSCs showed higher chemoresistance to temozolamide compared to 2D model. The
bioprinting of cells for 3D culture may offer unique advantages like predesigned
product configuration, flexible size of hydrogel filament, printing of several cell
types simultaneously, and multilevel biomolecular gradient formation [48]. However,
3D bioprinting is a relatively new dimension of the well-known field of 3D bioprint-
ing and addressal of various questions associated with tumor microenvironment
needs to be thoroughly evaluated before this technique may be widely used.
Growing research evidence indicates that the CSCs also behave similarly to nor-
mal stem cells in terms of the microenvironment niche. Various systems were used
to mimic the CSC niche to promote their growth and proliferation. A recent study
with porous, tunable alginate hydrogels could mimic the CSC niche and promoted
enrichment of CSC of mouse 4T1 breast cancer model. Roles of different parame-
ters including mechanical properties of polymer, cytokine immobilization, and the
composition of ECM on CSC niche were investigated for their role on the prolifera-
tion. Interestingly, modulation of matrix stiffness (1%-190 Pa, 1.1%-210 Pa, 1.2%-
270 Pa, 1.3%-710 Pa, 1.4%-950 Pa, 1.5%-1070 Pa, and 1.6%-4700 Pa) led to
significant variation in the size of tumor spheroids, and continuous growth for longer
6 Alginate-Based Three-Dimensional In Vitro Tumor Models: A Better Alternative… 177
Fig. 6.7 Growth of CD133+ GBM cells on CA scaffolds. (a) SEM images comparing the morphol-
ogy and proliferation of human U-118 MG GBM cells cultured on 3D CA scaffolds and 2D mono-
layers over 15 days. Scale bar corresponds to 10 mm. (b) Comparison of the change in fraction of
CD133+ cells in U-118 MG cell population grown for 15 days on CA scaffolds or as monolayers.
Immunopositivity for CD133 was determined by flow cytometry. (c) CD133 mRNA content deter-
mined by real-time PCR in U-87 MG GBM cells grown for 10 days on CA scaffolds compared to
that of monolayer cultures. CD133 mRNA content was normalized to the monolayer condition.
deg) Immunostaining for CD133 (green) and SEM imaging of CA-scaffold cultured U-118 MG
GBM cells at day 10. The boxed regions in the top-row images correspond to the areas of the bot-
tom images. Blue color reflects DAPI counterstaining of nuclei. (d) Solitary U-118 MG cells
generally showed no CD133 staining. (e) Small clusters of U-118 MG cells showed faint CD133
staining. (f and g) Intensity of CD133 staining increases as clusters of U-118 MG cells grow larger.
Scale bars for panels d-g correspond to 50 μm for the upper row and 25 μm for the lower row
(Reproduced with permission from Ref. [46])
duration led to decrease in the spheroid colony numbers. Hydrogels with 950 Pa
(1.4% alginate) elasticity were found to promote the number of colonies, and 27±3
colonies were formed from day 3 to 7 (Fig. 6.8). Cytokines are essential for continu-
ous maintenance of CSC culture, and immobilization of epidermal growth factor
(EGF) and basic fibroblast growth factor (bFGF) was carried out on alginate hydro-
gels to support CSC proliferation and to produce a conducive niche for CSC enrich-
ment. The results showed better growth characteristics of spheroids grown on
immobilized cytokine hydrogels compared to control (no cytokines) and the solu-
tion group (cytokines added to the culture media). Further, the addition of low
178 A. Khurana and C. Godugu
Fig. 6.8 Tumor sphere formation in an alginate-based hydrogel with different degrees of stiffness.
(a) A single 4T1 breast cancer cell grew into tumor spheres in alginate hydrogels of different stiff-
ness levels throughout the course of culture from day 1 to day 7. Scale bar: 50 lm. (b) The observa-
tion of the cytoskeleton of the multicellular 4T1 tumor spheroid after 7 days in culture in an
alginate hydrogel with different levels of stiffness; the multicellular 4T1 tumor spheroid was
released from the hydrogel and was stained with phalloidin for the cytoskeleton (red) and DAPI for
the nucleus (blue); it was imaged with an inverted fluorescent microscope. Scale bar: 20 lm. (c)
Tumor sphere (round colony) number as a function of culture time: day 1 to day 7. The 950 Pa
alginate hydrogel seems to be optimal for sustaining the spheroid colony number. Mean ± SD;
n = 3 (for the 190 Pa, 270 Pa, 950 Pa, and 4700 Pa alginate hydrogels); independent experiments.
(d) Colony size of the tumor spheres as a function of culture time and hydrogel stiffness. Apparently,
the 950 Pa hydrogel best promotes tumor growth. Mean ± SD; n = 3 (for 190 Pa, 270 Pa, 950 Pa,
and 4700 Pa alginate hydrogels); independent experiments (Reproduced with permission from
Ref. [49])
molecular weight hyaluronic acid (HA) to alginate hydrogel was found to promote
the spheroid number. However, there were no significant changes in the spheroid
size, whereas use of high molecular weight HA was found detrimental to growth of
CSCs. The alginate scaffold enriched the 4T1 CSCs compared to 2D cultured cells
as revealed by the gene expression of Nestin, Tert, Nanog, and SOX2. Immunostaining
for CD44, CD24, MDR1, and Dclk1 also suggested superiority of alginate hydro-
gels. Moreover, the tumorigenicity study in Balb/c mice also revealed superiority of
the proposed hydrogels [49]. Thus, this unique alginate hydrogel-based platform
may provide a means to study the CSCs biology from preclinical and patient-derived
samples. However, to harness the real potential of such CSC-based platforms,
6 Alginate-Based Three-Dimensional In Vitro Tumor Models: A Better Alternative… 179
detailed studies in different lab experiments other than proof of concept assays
should be carried out to furnish robust, reliable, and reproducible results.
6.5 A
dvantages of Alginate-Based 3D Models Compared
to Other 3D Models
Cell culture is an important tool for understanding the physiological behavior of cell
pharmacodynamics. Primary cell culture and secondary cell culture serve to aid in
the basic and molecular pathology for direct and indirect clinical applications.
Three-dimensional cultures have significantly contributed in the overall progress of
cell culture techniques for drug discovery and development. Controlling the matrix
for better simulation of in vivo microenvironment is of utmost importance and needs
due consideration during designing of 3D systems. Although 3D cell culture models
have a plethora of applications, however, there are still certain areas where further
improvement is desired. Understanding of 3D complexity of tumor must be consid-
ered before selection and designing of such scaffolds as matrix interactions in terms
of the elasticity and reactivity may badly hamper the outcomes of 3D cell culture.
Optimization of 3D scaffold by using high-end techniques of 3D bioprinting may
advance the field and may help in growing multiple cell lines on a single bioprinted
sheet. Another area of improvement includes the development of 3D friendly imag-
ing techniques. The traditional 3D cultured cells are difficult to be imaged with the
regularly used inverted microscope as the density of tumor spheroid may not be
180 A. Khurana and C. Godugu
optimally reached. The periphery and the core of tumor spheroids contain different
regions of metabolism and bioactivity as the distance from nutrient supply dictates
the formation of hypoxia and the subsequent necrotic lesion formation. Another
dimension of 3D culture is the growth and proliferation of CSCs using special matri-
ces. Understanding the biology of CSCs requires an ideal matrix that can mimic the
tumor stroma interactions and at the same time may aid the selective proliferation of
CSC population. Last, but not the least, overcoming the economical barrier to trans-
late the wide-scale use of 3D scaffolds is of urgent need. Thus far, the techniques
used to synthesize 3D matrices are largely confined to tissue engineering and bio-
material laboratories. To promote the utility of 3D scaffolds for cancer drug discov-
ery, one needs to establish simple, robust, reproducible, economical, and
commercially viable products. The 3D scaffolds available in market are relatively
costly and out of reach of developing countries.
6.7 Summary
Three-dimensional cell cultures offer a versatile platform for different needs during
various phases of drug discovery and development. In this chapter, we summarized
a brief introduction of 3D cancer models and the advantages of 3D models over 2D
cell culture models. Studies confirm regarding better in vivo correlation of results
obtained with 3D cell cultures owing to better similarity to the in vivo 3D matrix.
Alginate is a versatile biomaterial with novel advantages like inertness, animal-free
product, no intrinsic bioactivity, and feasible chemical modifications for tailored
applications. AlgiMatrix™-based 3D scaffolds are discussed with unique feature of
matrix dissolution for further processing of the grown spheroids. A significant num-
ber of studies have been carried out with alginate-based matrices like 3D scaffolds,
alginate beads, alginate hydrogels, alginate–collagen complexes, and chitosan–algi-
nate–gelatin mixed scaffolds. These 3D scaffolds have been used for wide applica-
tions ranging from anticancer HTS drug screening, for understanding drug molecular
mechanisms, molecular pathology of various cancers, enrichment of cancer stem
cells, 3D bioprinting, and so on. In conclusion, alginate-based biomaterials can con-
tribute significantly to expand the horizon of 3D culture and may serve to cut down
the cost of commercial 3D anticancer screening platforms for wide-scale applica-
tion of 3D culture models in cancer research programs.
Acknowledgments The authors would like to thank the Department of Biotechnology (DBT),
Government of India for the financial support to Dr. CG via North East Twinning Grant,
MAP/2015/58; Indo-Brazil Grant, DBT/IC-2/Indo-Brazil/2016-19/01; and Science and
Engineering Board Early Career Research Award (SERB-ECR) Grant, ECR/2016/000007. In addi-
tion, the authors would also like to thank the Department of Pharmaceuticals, the Ministry of
Chemicals and Fertilizers, the Government of India, and the Project Director, NIPER-Hyderabad.
6 Alginate-Based Three-Dimensional In Vitro Tumor Models: A Better Alternative… 181
References
1. Sharma SV, Haber DA, Settleman J (2010) Cell line-based platforms to evaluate the therapeutic
efficacy of candidate anticancer agents. Nat Rev Cancer 10(4):241–253
2. Maltman DJ, Przyborski SA (2010) Developments in three-dimensional cell culture technol-
ogy aimed at improving the accuracy of in vitro analyses. Biochem Soc Trans 38(4):1072–1075
3. Smalley KS, Lioni M, Herlyn M (2006) Life isn’t flat: taking cancer biology to the next dimen-
sion. In Vitro Cell Dev Biol Anim 42(8-9):242–247
4. Hongisto V, Jernstrom S, Fey V et al (2013) High-throughput 3D screening reveals differ-
ences in drug sensitivities between culture models of JIMT1 breast cancer cells. PLoS One
8(10):e77232
5. Gillet JP, Calcagno AM, Varma S et al (2011) Redefining the relevance of established can-
cer cell lines to the study of mechanisms of clinical anti-cancer drug resistance. PNAS
108(46):18708–18713
6. Wrzesinski K, Fey SJ (2015) From 2D to 3D-a new dimension for modelling the effect of
natural products on human tissue. Curr Pharm Des 21(38):5605–5616
7. Kim JB (2005) Three-dimensional tissue culture models in cancer biology. Semin Cancer Biol
15(5):365–377
8. Pampaloni F, Reynaud EG, Stelzer EH (2007) The third dimension bridges the gap between
cell culture and live tissue. Nat Rev Mol Cell Biol 8(10):839–845
9. Breslin S, Driscoll LO (2013) Three-dimensional cell culture: the missing link in drug discov-
ery. Drug Discov Today 18(5):240–249
10. Misfeldt DS, Hamamoto ST, Pitelka DR (1976) Transepithelial transport in cell culture. PNAS
73(4):1212–1216
11. Santini M, Rainaldi G, Indovina P (1999) Multicellular tumour spheroids in radiation biology.
Int J Radiat Biol 75(7):787–799
12. Van Wezel A (1967) Growth of cell-strains and primary cells on micro-carriers in homoge-
neous culture. Nature 216:64–65
13. Hutmacher DW (2001) Scaffold design and fabrication technologies for engineering tissues
state of the art and future perspectives. J Biomater Sci Polym Ed 12(1):107–124
14. Bolz J, Novak N, Staiger V (1992) Formation of specific afferent connections in organotypic
slice cultures from rat visual cortex cocultured with lateral geniculate nucleus. J Neurosci
12(8):3054–3070
15. Nirmalanandhan VS, Duren A, Hendricks P et al (2010) Activity of anticancer agents in a
three-dimensional cell culture model. Assay Drug Dev Technol 8(5):581–590
16. Fitzgerald KA, Malhotra M, Curtin CM et al (2015) Life in 3D is never flat: 3D models to
optimise drug delivery. J Control Release 215:39–54
17. Godugu C, Patel AR, Desai U et al (2013) AlgiMatrix™ based 3D cell culture system as an
in-vitro tumor model for anticancer studies. PLoS One 8(1):e53708
18. Thoma CR, Zimmermann M, Agarkova I et al (2014) 3D cell culture systems modeling tumor
growth determinants in cancer target discovery. Adv Drug Deliv Rev 69-70:29–41
19. Xu X, Sabanayagam CR, Harrington DA et al (2014) A hydrogel-based tumor model for the
evaluation of nanoparticle-based cancer therapeutics. Biomaterials 35(10):3319–3330
20. Rohwer N, Cramer T (2011) Hypoxia-mediated drug resistance: novel insights on the func-
tional interaction of HIFs and cell death pathways. Drug Resist Updat 14(3):191–201
21. Imamura Y, Mukohara T, Shimono Y et al (2015) Comparison of 2D-and 3D-culture models as
drug-testing platforms in breast cancer. Oncol Rep 33(4):1837–1843
22. Lee KY, Mooney DJ (2012) Alginate: properties and biomedical applications. Prog Polym Sci
37(1):106–126
23. Chen L, Xiao Z, Meng Y et al (2012) The enhancement of cancer stem cell properties of MCF-7
cells in 3D collagen scaffolds for modeling of cancer and anti-cancer drugs. Biomaterials
33(5):1437–1444
182 A. Khurana and C. Godugu
24. Cardoso T, Sakamoto SS, Stockman D et al (2016) A three-dimensional cell culture system as
an in vitro canine mammary carcinoma model for the expression of connective tissue modula-
tors. Vet Comp Oncol 2016:1–9
25. XX X, Liu C, Liu Y et al (2014) Enrichment of cancer stem cell-like cells by culture in alginate
gel beads. J Biotechnol 177:1–12
26. Nakaoka R, Hirano Y, Mooney DJ et al (2013) Study on the potential of RGD-and PHSRN-
modified alginates as artificial extracellular matrices for engineering bone. J Artif Organs
16(3):284–293
27. Hahn MS, Teply BA, Stevens MM et al (2006) Collagen composite hydrogels for vocal fold
lamina propria restoration. Biomaterials 27(7):1104–1109
28. Akeda K, Nishimura A, Satonaka H et al (2009) Three-dimensional alginate spheroid culture
system of murine osteosarcoma. Oncol Rep 22(5):997–1003
29. Leung M, Kievit FM, Florczyk SJ et al (2010) Chitosan-alginate scaffold culture sys-
tem for hepatocellular carcinoma increases malignancy and drug resistance. Pharm Res
27(9):1939–1948
30. Kievit FM, Florczyk SJ, Leung MC et al (2010) Chitosan–alginate 3D scaffolds as a mimic of
the glioma tumor microenvironment. Biomaterials 31(22):5903–5910
31. Florczyk SJ, Liu G, Kievit FM et al (2012) 3D porous chitosan–alginate scaffolds: a new
matrix for studying prostate cancer cell–lymphocyte interactions in vitro. Adv Healthc Mater
1(5):590–599
32. Wang K, Kievit FM, Florczyk SJ et al (2015) 3D porous chitosan–alginate scaffolds as an
in vitro model for evaluating nanoparticle-mediated tumor targeting and gene delivery to pros-
tate cancer. Biomacromolecules 16(10):3362–3372
33. Ginzberg DM, Konson A, Cohen S et al (2007) Entrapment of retroviral vector producer cells
in three-dimensional alginate scaffolds for potential use in cancer gene therapy. J Biomed
Mater Res B Appl Biomater 80(1):59–66
34. Wang JZ, Zhu YX, Ma HC et al (2016) Developing multi-cellular tumor spheroid model
(MCTS) in the chitosan/collagen/alginate (CCA) fibrous scaffold for anticancer drug screen-
ing. Mater Sci Eng C 62:215–225
35. Wenzel C, Riefke B, Gründemann S et al (2014) 3D high-content screening for the iden-
tification of compounds that target cells in dormant tumor spheroid regions. Exp Cell Res
323(1):131–143
36. Meli L, Jordan ET, Clark DS et al (2012) Influence of a three-dimensional, microarray envi-
ronment on human cell culture in drug screening systems. Biomaterials 33(35):9087–9096
37. XX X, Liu C, Liu Y et al (2013) Encapsulated human hepatocellular carcinoma cells by algi-
nate gel beads as an in vitro metastasis model. Exp Cell Res 319(14):2135–2144
38. Shakibaei M, Kraehe P, Popper B et al (2015) Curcumin potentiates antitumor activity of
5-fluorouracil in a 3D alginate tumor microenvironment of colorectal cancer. BMC Cancer
15(1):1–15
39. Buhrmann C, Shayan P, Kraehe P et al (2015) Resveratrol induces chemosensitization to
5-fluorouracil through up-regulation of intercellular junctions, Epithelial-to-mesenchymal
transition and apoptosis in colorectal cancer. Biochem Pharmacol 98(1):51–68
40. Buhrmann C, Shayan P, Popper B et al (2016) Sirt1 is required for resveratrol-mediated che-
mopreventive effects in colorectal cancer cells. Nutrients 8(3):145–166
41. Malarvizhi GL, Retnakumari AP, Nair S et al (2014) Transferrin targeted core-shell nanomedi-
cine for combinatorial delivery of doxorubicin and sorafenib against hepatocellular carcinoma.
Nanomedicine 10(8):1649–1659
42. Lan SF, Mroczka BS, Starly B (2010) Long-term cultivation of HepG2 liver cells encapsulated
in alginate hydrogels: a study of cell viability, morphology and drug metabolism. Toxicol in
Vitro 24(4):1314–1323
43. Zhou S, Li F, Xiao J et al (2010) Isolation and identification of cancer stem cells from human
osteosarcom by serum-free three-dimensional culture combined with anticancer drugs.
J Huazhong Univ Sci Technol Med Sci 30:81–84
6 Alginate-Based Three-Dimensional In Vitro Tumor Models: A Better Alternative… 183
Zhengfan Xu and Mai T. Lam
Abstract Alginate biomaterial has been extensively investigated and used for many
biomedical applications due to its biocompatibility, low toxicity, relatively low cost,
and ease of use. Its use toward cardiovascular application is no exception. Alginate
is approved by the Food and Drug Administration (FDA) for various medical
applications, such as a thickening, gel forming, and as a stabilizing agent for dental
impression materials, wound dressings, and more. In this chapter, we describe the
versatile biomedical applications of alginate, from its use as supporting extracel-
lular matrices (ECM) in patients after acute myocardial infarction (MI), to its
employment as a vehicle for stem cell delivery, to controlled delivery of multiple
combinations of bioactive molecules. We also cover the application of alginate in
creating solutions for treatment of other cardiovascular diseases by capitalizing on
the natural properties of alginate to improve creation of heart valves, blood vessels,
and drug and stem cell delivery vehicles.
7.1 Introduction
Alginate is a naturally occurring anionic polymer obtained from brown seaweed and
has gained popularity in a wide range of biomedical applications. Alginate has many
beneficial properties conducive to cardiovascular applications due to its relatively low
reactivity in vivo, low cost, non-thrombogenic nature, mild gelation process, and
structural resemblance to the ECM [1–4]. Two injectable alginate implants have
already reached clinical investigation phase, demonstrating the promising potential
of alginate-based treatments for myocardial repair and regeneration [4–8].
The more recent evolution in cardiac application of alginate has now led to new
standards in biomaterial application in the cardiac system to not only “fill the gap”
in the injured area, but to act as an interface with the cardiac biological systems as
well [9]. Such applications focus on four major areas: (1) using alginate hydrogels
as an ECM substitute in heart tissues to promote tissue regeneration due to the struc-
tural similarity between alginate and natural heart ECM, (2) using alginate hydro-
gels as a delivery vehicle for cardiac stem cells or adult cardiomyocytes to the injury
site to facilitate regeneration of functional heart tissue, (3) using alginate as a plat-
form for sustained delivery of growth factors to mimic natural physiology, and (4)
using alginate gels to control drug release. As a drug release vehicle, an alginate-
based system can be fine-tuned to control the speed of cardiac medicine release
(such as with antihypertensive or antiarrhythmia medications) by controlling the
cross-linker type and the cross-linking method.
In this chapter, we present an up-to-date view of alginate-based applications in
the treatment of cardiovascular diseases, with emphasis on cardiac tissue engineer-
ing, myocardial regeneration, as well as their translational status and clinical
advancement.
7.2.1 M
yocardial Infarction: Inflammation, Wound Healing,
and Remodeling of the Left Ventricle
Fig. 7.1 Summary of alginate application in myocardial infarction. Alginate has been used as an extracellular matrix (ECM) replacement, for cell delivery and
as a growth factor delivery vehicle
187
188 Z. Xu and M.T. Lam
LV dilatation is enlargement of the LV wall, creating a stretched and thin heart wall.
Decreasing the degree of LV dilatation is of utmost importance, as it is considered
to be one of the most important predictors of mortality in patients with MI.
7.2.2.1 A
lginate Hydrogel for Treating MI and Congestive Heart Failure
Patients
In a study by Yu et al., injections of fibrin or alginate gel were made directly into
infarcted myocardium 5 weeks after induced infarction in rats [21]. Echocardiographic
results at 5 weeks after injection of alginate demonstrated persistent improvement
of left ventricular fractional shortening and prevention of continued enlargement of
left ventricular dimensions indicating dilatation, whereas a comparison of fibrin
glue demonstrated no progression of left ventricular negative remodeling. Following
that study, the same research group conducted in vitro cell culture and in vivo rat
studies to evaluate efficacy of RGD peptide-conjugated alginate hydrogel [22].
RGD peptide-conjugated alginate improved human umbilical vein endothelial cell
(HUVEC) proliferation and adhesion in culture. Injection of the alginate hydrogel
into the infarct area of rats 5 weeks post-MI demonstrated that while alginate hydro-
gel could reshape a dilated aneurysmal LV and lead to improved LV function, a
RGD modified alginate enhanced angiogenesis. Both modified and non-modified
alginate improved heart function, while LV function in the control group deterio-
rated. Both the RGD modified alginate and non-modified alginate increased the
arteriole density compared to control, with the RGD modified alginate resulting in
the greatest angiogenic response. These results suggest that in situ use of modified
polymers may influence the tissue microenvironment and serve as potential thera-
peutic agents for patients with chronic heart failure.
Next, Sabbah et al. tested the efficacy of Algisyl-LVR™ in dogs with heart failure
(HF) [23]. The heart failure was induced by repetitive coronary microembolization in
dogs. The final injection mixture was prepared by combining sodium-alginate aque-
ous solution mixed with calcium cross-linked alginate hydrogel. The material was
applied at least 2 weeks after the last coronary microembolization through 7 injec-
tions (0.25–0.35 ml each) directly into the LV wall. Injections were made circumfer-
entially along the LV mid-ventricular level halfway between the apex and base. The
treatment was well tolerated, and ambulatory 24 h Holter electrocardiography
showed no significant differences between groups with respect to heart rate and
frequency and severity of ventricular arrhythmias. Seventeen weeks post treatment,
histological analysis showed that pockets of the implant material were still within
the LV free wall and were encapsulated by a thin layer of connective tissue with no
evidence of inflammation. Compared to saline-treated animals, the alginate implant
significantly reduced LV end-diastolic and end-systolic volumes and end-diastolic
pressure, improved LV sphericity, and increased wall thickness. The treatment also
significantly increased ejection fraction (EF) from ~26% at baseline to ~31% after
treatment, compared to decreased EF from~27% at baseline to ~24% after saline
injection seen in control dogs [23]. These promising results led to clinical trials.
190 Z. Xu and M.T. Lam
Lee et al. combined coronary artery bypass grafting (CABG) with Algisyl-LVR™
injection to treat three congestive heart failure patients (NYHA class 3) [24]. Magnetic
resonance images obtained before treatment (n = 3) and at 3 months (n = 3) and
6 months (n = 2) afterward were used to reconstruct the LV geometry. The LV became
more ellipsoidal after treatment, and both end-diastolic volume (EDV) and end-sys-
tolic volume (ESV) decreased substantially 3 months after treatment in all patients.
Ejection fraction increased from 32 ± 8% to 47 ± 18% during that period. Volumetric-
averaged wall thickness increased in all patients. These changes were accompanied by
about a 35% decrease in myofiber stress at end-diastole and at end-systole. Although
the statistical power of this finding is limited by the small sample size, the concept
of this treatment is novel and the effects are interesting and remarkable.
In a study by Lee et al., eleven male patients aged 44–74 years with advanced
heart failure (NYHA class 3 or 4), a left ventricular ejection fraction (LVEF) of <
40%, and a need for conventional heart surgery received Algisyl-LVR delivered into
the LV myocardial free wall [25]. Serial echocardiography, assessment of NYHA
class, Kansas City Cardiomyopathy Questionnaire (KCCQ), and 24-hour Holter
monitoring were obtained at baseline; 3–8 days (for echocardiography and Holter
monitoring); and at 3, 6, 12, 18, and 24 months. A total of 9 (81.8%) patients com-
pleted 24 months of follow-up. There were no adverse events attributed to the use of
Algisyl-LVR. Improvement in LV function was evidenced by mean LVEF, improved
NYHA class, as well as improved KCCQ summary scores. Holter monitor data
showed a significant decrease in the episodes of non-sustained ventricular tachycar-
dia following administration of Algisyl-LVR. This first-in-man study confirmed fea-
sibility and safety of administering Algisyl-LVR to patients with heart failure at the
time of open-heart surgery. Thus, a randomized controlled study to evaluate Algisyl-
LVR™ solely as a method of LV augmentation for patients with severe heart failure
(AUGMENT-HF) has been initiated and should provide definitive conclusions
about the clinical benefits of Algisyl-LVR™ therapy [26].
Recently, Rocca et al. developed a novel hydrogel composed of gelatinized algi-
nate with parallel capillary-like channels, termed Capgel [27]. To make this hydro-
gel, a solution of 2% (w/v) alginate and 2.6% (w/v) oligo-gelatin was poured into
an alginate-coated glass petri dish. The mixture was then submerged in a tank of
0.5 M CuSO4 and undisturbed for at least 36 h. The gel was then rinsed and cross-
linked using N-ethyl-N′-(3-dimethylamino propyl) carbodiimide hydrochloride
and N-hydroxysulfosuccinimide. The modulus of elasticity of the resultant mate-
rial was around 4 kPa, above the baseline modulus shown to support cardiomyo-
cyte function (1–3 kPa, [28–31]) and below the fibrotic modulus shown to induce
dysfunction (40–70 kPa). Angiotensin-(1–7) is a small peptide previously shown to
be cardioprotective for infarcted myocardium, although the main disadvantage is
its short half-life. Angiotensin was incorporated into Capgel to sustain local mol-
ecule release, prolong molecule bioactivity, and reduce off-target complications.
Forty-eight hours after induced anterior MI, Sprague Dawley rats received intra-
myocardial injection of Capgel directly into the anteroseptal wall at the infarct
border zone or no injection. Echocardiograms were performed at 48 h and 4 weeks
to evaluate left ventricular function. Echocardiograms showed improvement of left
7 Alginate Application for Heart and Cardiovascular Diseases 191
ventricular systolic function over time with gel injection. Histologically, the
authors observed that: (1) Capgel was present at the injection site after 4 weeks but
was minimal at 8 weeks, (2) the remaining gel was populated with blood vessels to
promote regeneration of cardiomyocytes, and (3) the remaining gel was heavily
populated by CD 68+ macrophages with CD 206+ clusters. CD206+ macrophages
are known as M2 (alternative) activation CD 206+ macrophages. This activation is
more consistent with a regenerative healing response rather than the classic inflam-
matory response [32]. An in vitro experiment was performed and demonstrated.
Angiotensin-(1–7) was released from the Capgel in a sustained manner for 90 days.
Lastly, the study showed intramyocardial injection of Capgel did not result in
arrhythmic events, and all animals survived after injection. Capgel injection did
not change systolic or diastolic LV function indices at 4 weeks compared with the
saline control group, indicating that the gel appeared safe for this intraventricular/
intramyocardial application.
7.2.3 A
lginate Hydrogel as a Vehicle for Cell Delivery
into Infarcted Tissue
Given its innate physical properties, alginate hydrogel could significantly improve
cell retention, survival, and function by serving as an artificial biomimetic
ECM. Alginate can provide the required temporal support for cell growth and func-
tion until the cells are able to support themselves with their own cell-secreted ECM
[40]. The aid of a temporary scaffold is critically important in conditions to which
transplanted cells are presented, that is, the “hostile” environment of an injury site,
such as in ischemic myocardium. Of note, poor cell retention is likely to be a major
factor responsible for inconsistent and limited efficacy of cell transplantation stud-
ies for MI treatment thus far [41, 42]. Roche et al. compared four biomaterial deliv-
ery vehicles for improving mesenchymal stem cell (MSC) retention in a rat MI
model – two hydrogels (RGD-modified alginate and chitosan/β-glycerophosphate)
and two epicardial patches (RGD-modified alginate and collagen) [43]. From 0 to
24 h, an average of 50–62% cell retention in all biomaterial-treated hearts is seen,
comparing to 9% cell retention in saline control-treated hearts. At the same time, no
significant difference was observed between the individual biomaterials.
Alginate hydrogel is most commonly used for intramyocardial delivery of MSCs.
Alginate is used to increase cell retention and enhance cell-mediated myocardial
repair. In a study by Levit et al., human bone marrow MSCs were encapsulated in
alginate hydrogel and then attached to the infarcted rat heart in a biocompatible
poly(ethyleneglycol) (PEG) hydrogel patch to secure the encapsulated cells [44].
Transthoracic echocardiography and cardiac magnetic resonance imaging showed
significantly improved cardiac function in encapsulated hMSC-treated animals,
associated with reduced scar size and increased microvascular density, compared to
directly injected cells. Bioluminescence imaging confirmed that encapsulation
resulted in significantly greater hMSC retention over a period of 7 days.
A macroporous alginate scaffold, commercially available from Life Technologies,
Inc. as the AlgiMatrix™ 3D Cell Culture System, was developed by the Cohen
group [45, 46]. This scaffold is created by controlled freeze drying of calcium cross-
linked alginate solution. The porous structure of the scaffold is dependent on the
freezing protocol, namely, parameters of rate and spatial direction of the tempera-
ture gradient. When calcium cross-linked alginate solutions are slowly frozen at
−20 °C in a nearly homogeneous cold atmosphere, the resultant scaffold has an
isotropic pore structure, creating pores that are spherical and interconnected.
Conversely, when the cooling process is performed under a unidirectional tempera-
ture gradient along the freezing solution, an anisotropic pore structure is attained.
Alginate scaffolds are characterized by an interconnected pore structure with
large pore size (50–200 μm in diameter) and high matrix porosity (70–90%) [45,
46]. The pore size in scaffolds should ideally be at least 50 μm in diameter to allow
high mass transport during in vitro culture and enable vascularization after implan-
tation. Macroporous solid 3D scaffolds are an ideal platform for in vitro cardiac
tissue bioengineering and creation of 3D tissue grafts, as they are capable of com-
7 Alginate Application for Heart and Cardiovascular Diseases 195
7.2.4 A
lginate Hydrogel as Vehicle for Delivery of Growth
Factors into Infarcted Tissue
The use of bioactive molecules (growth factors, cytokines, and stem cell-mobilizing
factors) has always been of interest in the field of therapeutic myocardial regenera-
tion. The effect exerted by these molecules includes cell proliferation, vasculariza-
tion apoptosis inhibition, progenitor cell migration, and progenitor cell differentiation
[12, 55–60]. However, systemic cytokine or growth factor administration is
currently considered unpractical because of several reasons. First, systemic admin-
istration is associated with numerous safety concerns including elevated blood
pressure, increased incidence of restenosis, thrombolytic events, arrhythmia, etc.
[61]. Secondly, systemic administration requires high doses of the protein due to
extremely low protein stability in the circulation. Lastly, most of these molecules
have pleiotropic functions. Thus, a careful, local, and time-adjusted intervention is
needed for administration. Biomaterials could be engineered to produce a sustained
delivery system for bioactive molecule combinations. In such a system, the bioma-
terial will provide structural temporary matrix support and direct the formation of
functional tissue in situ. Simultaneously, it will serve as a temporary depot for sus-
tained delivery and presentation of bioactive molecules with spatial and controlled
distribution of the desired agent to induce multiple reparative/regenerative pro-
cesses. In general, due to low protein adsorption and its highly porous and hydro-
philic nature, the pristine unmodified alginate matrix yields fast and unpredictable
7 Alginate Application for Heart and Cardiovascular Diseases 197
protein release kinetics. Thus, several modification designs and engineering schemes
were applied for the use of alginate in protein delivery.
Hao et al. used alginate consisting of both high and low molecular weight (MW)
components, otherwise known as binary molecular weight alginate. Increasing the
molecular weight in alginate gel leads to enhanced mechanical properties in resultant
gels, although the gel becomes more difficult to be cleared by the body [62]. The
binary MW alginate also supplies a partially oxidized formulation for sequential
delivery of vascular endothelial growth factor (VEGF) and platelet-derived growth
factor (PDGF)-BB into the infarcted myocardium. The factors were adsorbed to the
hydrogel via electrostatic interaction, and the sequential factor delivery was achieved
due to the varying degradation rates of the partially oxidized alginates constituting
the hydrogel. One week after MI was induced in rats, the modified alginate hydrogels
loaded with VEGF and PDGF-BB were injected intramyocardially along the border
zone of the infarct. Four weeks after injection, the sequential growth factor release
led to a higher density of α-SMA-positive (mature) vessels compared to the delivery
of single factors. Sequential delivery of VEGF and PDGF-BB increased the systolic
velocity-time integral compared to delivery of VEGF alone.
Alginate has also been used as a component in composite carriers. Banquet et al.
developed cross-linked albumin-alginate microcapsules that sequentially released
fibroblast growth factor-2 (FGF-2) and hepatocyte growth factor (HGF) [63]. The
developed microcapsules (~100 μm in diameter) contained a thin, covalently cross-
linked human serum albumin and propylene glycol alginate membrane surrounding
a liquid center. Both proteins were bound to a microcapsule surface layer, and FGF-2
was also found in the liquid center. The microcapsules displayed differential release
profiles for FGF-2 (released first) and HGF (release delayed for >1 week) in vitro. In
a rat model of MI, immediate post-MI intramyocardial injection of FGF-2/HGF algi-
nate-albumin microcapsules stimulated angiogenesis and arteriogenesis and pre-
vented cardiac hypertrophy and fibrosis, as determined by immunohistochemistry.
Cardiac perfusion was improved after 3 months, as shown by magnetic resonance
imaging. These multiple beneficial effects resulted in reduced adverse cardiac remod-
eling and improved left ventricular function, as shown by echocardiography.
Ruvinov et al. established an alginate-based scaffold which binds to various bio-
active molecules, collectively known as heparin-binding proteins [64]. To enable
this process, sulfation of the uronic acid monomers on alginate was performed,
presenting affinity-binding sites for heparin-binding proteins. The sequential deliv-
ery of different growth factors was achieved by (1) the varied equilibrium binding
constants of the modified alginate as determined by surface plasmon resonance
(SPR) analysis and (2) manipulating the initial amount of factors added to the sys-
tem. The research group chose insulin-like growth factor (IGF)-1 and HGF for the
injectable alginate system. These two growth factors have established and compli-
mentary beneficial effects on infarcted myocardium [61, 65–67]. Biologically active
IGF-1 followed by HGF proteins were sequentially released from the alginate
hydrogel. This sequential delivery system of IGF-1 and HGF could be applied for
the proper execution of reparative processes in the infarcted myocardium and for
achieving a more favorable course of repair. The faster released IGF-1 can provide
198 Z. Xu and M.T. Lam
cardiac water indicating there is less edema in alginate particle-treated hearts. The
IHC staining at week 3 post MI showed less CD 68+ macrophage at infarcted area.
Picrosirius Red stain at the same time showed less collagen formation at the same
area. These results indicated cardiac inflammation and fibrosis were attenuated.
Finally, magnet resonance (MR) and ultrasound were conducted at week 6 post MI
to evaluate cardiac function. MR showed increased cardiac perfusion in the alginate
particle-treated group compared with the untreated control. Ultrasound showed sig-
nificant increase in function of the treated rat hearts compared to the control. This
study showed that, despite the endogenous cardiac lymphangiogenic response post-
MI, the remodeling and dysfunction of collecting ducts contribute to the develop-
ment of chronic myocardial edema and inflammation-aggravating cardiac fibrosis
and dysfunction. Moreover, the study revealed that therapeutic lymphangiogenesis
may be a promising new approach for the treatment of cardiovascular diseases.
Chang et al. produced genome-modified mesenchymal stem cells that secreted
HGF upon drug-specific induction [72]. The modified MSCs were then integrated
in arginine-glycine-aspartic acid (RGD)-alginate microgel. This microgel was then
transplanted into a hindlimb ischemia model in rats. First, a TetOn-HGF-expression
construct was generated. The drug-inducible HGF expression was evidenced by
transiently transfecting the construct into HEK293t cells. Then, the TetOn-HGF-
expression construct was integrated into a safe harbor site in an MSC chromosome
using the transcription activator-like effector nuclease (TALEN) system, resulting
in the production of TetOn-HGF/human umbilical cord blood-derived (hUCB)-
MSCs. Gel migration test of the TetOn-HGF/hUCB-MSCs showed that they had
enhanced mobility upon the induction of HGF expression. Moreover, long-term
induction of HGF expression in TetOn-HGF/hUCB-MSCs enhanced the antiapop-
totic responses of these cells subjected to oxidative stress. Long-term HGF induc-
tion in the same cells also improved the tube formation ability evidenced by a
matrigel tube forming assay. Lastly, TetOn-HGF/hUCB-MSCs encapsulated by
RGD-alginate microgel induced to express HGF improved in vivo angiogenesis in
a mouse hindlimb ischemia model. After one week of induced hindlimb ischemia,
mice were treated with phosphate-buffered saline, UCB-MSCs only, hrHGF only,
an RGD-alginate microgel containing UCB-MSCs, an RGD-alginate microgel con-
taining HGF integrated UCB-MSC with sustained inducing of HGF. After treat-
ment, the levels of blood perfusion of the hindlimbs were measured using a
Laser-Doppler flowmeter weekly for 4 weeks. Blood flow was significantly higher
in the microgel group, and a statistically significant difference was observed between
the microgel group and other groups. The IHC stain of the ischemic limb also
showed positive stain for von Willebrand factor in microgel group, further confirm-
ing angiogenesis in the lesion. In this study, by genetically modifying mesenchymal
stem cells, the researchers achieved sustained secretion of HGF and increased the
viability and migration ability of the MSCs. This system also overcomes the issue
of the short half-life of HGF, a major limitation for the usage of recombinant
HGF. Thus, the MSCs that express HGF in an inducible manner are a useful thera-
peutic modality for the treatment of vascular diseases requiring angiogenesis.
7 Alginate Application for Heart and Cardiovascular Diseases 201
7.3 A
pplication of Alginate in Other Cardiovascular
Diseases
Beyond myocardial infarction, alginate has also been used in potential treatments
for other cardiovascular diseases, as summarized in Fig. 7.2 and detailed in the fol-
lowing sections.
7.3.1 T
hree-Dimensional (3D) Printed Aortic Valve and Its
Potential Application in Aortic Valve Disease
Fig. 7.2 Summary of alginate application in other cardiovascular diseases. Alginate has been explored for its utility as a bio-ink in 3D printing approaches, in
the fabrication of blood vessels, in drug delivery systems, and in other applications for potential treatment in cardiovascular disease
Z. Xu and M.T. Lam
7 Alginate Application for Heart and Cardiovascular Diseases 203
7.3.2 F
abrication of Bioengineered Blood Vessels Using
Alginate-Based Methods
7.3.3 A
lginate-Based Drug Delivery Systems in Cardiovascular
Diseases
Alginate gels have been investigated for the delivery of a variety of low molecular
weight drugs, including drugs used to treat cardiovascular diseases. Alginate-
chitosan systems are often used to decrease release of medicine into the stomach. In
one study by Kevadiya et al., antiarrhythmia procainamide (PA) was intercalated
into the interlayer of montmorillonite (MMT) via an ion exchange mechanism [78].
The prepared PA–MMT composite was then compounded within an alginate (AL)
and chitosan (CS) complex. The release performance of PA was found to be delayed
in the PA-MMT-AL/CS composite in the gastric environment compared to the
PA-MMT composite. Release of the drug in the PA-MMT-AL/CS composite in the
intestinal environments exhibited a controlled manner.
Alginate-based local drug delivery systems are being tested on variant applica-
tions. In the Beckerman et al. study, alginate-based glue (SEAlantis) with amioda-
rone was applied pericardially to the right atrium [79]. Rapid atrial response (RAR,
an abnormal rhythm associated with atrial fibrillation) to burst pacing was assessed
before application and in the third postoperative day (POD3). The results yielded a
7 Alginate Application for Heart and Cardiovascular Diseases 205
perfusion is a plausible mechanism of both drug distribution and clearance from the
myocardium. These critical insights suggest that inotropic compounds with physi-
cochemical properties that lend themselves to tissue retention may be better suited
for epicardial applications. The demonstrated local myocardial pharmacokinetics
may allow for practical designs for epicardial drug therapy systems.
7.3.4 A
lginate-Based Stem Cell Delivery in Reverting
Doxorubicin-Related Cardiomyopathy
A study by Liu et al. showed that encapsulation of cardiac stem cells (CSCs) in
superoxide dismutase (SOD)-loaded alginate hydrogel prevents doxorubicin
(DOX)-mediated toxicity in vitro [84]. The study was based on the concept that
DOX-induced cardiotoxicity can be viewed as a stem cell disease, whereby the
formation of reactive oxygen species (ROS) by DOX is seen to predominantly hin-
der regenerative capability. Increased cell survival was confirmed by fluorescent
microscopy and assays measuring metabolic activity, cell viability, cytotoxicity, and
apoptosis. Encapsulation of CSCs in alginate alone failed to prevent apoptosis in
DOX-conditioned cell culture medium while encapsulation in SOD-loaded alginate
reduced apoptosis to near-normal levels, while metabolic activity was returned to
baseline. Although this study explored the protective effects of CSC encapsulation
against DOX, this technology may also be successfully implemented in the treat-
ment of other diseases in which sustained oxidative stress contributes to pathology.
plasma atrial natriuretic peptide (ANP) content, heart and renal weight indices, and
decreased plasma aldosterone (ALD). The authors claimed L-PA offered a novel
form of potassium supplementation with greater antihypertensive and sodium excre-
tion actions than KCl.
Alginate has demonstrated great utility and potential as a biomaterial for use in
cardiovascular diseases, particularly in the applications such as ECM replacement,
3D microenvironment design for functional cardiac tissue formation, stem cell
delivery, and controlled release and presentation of multiple combinations of bioac-
tive molecules and regenerative factors. The most attractive features of alginate for
these applications include biocompatibility, mild gelation conditions, and simple
modifications to prepare alginate derivatives with new properties. Moreover, based
on encouraging preclinical data, acellular injectable alginate implants for myocar-
dial repair and tissue reconstruction have already reached the clinical investigation
phase in MI and HF patients.
In the near future, the use of alginate-based materials in cardiovascular diseases is
likely to evolve considerably. However, several challenges and needs have still not
been fully addressed. First, the design of a more cell-interactive biomaterial is required.
For example, RGD peptides have been extensively exploited to date as a cell adhesion
ligand. However, multiple other ligands (such as heparin binding proteins) may be
required to properly produce replacement tissues in cardiac diseases. One or more
bioactive agents can be incorporated into alginate hydrogel to facilitate cardiac regen-
eration, as these gels have demonstrated utility in maintaining local concentrations of
biological factors, such as proteins, for extended time periods. Introduction of multi-
ple signaling ligands and topographical cues and control over their spatial distribution
at the nanoscale are the key variables in such design process.
Next, the development of “smart” alginate-based release systems will continue,
and this will potentially improve safety, increase sustained and local effects, and
provide new therapies. Precise control over the delivery of multiple drug combina-
tions, spanning multiple drug families (proteins, nucleic acids, small molecules) of
sustained vs. sequential release in response to external environmental changes
(mechanical signals, electromagnetic fields, changes in cell populations, ECM
degradation, etc.) is highly desirable.
It is clear that biomaterials combined with adult stem cells may become an
optimal and effective myocardial recovery and usually show synergistic effects that
play an important role in recovery progress of the infarcted myocardium. Results
show that even the small remaining numbers of transplanted stem cells may lead to
restoration of the infarcted myocardium. The majority of researchers believe it is
due to the paracrine effects through which more CSCs are recruited from the host
tissue. Subsequently, the CSCs undergo differentiation, restoring the cardiomyocyte
compartment. The lack of true regeneration in terms of increasing the fraction of
208 Z. Xu and M.T. Lam
contractile units is one of the main reasons that limit the success in clinical trials of
stem cell transplantation. When alginate hydrogel is used in cell encapsulation strat-
egies, the commonly used covalent cross-linking reactions can cause toxicity to the
encapsulated cells, making the appropriate choice of cell-compatible chemical
reagents (e.g., initiator), and thorough removal of unreacted reagents and by-prod-
ucts necessary. Future research may focus on the following areas: (1) design of a
smart hydrogel that can degrade upon an activation signal or at a specific pH, (2)
optimization of the hydrogel’s properties to improve its resistance to cardiac cyclic
loading movement, and (3) progressing in vivo studies to large animal models to
create more relevancy to humans.
In the future, we expect the evolution of combinatory and more complex strate-
gies, where combinations of the abovementioned approaches will be used together,
to achieve more powerful and synergistic effects on tissue repair and regeneration.
References
1. Ueno M, Oda T (2014) Chapter six – biological activities of alginate. Adv Food Nutr Res
72:95–112
2. Laurienzo P (2010) Marine polysaccharides in pharmaceutical applications: an overview. Mar
Drugs 8:2435–2465
3. Senni K et al (2011) Marine polysaccharides: a source of bioactive molecules for cell therapy
and tissue engineering. Mar Drugs 9:1664–1681
4. Andersen T, Strand BL, Formo K, Alsberg E, Christensen BE (2012) Chapter 9 alginates as
biomaterials in tissue engineering. Carbohydr Chem 37:227–258
5. Alsberg E, Anderson KW, Albeiruti A, Franceschi RT, Mooney DJ (2001) Cell-interactive
alginate hydrogels for bone tissue engineering. J Dent Res 80:2025–2029
6. Boateng JS, Matthews KH, Stevens HN, Eccleston GM (2008) Wound healing dressings and
drug delivery systems: a review. J Pharm Sci 97:2892–2923
7. Rabbany SY et al (2010) Continuous delivery of stromal cell-derived factor-1 from alginate
scaffolds accelerates wound healing. Cell Transplant 19:399–408
8. Murakami K et al (2010) Hydrogel blends of chitin/chitosan, fucoidan and alginate as healing-
impaired wound dressings. Biomaterials 31:83–90
9. Williams DF (2009) On the nature of biomaterials. Biomaterials 30:5897–5909
10. Martin GSJS, Norman S (2000) Left ventricular remodeling after myocardial infarction patho-
physiology and therapy. Circulation 101:2981–2988
11. Westman PC et al (2016) Inflammation and remodeling after myocardial infarction. J Am Coll
Cardiol 67:2050–2060
12. Ruvinov E, Sapir Y, Cohen S (2012) Cardiac tissue engineering: principles, materials, and
applications. Morgan & Claypool Publishers, San Rafael
13. Vunjak-Novakovic G, Lui KO, Tandon N, Chien KR (2011) Bioengineering heart muscle: a
paradigm for regenerative medicine. Annu Rev Biomed Eng 13:245–267
14. Hirt MN, Hansen A, Eschenhagen T (2014) Cardiac tissue engineering: state of the art. Circ
Res 114:354–367
15. Akhyari P, Kamiya H, Haverich A, Karck M, Lichtenberg A (2008) Myocardial tissue engi-
neering: the extracellular matrix. Eur J Cardiothorac Surg 34:229–241
16. Dobaczewski M, Gonzalez-Quesada C, Frangogiannis NG (2010) The extracellular matrix as a
modulator of the inflammatory and reparative response following myocardial infarction. J Mol
Cell Cardiol 48:504–511
7 Alginate Application for Heart and Cardiovascular Diseases 209
40. Segers VF, Lee RT (2011) Biomaterials to enhance stem cell function in the heart. Circ Res
109:910–922
41. Templin C, Luscher TF, Landmesser U (2011) Cell-based cardiovascular repair and regenera-
tion in acute myo-cardial infarction and chronic ischemic cardiomyopathy – current status and
future developments. Int J Dev Biol 55:407–417
42. Singelyn JM, Christman KL (2010) Injectable materials for the treatment of myocardial
infarction and heart failure: the promise of decellularized matrices. J Cardiovasc Transl Res
3:478–486
43. Roche ET et al (2014) Comparison of biomaterial delivery vehicles for improving acute reten-
tion of stem cells in the infarcted heart. Biomaterials 35:6850–6858
44. Levit RD (2013) Cellular encapsulation enhances cardiac repair. J Am Heart Assoc 2:e000367
45. Shapiro L, Cohen S (1997) Novel alginate sponges for cell culture and transplantation.
Biomaterials 18:583–590
46. Zmora S, Glicklis R, Cohen S (2002) Tailoring the pore architecture in 3-D alginate scaffolds
by controlling the freezing regime during fabrication. Biomaterials 23:4087–4094
47. Leor J et al (2000) Bioengineered cardiac grafts: a new approach to repair the infarcted myo-
cardium? Circulation 102(Suppl. II):56–61
48. Dar A, Shachar M, Leor J, Cohen S (2002) Optimization of cardiac cell seeding and distribu-
tion in 3D porous alginate scaffolds. Biotechnol Bioeng 80:305–312
49. Rosso F, Giordano A, Barbarisi M, Barbarisi A (2004) From cell–ECM interactions to tissue
engineering. J Cell Physiol 199:174–180
50. Shachar M, Tsur-Gang O, Dvir T, Leor J, Cohen S (2011) The effect of immobilized RGD
peptide in alginate scaffolds on cardiac tissue engineering. Acta Biomater 7:152–162
51. Sapir Y, Kryukov O, Cohen S (2011) Integration of multiple cell–matrix interactions into algi-
nate scaffolds for promoting cardiac tissue regeneration. Biomaterials 32:1838–1847
52. Cardin AD, Weintraub HJ (1989) Molecular modeling of protein–glycosaminoglycan interac-
tions. Arteriosclerosis 9:21–32
53. Dvir T et al (2011) Nanowired three-dimensional cardiac patches. Nat Nanotechnol 6:720–725
54. Sapir Y, Polyak B, Cohen S (2014) Cardiac tissue engineering in magnetically actuated scaf-
folds. Nanotechnology 25:014009
55. Lee K, Silva EA, Mooney DJ (2011) Growth factor delivery-based tissue engineering: general
approaches and a review of recent developments. J R Soc Interface 8:153–170
56. Maltais S, Tremblay JP (2010) The paracrine effect: pivotal mechanism in cell-based cardiac
repair. J Cardiovasc Transl Res 3:652–662
57. Gnecchi M, Zhang Z, Ni A, Dzau VJ (2008) Paracrine mechanisms in adult stem cell signaling
and therapy. Circ Res 103:1204–1219
58. Mirotsou M, Jayawardena TM, Schmeckpeper J, Gnecchi M, Dzau VJ (2011) Paracrine
mechanisms of stem cell reparative and regenerative actions in the heart. J Mol Cell Cardiol
50:280–289
59. Ratajczak MZ et al (2012) Pivotal role of paracrine effects in stem cell therapies in regenera-
tive medicine: can we translate stem cell-secreted paracrine factors and microvesicles into bet-
ter therapeutic strategies? Leukemia 26:1166–1173
60. Camussi G, Deregibus MC, Cantaluppi V (2013) Role of stem-cell-derived microvesicles in
the paracrine action of stem cells. Biochem Soc Trans 41:283–287
61. Vandervelde S, van Luyn MJ, Tio RA, Harmsen MC (2005) Signaling factors in stem cell-
mediated repair of infarcted myocardium. J Mol Cell Cardiol 39:363–376
62. Hao X et al (2007) Angiogenic effects of sequential release of VEGF-A(165) and PDGF-BB
with alginate hydrogels after myocardial infarction. Cardiovasc Res 75:178–185
63. Banquet S et al (2011) Arteriogenic therapy by intramyocardial sustained delivery of a novel
growth factor combination prevents chronic heart failure. Circulation 124:1059–1069
64. Ruvinov E, Leor J, Cohen S (2011) The promotion of myocardial repair by the sequential
delivery of IGF-1 and HGF from an injectable alginate biomaterial in a model of acute myo-
cardial infarction. Biomaterials 32:565–578
7 Alginate Application for Heart and Cardiovascular Diseases 211
65. Conti E et al (2004) Insulin like growth factor-1 as a vascular protective factor. Circulation
110:2260–2265
66. Nakamura T, Mizuno S, Matsumoto K, Sawa Y, Matsuda H, Nakamura T (2000) Myocardial
protection from ischemia/reperfusion injury by endogenous and exogenous HGF. J Clin Invest
106:1511–1519
67. Hausenloy DJ, Yellon DM (2009) Cardioprotective growth factors. Cardiovasc Res 83:179–194
68. Ruvinov E, Leor J, Cohen S (2010) The effects of controlled HGF delivery from an affin-
ity binding alginate biomaterial on angiogenesis and blood perfusion in a hindlimb ischemia
model. Biomaterials 31:4573–4582
69. Dvir T et al (2009) Prevascularization of cardiac patch on the omentum improves its therapeu-
tic outcome. Proc Natl Acad Sci U S A 106:14990–14995
70. Rodness J et al (2016) VEGF-loaded microsphere patch for local protein delivery to the
ischemic heart. Acta Biomaterialia. pii: S1742-7061(16)30472-X. doi: 10.1016/j.act-
bio.2016.09.009. [Epub ahead of print]
71. Henri O et al (2016) Selective stimulation of cardiac lymphangiogenesis reduces myocar-
dial edema and fibrosis leading to improved cardiac function following myocardial infarction.
Circulation 133:1484–1497
72. Kang HK et al (2016) Inducible HGF-secreting human umbilical cord blood-derived
MSCs Produced via TALEN-mediated genome editing promoted angiogenesis. Mol Ther
24:1644–1654
73. Bin D, Hockaday AL, Kang KH, Butcher JT (2013) 3d bioprinting of heterogeneous aortic
valve conduits with alginate/gelatin hydrogels. J Biomed Mater Res A 101:1255–1264
74. Hockaday LA et al (2012) Rapid 3D printing of anatomically accurate and mechanically het-
erogeneous aortic valve hydrogel scaffolds. Biofabrication 4(3):035005
75. Liu Y, Sakai S, Taya M (2016) Engineering tissues with a perfusable vessel-like network
using endothelialized alginate hydrogel fiber and spheroid-enclosing microcapsules. Heliyon
2:e00067
76. Kinoshita K et al (2016) Fabrication of multilayered vascular tissues using microfluidic aga-
rose hydrogel platforms. Biotechnology Journal. doi: 10.1002/biot.201600083. [Epub ahead
of print]
77. Jia W et al (2016) Direct 3D bioprinting of perfusable vascular constructs using a blend bioink.
Biomaterials 106:58–68
78. Kevadiya BD, Joshi GV, Bajaj HC (2010) Layered bionanocomposites as carrier for procain-
amide. Int J Pharm 388:280–286
79. Beckerman Z et al (2014) A novel amiodarone-eluting biological glue for reducing post-
operative atrial fibrillation: first animal trial. J Cardiovasc Pharmacol Ther 19:481–491
80. Segale L, Mannina P, Giovannelli L, Muschert S, Pattarino F (2016) Formulation and coating
of alginate and alginate-hydroxypropylcellulose pellets containing Ranolazine. J Pharm Sci
S0022-3549(16):41633–41633. https://doi.org/10.1016/j.xphs.2016.08.001. [Epub ahead of
print]
81. Lovich MA, Wei A, Maslov MY, Wu PI, Edelman ER (2011) Local epicardial inotropic drug
delivery allows targeted pharmacologic intervention with preservation of myocardial loading
conditions. J Pharm Sci 100(11):4993–5006
82. Maslov MY, Edelman ER, Wei AE, Pezone MJ, Lovich MA (2013) High concentrations of
drug in target tissues following local controlled release are utilized for both drug distribution
and biologic effect: an example with epicardial inotropic drug delivery. J Control Release
171(2):201–207
83. Maslov MY et al (2014) Myocardial drug distribution generated from local epicardial appli-
cation: potential impact of cardiac capillary perfusion in a swine model using epinephrine.
J Control Release 194:257–265
84. Liu TC, Ismail S, Brennan O, Hastings C, Duffy GP (2013) Encapsulation of cardiac stem cells
in superoxide dismutase-loaded alginate prevents doxorubicin mediated toxicity. J Tissue Eng
Regen Med 7:302–311
212 Z. Xu and M.T. Lam
85. Terakado S et al (2012) Sodium alginate oligosaccharides attenuate hypertension and asso-
ciated kidney damage in Dahl salt-sensitive rats fed a high-salt diet. Clin Exp Hypertens
34(2):99–106
86. Ueno M et al (2012) Sodium alginate oligosaccharides attenuate hypertension in spontane-
ously hypertensive rats fed a low-salt diet. Clin Exp Hypertens 34(5):305–310
87. Moriya C et al (2013) Subcutaneous administration of sodium alginate oligosaccharides pre-
vents salt-induced hypertension in Dahl salt-sensitive rats. Clin Exp Hypertens 35(8):607–613
88. Chen YY et al (2010) Preventive effects of low molecular mass potassium alginate extracted
from brown algae on DOCA salt-induced hypertension in rats. Biomed Pharmacother
64:291–295
89. Cohn JN, Ferrari R, Sharpe N (2000) Cardiac remodeling – concepts and clinical implications:
a consensus paper from an international forum on cardiac remodeling. J Am Coll Cardiol
35:569–582
90. Zhang P, Zhang H, Wang H, Wei Y, Hu S (2006) Artificial matrix helps neonatal cardiomyo-
cytes restore injured myocardium in rats. Artif Organs 30:86–93
Chapter 8
Alginates in Dressings and Wound
Management
Michael Clark
Abstract This chapter considers how wound dressings are used in the treatment of
wounds identifying the ideal properties of a wound dressing. Changes in the treat-
ment of wounds with dressings since 1980 are discussed highlighting the current
availability of a wide range of advanced wound dressings that clinicians have to
select from for each wound they treat. Alginate wound dressings are introduced with
their chemistry briefly considered and their indications and contraindications for
clinical use reported. The clinical evidence supporting the use of alginate wound
dressings is discussed highlighting the generally weak evidence underpinning the
use of all advanced wound dressings. Recent reviews of the effectiveness of alginate
dressings noted that across all the studies, there were no statistically significant dif-
ferences between the outcomes achieved using the alginate dressings and the com-
parison groups. It is concluded that alginate dressings are currently not in widespread
use in the UK National Health Service and may now be considered as comparisons
against which new technologies may be compared.
8.1 Introduction
Wounds to the skin and underlying tissues are both common and costly to health-
care services. All surgical patients will have a wound and while most heal unevent-
fully, local infections are common with surgical site infections (SSI) ranging from
0.3 per 1000 in-patient days to 8.2 per 1000 in-patient days dependent upon the type
of surgery [1]. Other wound aetiologies include burns, traumatic wounds, pressure
M. Clark (*)
Welsh Wound Innovation Centre, Ynysmaerdy, Pontyclun, Wales
Wound Healing Practice Development Unit, Birmingham City University, Birmingham, UK
e-mail: Michael.Clark@wwic.wales
ulcers (bedsores), leg ulcers and diabetic foot ulcers. The final three wound aetiolo-
gies are often known as chronic wounds with these wounds often failing to heal
after 3 months of treatment [2].
There are many estimates of the number of patients with chronic wounds, for
example, Posnett and Franks [3] reported that at any moment in time there would be
around 70–100,000 people with leg ulcers in the United Kingdom, 64,000 with
diabetic foot ulcers and 20,000 with pressure ulcers. The cost to health services due
to wound treatment is high; Phillips et al. [4] calculated that in Wales (population
3,168,000), each year wound care costs 6% of all expenditure on health or
£330,000,000. Wounds are not only common and costly, but also they pose serious
consequences for patients in terms of increased morbidity (increased pain, malo-
dour, reduced activity and mobility) and mortality. Between 1990 and 2001, 114,380
people in the United States had pressure ulcers noted as a cause of death, with
21,365 (18.7%) having pressure ulcers as the primary cause of mortality [5].
The detailed biology of wound healing is beyond the scope of this chapter with a
recent reference covering this complex process by Martin and Nunan [2]. The heal-
ing of wounds is often very simplified into four overlapping phases [6]:
Bleeding and haemostasis
Platelets aggregate and attach to collagen surfaces and release vasoactive agents
prompting blood clotting while chemokine release attracts inflammatory cells
(neutrophils).
Inflammation
Initial attraction of neutrophils to kill microbes that have accessed the wound and
later entry of macrophages to control repair processes and remove neutrophils
and cell debris.
Cell proliferation and matrix deposition
Formation of granulation tissue and new blood vessels and formation of a new
epithelial layer
Matrix remodelling
Remodelling of collagen, reduction in blood vessels and scar formation
In simplistic terms, wounds are closed by primary or secondary intention [6]. A
wound healing by primary intention, for example, a surgical wound, has the two
edges of the wound brought together and held in place with sutures or tape. Whereas
in wound healing by secondary intention, the edges of the wound cannot be approxi-
mated, and the wound will heal through a combination of wound contraction, gen-
8 Alginates in Dressings and Wound Management 215
eration of new soft tissue and blood vessels and re-epithelialisation. Chronic wounds
are examples of wounds that will heal by secondary intention (Fig. 8.1). A wound
healing by secondary intention would typically be covered with a wound dressing.
areas of damaged but intact skin (Category I pressure ulcer). Dry dressings, c ommon
in 1983, were only used on five wounds [13]. By 2009, clinicians appeared to be
using advanced wound dressings selectively upon different presentations of pres-
sure ulcer. Where the skin was broken, but damage restricted to the epidermis
(Category II pressure ulcers), foam and hydrocolloid dressings were most prevalent.
As the severity of the wound increased involving tissues below the dermis, then
antimicrobial dressings were commonly used. So between 1983 and 1999, wound
dressings were both more commonly used, and their complexity increased.
Today, clinicians are faced with a wide range of choices of products with 958
products listed on an on-line compendium of wound care products, most of these
being wound dressings [14]. The vast range of wound dressings available to clini-
cians poses questions regarding how clinicians are expected to discriminate between
the hundreds of wound dressings now available and to make an appropriate dressing
selection for each wound they encounter.
This list of the properties of the ideal dressing dates to times when few advanced
wound dressings were available and concluded that the future would see dressings
specified for individual wound aetiologies and for different stages of the wound
healing continuum [15]. Perhaps we now practice within that ‘future’ where wound
dressings are selected for different presentations of wound and with greater empha-
sis placed upon the ability of a dressing to manage specific phases of wound heal-
ing, for example, odour control, debridement and exudate management.
The preceding discussion has explored the development and increased use of wound
dressings in wound management over the past 30 years. One of the earliest classes
of advanced wound dressing that remains in widespread use today is alginate dress-
ings, of which 48 types are currently available to clinicians in the UK National
Health Service [14, 16]. Seaweed has long been recognised as an aid in wound heal-
ing dating back to the Romans and then in later use among sailors to stop bleeding
and by doctors draining abdominal wall abscesses; however it is challenging to
verify these anecdotes [17–19]. After the Second World War, alginate dressings
were in use across 70 UK hospitals as haemostatic agents initially in surgical
wounds with their use extending into accident and emergency departments [17].
However, the early use of alginate dressings came to a halt in the early 1970s when
general alginate production reduced due to the availability of cheaper alternative
products [17] with a resurgence in interest in alginate dressings in the early 1980s
as interest in wound healing and the role of advanced dressing materials expanded.
Alginate dressings are founded upon the mixing of sodium carbonate or sodium
hydroxide with the alginic acid extracted from seaweed [17, 19] to form sodium
alginate. The sodium alginate is then forced under pressure into a solution of cal-
cium salt leading to the formation of fibres of calcium alginate. To the calcium
alginate, sodium alginate may also be added to help accelerate the process of gelling
when in contact with the fluid leaking from a wound (wound exudate). Commercially
manufactured alginate dressings will vary both in the mix of calcium alginate and
sodium alginate fibres but also in the proportion of the chains produced using the
constituting monomers of the alginic acid (β-D-mannuronic acid and α-L-guluronic
acid). M-group chains contain all β-D-mannuronic acid, while G-group chains con-
tain all α-L-guluronic acid; the final chain (MG group) contains alternate units of
β-D-mannuronic acid and α-L-guluronic acid [19]. The relative proportions of the
M and G groups present will influence the interaction of the alginate dressing with
the wound. When an alginate dressing is in contact with a wound, there is an ion
218 M. Clark
exchange between the calcium ions in the dressing and the sodium ions in blood or
wound exudate; when this exchange reaches a point where sufficient calcium ions
have been exchanged, the alginate fibres begin to swell and partially dissolve form-
ing a gel. Gel formation is faster where M groups predominate with the resulting gel
softer and more elastic than the gel produced by a predominantly G-group alginate
dressing. While dressings with predominantly M groups gel faster, the greater dis-
solution of the fibres affects dressing removal which is typically undertaken through
irrigation of the wound to remove the partially dissolved fibres. Predominantly
G-group alginates swell less and so can be removed from the wound as an intact
dressing [19].
There is a general lack of robust clinical evidence for the effect of wound dressings
within wound healing [32]. The reasons for this are varied – among these are a lack of
research funding to allow large-scale clinical trials; challenging methodological issues
around masking patients, clinicians and researchers to the dressings allocated to study
participants; and the strong commercial drive towards funding evaluations leading to
multiple small studies which are often uncontrolled. The general weakness in the evi-
dence base for the effect of wound dressings is equally applicable to alginate dressings.
Thomas [17] provided an excellent overview of the clinical studies of alginate
use across a range of wound aetiologies including pressure ulcers, leg ulcers, burns
and donor sites, foot care, surgery and dental practice. Of the 117 references cited
all bar 13 dated to before 2000 and often report the use of products no longer com-
mercially available. The Cochrane Wounds Group (http://wounds.cochrane.org)
reports systematic reviews of the literature reporting controlled trials of wound
healing interventions in human subjects. Since 2013, the Wounds Group have pub-
lished three reviews of the role of alginate dressings in the treatment of diabetic foot
ulcers [33], leg ulcers [34] and pressure ulcers [35]. Across the three reviews, there
were 17 controlled studies with 1006 participants. Alginate dressings were com-
220 M. Clark
8.9 Conclusions
While alginate dressings have been used for around 30 years in wound healing, their
use is now relatively limited in clinical practice. Vowden and Vowden reported the use
of advanced wound dressings used to treat leg ulcers [37], pressure ulcers [13] and
acute wounds [38] across the population of Bradford, UK. Of the 1671 wounds
reported, only 57 were dressed with an alginate dressing. The apparent currently lim-
ited use of alginate dressings may offer opportunities for a revival of alginate dress-
ings potentially through the manipulation of the sodium and calcium alginate fibre
M- and G-group composition along with the introduction of silver and other antimi-
crobial agents. Such refinements in alginate manufacture may lead to increased fluid
handling properties combined with antimicrobial effects. Thomas [17] suggested four
areas for exploration which might help revive interest in alginate wound dressings:
• Can changes in the chemical composition of alginate dressings be correlated
with either wound healing or wound infection rates?
• Can the composition of alginate dressings be altered to stimulate cytokine
production?
• Can alginate dressings be manipulated to increase their absorption of bacteria
and proteolytic enzymes?
• Can alginate dressings be used to treat infected or malodorous wounds?
Clark [19] added a fifth question to this list:
• Can alginate dressings be used to manage wounds that have blood as part of the
wound exudate?
8 Alginates in Dressings and Wound Management 221
Without investment to answer these questions, it is unlikely that the clinical use
of alginate dressings will increase with other dressing technologies, also designed
to manage exudate. Perhaps the time is fast approaching when alginate wound
dressings will have no role within wound management having been effectively
replaced by other wound technologies.
References
21. Chrisman CA (2010) Care of chronic wounds in palliative care and end-of-life patients. Int
Wound J 7:214–235
22. Sweeney IR, Miraftab M, Collyer G (2012) A critical review of modern and emerg-
ing absorbent dressings used to treat exuding wounds. Int Wound J. https://doi.
org/10.1111/j.1742-481X.2011.00923.x
23. Thomas S (1992) Observations on the fluid handling properties of alginate dressings. Pharm
J 248:85–851
24. Clark R, Bradbury S (2010) SILVERCEL® non-adherent made easy. Wounds International
1(5.) Available from http://www.woundsinternational.com
25. Harris CL, Holloway S (2012) Development of an evidence-based protocol for care of pilo-
nidal sinus wounds healing by secondary intent using a modified Reactive Delphi procedure.
Part 2: methodology, analysis and results. Int Wound J 9:173–188
26. Higgins L, Wasiak J, Spinks A, Cleland H (2012) Split-thickness skin graft donor site manage-
ment: a randomized controlled trial comparing polyurethane with calcium alginate dressings.
Int Wound J 9:126–131
27. Ravnskog FA, Espehaug B, Indrekvam K (2011) Randomised clinical trial comparing
Hydrofiber and alginate dressings post-hip replacement. J Wound Care 20(3):136–142
28. Wiegand C, Heinze T, Hipler U (2009) Comparative in vitro study on cytotoxicity, antimicro-
bial activity, and binding capacity for pathophysiological factors in chronic wounds of alginate
and silver-containing alginate. Wound Repair Regen 17:511–521
29. Percival SL, Slone W, Linton S, Okel T, Corum L, Thomas JG (2011) The antimicrobial effi-
cacy of a silver alginate dressing against a broad spectrum of clinically relevant wound iso-
lates. Int Wound J 8:237–243
30. Hooper SJ, Percival SL, Hill KE, Thomas DW, Hayes AJ, Williams DW (2012) The visualisa-
tion and speed of kill of wound isolates on a silver alginate dressing. Int Wound J. https://doi.
org/10.1111/j.1742-481X.2012.00927.x
31. Best Practice Statement: The use of topical antiseptic/antimicrobial agents in wound manage-
ment (2011). 2nd edition. Wounds UK, London
32. Evidence summary ESMPB2 (2016). Chronic wounds: advanced wound dressings and antimi-
crobial dressings. National Institute for Health and care Excellence. Accessed at https://www.
nice.org.uk/advice/esmpb2/chapter/Key-points-from-the-evidence on 23 Mar 2017
33. Dumville JC, O'Meara S, Deshpande S, Speak K (2012) Alginate dressings for healing dia-
betic foot ulcers. Cochrane Database Syst Rev., Issue 2. Art. No.: CD009110. https://doi.
org/10.1002/14651858.CD009110.pub2
34. O'Meara S, Martyn-St James M, Adderley UJ (2015) Alginate dressings for venous leg ulcers.
Cochrane Database Syst Rev., Issue 8. Art. No.: CD010182. https://doi.org/10.1002/14651858.
CD010182.pub3
35. Dumville JC, Keogh SJ, Liu Z, Stubbs N, Walker RM, Fortnam M (2015) Alginate dress-
ings for treating pressure ulcers. Cochrane Database Syst Rev., Issue 5. Art. No.: CD011277.
https://doi.org/10.1002/14651858.CD011277.pub2
36. Monsen C, Wann-Hansson C, Wictorsson C, Acosta A (2014) Vacuum-assisted wound closure
versus alginate for the treatment of deep perivascular wound infections in the groin after vas-
cular surgery. J Vasc Surg 59(1):145–151
37. Vowden KR, Vowden P (2009) The prevalence, management and outcome for patients with
lower limb ulceration identified in a wound care survey within one English health care district.
J Tiss Viab 18(1):13–19
38. Vowden KR, Vowden P (2009) The prevalence, management and outcome for acute wounds
identified in a wound care survey within one English health care district. J Tiss Viab 18(1):7–12
Chapter 9
Alginates in Metabolic Syndrome
Abstract Alginates extracted from seaweeds are widely used for nutrition, but they
are underutilised for the prevention or reversal of human disease. Alginates are long
chains of α-L-guluronic acid and β-D-mannuronic acid from brown seaweeds that
act as readily available, low cost, non-toxic and biodegradable biopolymers. Sodium
alginates are primarily used for the management of gastrointestinal tract disorders,
but they are of potential use to attenuate the components of the metabolic syndrome
including obesity, type 2 diabetes, hypertension, non-alcoholic fatty liver disease
and dyslipidaemia. As prebiotics, alginates changed the gut microbiome to increase
production of short-chain fatty acids as substrates for Bifidobacteria. Alginates
inhibited pancreatic lipases and so decreased triacylglycerol breakdown and uptake.
Treatment with alginates decreased food intake by inducing satiety and increased
weight loss in patients on a calorie-restricted diet. Both glucose and fatty acid
uptake were reduced. In rat models of hypertension, alginates decreased blood pres-
sure. An alginate-antacid combination is an effective treatment of gastric reflux dis-
ease by forming a raft on the gastric contents. Alginates are important as drug
carriers in microparticles and nanoparticles to increase drug bioavailability, for
example, in drugs used for treatment of metabolic syndrome. Alginates are also
used to protect cells during transplantation from immune responses of the host,
allowing potential long-term control of some endocrine disorders such as type 1
diabetes and increased thermogenesis by brown adipocytes in obesity. There are
many potential uses for these versatile biopolymers in the treatment of human
disease.
S.A. Kumar
Advanced Centre for Treatment, Research and Education in Cancer (ACTREC),
Tata Memorial Centre, Kharghar, Navi Mumbai, Maharashtra, 410210, India
L. Brown (*)
School of Health and Wellbeing, University of Southern Queensland, Toowoomba, Australia
e-mail: Lindsay.Brown@usq.edu.au
9.1 Introduction
The human diet in Japan, Korea, China, Vietnam and the Philippines has included
seaweeds for hundreds of years. As potential functional foods, seaweeds may pre-
vent or treat disease in addition to their nutritional advantages [1, 2], but their use-
fulness is underestimated. Seaweeds are aquatic photosynthetic plants separated
into macroalgae and microalgae, with macroalgae classified into three types: brown
algae (Phaeophyta), red algae (Rhodophyta) and green algae (Chlorophyta) [3].
Brown seaweeds contain alginates as viscous water-soluble polysaccharides that
consist of (1,4)-linked chains of α-L-guluronic acid and β-D-mannuronic acid as the
major sugar residues [4, 5]. The concentration of alginates can be high in seaweeds,
for example, 15–30% in Ascophyllum nodosum (rockweed or Norwegian kelp),
20–45% in Laminaria digitata (oarweed) and 21–42% in Alaria esculenta (dabber-
locks or winged kelp) [6]. Brown seaweeds such as Sargassum sp. are widely used
in food and have been used in Traditional Chinese Medicine for nearly 2000 years
as potential anticancer, anti-inflammatory, antibacterial and anti-viral medicines
[7]. Laminaria japonica, a source of alginate and fucoidan, as well as fat-soluble
components, such as fucoxanthin and fucosterol, is widely used in Japan as a healthy
food (kombu) that may prevent obesity and diabetes [8]. The physiological responses
to consumption of seaweeds containing alginates mainly involve the gastrointestinal
tract, including increased gastric distension, delayed gastric emptying and enhance-
ment of satiety together with delayed postprandial glycaemia and insulin responses
[9]. Sodium alginates have found applications in the management of gastrointestinal
and metabolic complications, primarily of the components of the metabolic syn-
drome including obesity, type 2 diabetes, hypertension, non-alcoholic fatty liver
disease and dyslipidaemia [10–12]. This chapter will discuss the role of alginates to
improve health, primarily based on their changes in gastrointestinal function.
Insulin resistance or diabetes together with dyslipidaemia are used as clinical signs
to define metabolic syndrome. High-viscosity dietary fibres, including guar gum
and alginates, when incorporated in edible crispy bars containing 50 g carbohy-
drate, attenuated postprandial glycaemia without any change in gastrointestinal tol-
erance in healthy adults [28]. Supplementation with alginates and calcium in rats
attenuated postprandial glycaemic responses in streptozotocin (STZ)-induced dia-
betes in rats, probably due to increased viscosity as well as calcium-induced gel
formation [29]. This increased gelling is likely to delay gastric emptying and
decrease nutrient absorption in the small intestine, and both changes will decrease
postprandial glycaemic responses and attenuate peak insulinaemic responses [26].
In male patients, cholesterol uptake from a fixed diet increased with increasing
body fat; a single administration of 1.5 g sodium alginate with calcium carbonate
decreased uptake of glucose, cholesterol and triacylglycerols to the levels in healthy
subjects [30]. In rats fed a high cholesterol diet, addition of 2% calcium alginate to
the diet decreased plasma cholesterol concentrations, possibly due to an increased
bile acid excretion due to reduced intestinal reabsorption [31]. The gelling of both
high and low molecular weight alginates from Laminaria angustata in the stomach
was proposed as the mechanism for the reduced glucose uptake and insulin response
and increased cholesterol excretion from the gastrointestinal tract [32]. These stud-
ies suggest that the gastrointestinal changes induced by alginates can reduce dys-
lipidaemia in overweight/obese patients.
Hypertension is one of the diagnostic criteria for metabolic syndrome, and, further,
metabolic syndrome increases the risk of cardiovascular disease. Oral administra-
tion of low molecular weight potassium alginates (250 or 500 mg/kg body weight)
extracted from brown seaweeds normalised the cardiovascular changes in DOCA-
salt hypertensive rats to a greater extent than the same dose of potassium chloride
[33]. Sodium alginate oligosaccharides (60 mg/kg given subcutaneously) almost
completely abolished the increased blood pressure in Dahl salt-sensitive rats fed 4%
sodium chloride; this response may be due to improved kidney function with
decreased sclerosis and vascular injury in the kidney, together with direct effects on
vascular function, rather than by reducing salt absorption [34]. Dietary sodium algi-
nate oligosaccharides given as a 4% intervention in the diet induced small changes
in systolic blood pressure in male SHR, but renal glomerular damage was markedly
decreased [35]. In obese patients, sodium alginates had no effect on borderline
hypertensive patients with a baseline systolic blood pressure of 132.7 ± 2.2 mmHg
[27]. No studies were found that reported changes in blood pressures in hyperten-
sive patients following alginate interventions.
Alginates derived from Sargassum vulgare have shown antitumour activity in
mice. However, these mice developed acute tubular necrosis, suggesting intrinsic
9 Alginates in Metabolic Syndrome 227
Alginates have been given to relieve gastric reflux for many years. They precipitate upon
contact with gastric acid to produce low-viscosity gels of near-neutral pH, triggering the
sodium bicarbonate in the formulation to release carbon dioxide in the gel, which then
floats on the stomach contents as a raft close to the oesophageal-gastric junction [37].
Combination of calcium carbonate and sodium bicarbonate with sodium alginate reduced
gastric reflux episodes and increased time to reflux symptoms compared to patients
given antacid only [38]. This study showed that the alginate-antacid raft was localised to
below the diaphragm in these gastric reflux patients [38]. Despite a substantial placebo
response, an alginate-antacid combination reduced heartburn, regurgitation and dyspep-
sia in a randomised trial of 1107 patients with mild-to-moderate symptoms [38].
Obesity increases the risk of developing non-alcoholic fatty liver disease (NAFLD),
which may develop into non-alcoholic steatohepatitis (NASH) and then progress to
hepatocellular carcinoma [39]. In monosodium glutamate-treated mice with NASH
symptoms, oral sodium alginate treatment improved liver steatosis, insulin resis-
tance and chronic inflammation, and prevented the progression to carcinoma [11].
Translation to humans with NAFLD or NASH has not been reported.
Obesity is defined as a chronic inflammation [40], yet no studies have reported anti-
inflammatory effects of alginates in obese rats or humans. Adjuvant-induced
arthritic rats as a model of rheumatoid arthritis when treated with alginate from
Sargassum wightii showed decreased paw oedema, reduced activities of inflamma-
tory enzymes and reduced plasma inflammatory biomarkers [41]. However, anti-
inflammatory compounds such as indomethacin will induce gastric and small
intestinal ulcers. Sodium alginate has been proposed as a treatment to prevent
indomethacin-induced small intestinal injury as mice showed reduced intestinal
injury and reduced expression of mucin following alginate treatment [42] (Fig. 9.1).
228
Fig. 9.1 Possible therapeutic effects of alginates in the attenuation of metabolic complications. NAFLD non-alcoholic fatty liver disease; (↑) increased
response; (↓) diminished response
S.A. Kumar and L. Brown
9 Alginates in Metabolic Syndrome 229
9.8 A
lginates as Drug Carriers for Treatment of Metabolic
Syndrome
Alginates are readily available, low cost, non-toxic, biodegradable and versatile bio-
polymers and are therefore useful as drug carriers for therapy, for example, as
hydrocolloids in sustained-release products [43]. Further development to produce
nanoparticles with improved drug delivery is one of the success stories in pharma-
ceutical technology in the last 20 years, with a wide range of techniques being avail-
able for their preparation [44]. The effectiveness of nanoparticles depends on their
size and surface area, with a wide range of possible shapes giving a range of poten-
tial applications [45]. As an example, the oral bioavailability of insulin has been
improved by formulation as polymer-based nanoparticles, including alginate, but
these products have not reached the market [46]. The preparation of alginate mic-
roparticles and nanoparticles has been summarised, and future challenges have been
outlined [47].
There is now clear evidence that alginate-containing microparticles of oral hypo-
glycaemic drugs could be effective in type 2 diabetic patients. Metformin encapsu-
lated in alginate floating beads produced greater decreases in blood glucose
concentrations in Sprague-Dawley rats made diabetic following injection of strep-
tozotocin (60 mg/kg ip for 3 days) than with metformin alone [48]. Microcapsules
of gliclazide prepared using taurocholic acid and sodium alginate decreased hyper-
glycaemic responses in alloxan-induced type 1 diabetic rats [49]. Exenatide deliv-
ered orally in microcapsules with alginates and hyaluronate to db/db mice normalised
the blood glucose concentrations for 2 h; this response could be prolonged until 4 h
with increased exenatide doses for effective control of type 2 diabetes [50]. These
studies show the potential of micro- and nanoparticles to increase treatment options
for type 2 diabetes. These techniques may also apply to insulin treatment of type 1
diabetes, now exclusively given subcutaneously. Oral administration of chitosan-
alginate insulin nanoparticles reduced blood glucose concentrations in alloxan-
diabetic mice more slowly than subcutaneous insulin, with bioavailability of
approximately 8% [51]. Liver damage is common in diabetes. One possible alterna-
tive for treatment of liver tumours is the use of alginate microspheres with the anti-
neoplastic drug, amonafide, that causes serious adverse effects with oral delivery, to
achieve targeted delivery with reduced systemic toxicity [52]. Another option for
intracellular targeting of liver tumour cells is the use of microspheres with meso-
porous silica nanoparticles together with alginate providing high biocompatibility
and sustained release [53].
Alginate-containing nanoparticles may also be useful to administer lipid-
lowering drugs such as probucol [54]. The physical characteristics of these probucol
nanoparticles were appropriate for treatment [54]; similar nanoparticles of probucol
improved insulin release and decreased TNF-alpha production by pancreatic beta
cells cultured in 25.5 mM glucose [55].
Hypertension is an important component of the metabolic syndrome. Many anti-
hypertensive drugs have been formulated in alginate-containing nanoparticles for
230 S.A. Kumar and L. Brown
Transplantation has a long history in endocrinology with recent studies using iso-
lated β-islet cells or stem cells showing the potential of this procedure to restore the
endocrine activity of the pancreas [64]. Procedures to improve success include treat-
ment of the cells with alginates. In immune-competent STZ-induced type 1 diabetic
C57BL/6J mice, transplantation of in vitro-derived glucose-responsive mature
β-cells from human embryonic stem cells encapsulated using chemically modified
alginates via the intraperitoneal route normalised blood glucose concentrations up
to 174 days after transplantation with minimal graft rejection [65]. The develop-
ment of an oxygenated chamber system with immune-isolating alginate and poly-
membrane covers allowed the survival and function of human pancreatic islets
without immunosuppression [66]. Transplantation of these cells into a 63-year-old
man with a history of type 1 diabetes for 54 years was followed by persistent graft
function and regulated insulin secretion for at least 10 months, without immunosup-
pression [66].
As myocytes cannot replicate, cell transplantation is an attractive alternative to
improve cardiac function after injury. Foetal cardiomyocytes grown on porous algi-
nate scaffolds were transplanted into rats 7 days after myocardial infarction [67].
After 9 weeks, the transplanted cells had stimulated intense neovascularisation and
attenuated left ventricular dilatation and cardiac failure [67].
Unlike the heart, the liver can regenerate, but hepatocyte transplantation may be
needed in acute liver failure to provide short-term support. Further, these patients
may require liver transplantation, a major challenge for the health system [68].
Transplantation of rat hepatocytes microencapsulated with alginate markedly
improved liver parameters in a rat model of D-galactosamine-induced acute liver
failure; further, recovery of microbeads on day 8 after transplantation showed no
signs of adhesion or inflammation [69]. Alginate-polyethylene glycol microspheres
9 Alginates in Metabolic Syndrome 231
of human mesenchymal stem cells transplanted into mice delayed the development
of fibrosis in bile duct-ligated or carbon tetrachloride-treated mice [70]. After partial
hepatectomy in mice, the use of implanted alginate scaffolds supported the growth
of the remaining kidney, decreasing liver injury and improving survival [71].
Alginate microspheres with adipose tissue-derived stem cells could be transplanted
into recipient mice where the stem cells underwent hepatogenic differentiation to
cells that secreted albumin in the liver [72].
In contrast to white adipocytes, brown adipocytes may help control obesity [73].
The encapsulation of mouse embryonic stem cells in alginate hydrogel microstrands
allowed differentiation into brown adipocytes confirmed by the expression of
uncoupling protein 1 which is characteristic of these cells, as well as increased
expression with β3-adrenoceptor agonists [74]. Cell entrapment within alginate
microcapsules allows the cells to avoid the immune responses of the host; the use of
this technique with catabolic cells that use lipids for thermogenesis may be appli-
cable for the treatment of obesity [75]. Alginate-poly-L-lysine microencapsulated
Chinese hamster ovary (CHO)-E3 cells secreted apolipoprotein E3 when given
intraperitoneally to mice, leading to decreased cholesterol and increased HDL con-
centrations in the plasma [76]. This technology may be feasible to minimise athero-
sclerosis in obese and diabetic patients.
In conclusion, alginates are low cost, mostly non-toxic and versatile biopolymers
that can be used for treatment of many gastrointestinal problems. In addition, they
are useful in microparticles and nanoparticles as drug carriers and to protect cells
during transplantation. However, the full potential of these natural products as func-
tional foods needs to be further researched.
References
1. Kumar SA, Brown L (2013) Seaweeds as potential therapeutic interventions for the metabolic
syndrome. Rev Endocr Metab Disord 14(3):299–308
2. Brown ES, Allsopp PJ, Magee PJ, Gill CI, Nitecki S, Strain CR, McSorley EM (2014) Seaweed
and human health. Nutr Rev 72(3):205–216
3. MacArtain P, Gill CI, Brooks M, Campbell R, Rowland IR (2007) Nutritional value of edible
seaweeds. Nutr Rev 65(12 Pt 1):535–543
4. Fischer FG, Dorfel H (1955) Die Polyuronsäuren der Braunalgen (Kohlenhydrate der Algen-1).
Hoppe Seylers Z Physiol Chem 302(4-6):186–203
5. Lee KY, Mooney DJ (2012) Alginate: properties and biomedical applications. Prog Polym Sci
37(1):106–126
6. Chemical composition of seaweeds; The Seaweed Site: information on marine algae (2016)
Algaebase. http://www.seaweed.ie/nutrition/index.php
7. Liu L, Heinrich M, Myers S, Dworjanyn SA (2012) Towards a better understanding of medici-
nal uses of the brown seaweed Sargassum in Traditional Chinese Medicine: a phytochemical
and pharmacological review. J Ethnopharmacol 142(3):591–619
8. Shirosaki M, Koyama T (2011) Laminaria japonica as a food for the prevention of obesity and
diabetes. Adv Food Nutr Res 64:199–212
9. El Khoury D, Goff HD, Anderson GH (2015) The role of alginates in regulation of food intake
and glycemia: a gastroenterological perspective. Crit Rev Food Sci Nutr 55(10):1406–1424
232 S.A. Kumar and L. Brown
10. Chater PI, Wilcox MD, Houghton D, Pearson JP (2015) The role of seaweed bioactives in the
control of digestion: implications for obesity treatments. Food Funct 6(11):3420–3427
11. Miyazaki T, Shirakami Y, Kubota M, Ideta T, Kochi T, Sakai H, Tanaka T, Moriwaki H,
Shimizu M (2016) Sodium alginate prevents progression of non-alcoholic steatohepatitis and
liver carcinogenesis in obese and diabetic mice. Oncotarget 7(9):10448–10458
12. Terakado S, Ueno M, Tamura Y, Toda N, Yoshinaga M, Otsuka K, Numabe A, Kawabata Y,
Murota I, Sato N, Uehara Y (2012) Sodium alginate oligosaccharides attenuate hyperten-
sion and associated kidney damage in Dahl salt-sensitive rats fed a high-salt diet. Clin Exp
Hypertens 34(2):99–106
13. de Jesus Raposo MF, de Morais AM, de Morais RM (2016) Emergent sources of prebiotics:
seaweeds and microalgae. Mar Drugs 14(2):27
14. Terada A, Harat T, Mitsuoka T (1995) Effect of dietary alginate on the faecal microbiota and
faecal metabolic activity in humans. Microb Ecol Health Dis 8:259–266
15. Sanmiguel C, Gupta A, Mayer EA (2015) Gut microbiome and obesity: a plausible explana-
tion for obesity. Curr Obes Rep 4(2):250–261
16. John GK, Mullin GE (2016) The gut microbiome and obesity. Curr Oncol Rep 18(7):45
17. Wilcox MD, Brownlee IA, Richardson JC, Dettmar PW, Pearson JP (2014) The modulation of
pancreatic lipase activity by alginates. Food Chem 146:479–484
18. Houghton D, Wilcox MD, Chater PI, Brownlee IA, Seal CJ, Pearson JP (2015) Biological
activity of alginate and its effect on pancreatic lipase inhibition as a potential treatment for
obesity. Food Hydrocoll 49:18–24
19. Hellstrom PM (2013) Satiety signals and obesity. Curr Opin Gastroenterol 29(2):222–227
20. Jensen MG, Kristensen M, Belza A, Knudsen JC, Astrup A (2012) Acute effect of alginate-
based preload on satiety feelings, energy intake, and gastric emptying rate in healthy subjects.
Obesity (Silver Spring) 20(9):1851–1858
21. Hoad CL, Rayment P, Spiller RC, Marciani L, Alonso B de C, Traynor C, Mela DJ, Peters HP,
Gowland PA (2004) In vivo imaging of intragastric gelation and its effect on satiety in humans.
J Nutr 134(9):2293–2300
22. Pelkman CL, Navia JL, Miller AE, Pohle RJ (2007) Novel calcium-gelled, alginate-pectin
beverage reduced energy intake in nondieting overweight and obese women: interactions with
dietary restraint status. Am J Clin Nutr 86(6):1595–1602
23. Paxman JR, Richardson JC, Dettmar PW, Corfe BM (2008) Daily ingestion of alginate reduces
energy intake in free-living subjects. Appetite 51(3):713–719
24. Odunsi ST, Vazquez-Roque MI, Camilleri M, Papathanasopoulos A, Clark MM, Wodrich L,
Lempke M, McKinzie S, Ryks M, Burton D, Zinsmeister AR (2010) Effect of alginate on
satiation, appetite, gastric function, and selected gut satiety hormones in overweight and obe-
sity. Obesity (Silver Spring) 18(8):1579–1584
25. Carnahan S, Balzer A, Panchal SK, Brown L (2014) Prebiotics in obesity. Panminerva Med
56(2):165–175
26. Jensen MG, Pedersen C, Kristensen M, Frost G, Astrup A (2013) Review: efficacy of alginate
supplementation in relation to appetite regulation and metabolic risk factors: evidence from
animal and human studies. Obes Rev 14(2):129–144
27. Jensen MG, Kristensen M, Astrup A (2012) Effect of alginate supplementation on weight loss
in obese subjects completing a 12-wk energy-restricted diet: a randomized controlled trial. Am
J Clin Nutr 96(1):5–13
28. Williams JA, Lai CS, Corwin H, Ma Y, Maki KC, Garleb KA, Wolf BW (2004) Inclusion of
guar gum and alginate into a crispy bar improves postprandial glycemia in humans. J Nutr
134(4):886–889
29. Ohta A, Taguchi A, Takizawa T, Adachi T, Kimura S, Hashizume N (1997) The alginate reduce
the postprandial glycaemic response by forming a gel with dietary calcium in the stomach of
the rat. Int J Vitam Nutr Res 67(1):55–61
9 Alginates in Metabolic Syndrome 233
30. Paxman JR, Richardson JC, Dettmar PW, Corfe BM (2008) Alginate reduces the increased
uptake of cholesterol and glucose in overweight male subjects: a pilot study. Nutr Res
28(8):501–505
31. Idota Y, Kogure Y, Kato T, Ogawa M, Kobayashi S, Kakinuma C, Yano K, Arakawa H,
Miyajima C, Kasahara F, Ogihara T (2016) Cholesterol-lowering effect of calcium alginate in
rats. Biol Pharm Bull 39(1):62–67
32. Kimura Y, Watanabe K, Okuda H (1996) Effects of soluble sodium alginate on cholesterol
excretion and glucose tolerance in rats. J Ethnopharmacol 54(1):47–54
33. Chen YY, Ji W, Du JR, Yu DK, He Y, Yu CX, Li DS, Zhao CY, Qiao KY (2010) Preventive
effects of low molecular mass potassium alginate extracted from brown algae on DOCA salt-
induced hypertension in rats. Biomed Pharmacother 64(4):291–295
34. Moriya C, Shida Y, Yamane Y, Miyamoto Y, Kimura M, Huse N, Ebisawa K, Kameda Y, Nishi
A, Du D, Yoshinaga M, Murota I, Sato N, Uehara Y (2013) Subcutaneous administration of
sodium alginate oligosaccharides prevents salt-induced hypertension in Dahl salt-sensitive
rats. Clin Exp Hypertens 35(8):607–613
35. Ueno M, Tamura Y, Toda N, Yoshinaga M, Terakado S, Otsuka K, Numabe A, Kawabata Y,
Murota I, Sato N, Uehara Y (2012) Sodium alginate oligosaccharides attenuate hypertension
in spontaneously hypertensive rats fed a low-salt diet. Clin Exp Hypertens 34(5):305–310
36. de Paula Alves Sousa A, Barbosa PS, Torres MR, Martins AM, Martins RD, de Sousa Alves R,
de Sousa DF, Alves CD, Costa-Lotufo LV, Monteiro HS (2008) The renal effects of alginates
isolated from brown seaweed Sargassum vulgare. J Appl Toxicol 28(3):364–369
37. Rohof WO, Bennink RJ, Smout AJ, Thomas E, Boeckxstaens GE (2013) An alginate-antacid
formulation localizes to the acid pocket to reduce acid reflux in patients with gastroesophageal
reflux disease. Clin Gastroenterol Hepatol 11(12):1585–1591
38. Sun J, Yang C, Zhao H, Zheng P, Wilkinson J, Ng B, Yuan Y (2015) Randomised clinical trial:
the clinical efficacy and safety of an alginate-antacid (Gaviscon Double Action) versus pla-
cebo, for decreasing upper gastrointestinal symptoms in symptomatic gastroesophageal reflux
disease (GERD) in China. Aliment Pharmacol Ther 42(7):845–854
39. Noureddin M, Rinella ME (2015) Nonalcoholic fatty liver disease, diabetes, obesity, and hepa-
tocellular carcinoma. Clin Liver Dis 19(2):361–379
40. Gregor MF, Hotamisligil GS (2011) Inflammatory mechanisms in obesity. Annu Rev Immunol
29:415–445
41. Sarithakumari CH, Renju GL, Kurup GM (2013) Anti-inflammatory and antioxidant potential
of alginic acid isolated from the marine algae, Sargassum wightii on adjuvant-induced arthritic
rats. Inflammopharmacology 21(3):261–268
42. Horibe S, Tanahashi T, Kawauchi S, Mizuno S, Rikitake Y (2016) Preventative effects of
sodium alginate on indomethacin-induced small-intestinal injury in mice. Int J Med Sci
13(9):653–663
43. Tønnesen HH, Karlsen J (2002) Alginate in drug delivery systems. Drug Dev Ind Pharm
28(6):621–630
44. Rao JP, Geckeler KE (2011) Polymer nanoparticles: preparation techniques and size-control
parameters. Prog Polym Sci 36(7):887–913
45. Duchêne D, Gref R (2016) Small is beautiful: surprising nanoparticles. Int J Pharm
502(1–2):219–231
46. Fonte P, Araujo F, Silva C, Pereira C, Reis S, Santos HA, Sarmento B (2015) Polymer-
based nanoparticles for oral insulin delivery: revisited approaches. Biotechnol Adv 33(6 Pt
3):1342–1354
47. Paques JP, van der Linden E, van Rijn CJ, Sagis LM (2014) Preparation methods of alginate
nanoparticles. Adv Colloid Interf Sci 209:163–171
48. Nayak A, Jain SK, Pandey RS (2011) Controlling release of metformin HCl through incorpo-
ration into stomach specific floating alginate beads. Mol Pharm 8(6):2273–2281
234 S.A. Kumar and L. Brown
67. Leor J, Aboulafia-Etzion S, Dar A, Shapiro L, Barbash IM, Battler A, Granot Y, Cohen S
(2000) Bioengineered cardiac grafts: a new approach to repair the infarcted myocardium?
Circulation 102(19 Suppl 3):III56–III61
68. Koehler E, Watt K, Charlton M (2009) Fatty liver and liver transplantation. Clin Liver Dis
13(4):621–630
69. Jitraruch S, Dhawan A, Hughes RD, Filippi C, Soong D, Philippeos C, Lehec SC, Heaton ND,
Longhi MS, Mitry RR (2014) Alginate microencapsulated hepatocytes optimised for trans-
plantation in acute liver failure. PLoS One 9(12):e113609
70. Meier RP, Mahou R, Morel P, Meyer J, Montanari E, Muller YD, Christofilopoulos P, Wandrey
C, Gonelle-Gispert C, Buhler LH (2015) Microencapsulated human mesenchymal stem cells
decrease liver fibrosis in mice. J Hepatol 62(3):634–641
71. Shteyer E, Ben Ya’acov A, Zolotaryova L, Sinai A, Lichtenstein Y, Pappo O, Kryukov
O, Elkayam T, Cohen S, Ilan Y (2014) Reduced liver cell death using an alginate scaffold
bandage: a novel approach for liver reconstruction after extended partial hepatectomy. Acta
Biomater 10(7):3209–3216
72. Chen MJ, Lu Y, Simpson NE, Beveridge MJ, Elshikha AS, Akbar MA, Tsai HY, Hinske S,
Qin J, Grunwitz CR, Chen T, Brantly ML, Song S (2015) In situ transplantation of alginate
bioencapsulated adipose tissues derived stem cells (ADSCs) via hepatic injection in a mouse
model. PLoS One 10(9):e0138184
73. Schulz TJ, Tseng YH (2013) Brown adipose tissue: development, metabolism and beyond.
Biochem J 453(2):167–178
74. Unser AM, Mooney B, Corr DT, Tseng YH, Xie Y (2016) 3D brown adipogenesis to create
“Brown-Fat-in-Microstrands”. Biomaterials 75:123–134
75. Xu L, Shen Q, Mao Z, Lee LJ, Ziouzenkova O (2015) Encapsulation thermogenic preadipo-
cytes for transplantation into adipose tissue depots. J Vis Exp 100:e52806
76. Tagalakis AD, Diakonov IA, Graham IR, Heald KA, Harris JD, Mulcahy JV, Dickson G, Owen
JS (2005) Apolipoprotein E delivery by peritoneal implantation of encapsulated recombinant
cells improves the hyperlipidaemic profile in apoE-deficient mice. Biochim Biophys Acta
1686(3):190–199
Chapter 10
Alginate Oligomers and Their Use as Active
Pharmaceutical Drugs
Abstract Alginate oligomers retain most of the chemical and physical properties
of the higher molecular weight commercial alginates, retaining affinity towards
monovalent and divalent ions, which is dependent on the chemical composition of
the oligomer. However, due to their low molecular weight, they will normally not
form gels in the presence of divalent cations. This property is exploited in biological
systems to chelate multivalent ions and disrupt Ca2+-mediated cross-linking. Studies
have also identified interactions between alginate oligomers and complex mucin
polymer systems, bacteria and extracellular polymeric substance (EPS), which sug-
gests that these interactions are not simply the result of cationic chelation. By virtue
of their low molecular weight, alginate oligomers stay in solution at high concentra-
tion without significant increase in viscosity and can be tailor-made to precisely
defined chemical composition and molecular weight. This affords the opportunity to
design effective formulations with precisely defined properties and biological
effects. The properties now being identified for alginate oligomers represent a
promising new approach in the management of chronic lung diseases, biofilm infec-
tions and antibiotic use. This chapter outlines the research performed to date, high-
lighting the excellent safety profile and novel chemical characteristics of alginate
oligomers that emphasize their potential in multiple therapeutic applications.
10.1 A
lginate Oligomers: Composition, Structure
and Production
10.1.2 Production
Alginate oligomers, like their high molecular weight siblings, are commercially
manufactured by extraction of harvested brown algae such as Macrocystis,
Laminaria and Ascophyllum spp. The G-content of alginates derived from these
seaweeds varies greatly, not only between species but also within the different parts
of the plant. The G-content of alginate from the stipe and leaves of Laminaria
hyperborea can contain 75% and 35% G residues, respectively. The specific
G-content of alginates can be determined by nuclear magnetic resonance spectros-
copy (NMR). Alginates are also produced by several species of the bacteria
Pseudomonas and Azotobacter. The role of these bacterial producers has attracted a
great deal of attention for several reasons. Initially the focus was motivated by iden-
tifying the role of alginate in the pathogenesis of Pseudomonas aeruginosa lung
infections and developing an understanding of alginate biosynthesis and regulation
[1–6]. During recent years there has been an increasing focus on microbial strain
engineering to potentially utilize these bacterially produced or modified alginates
for high-value applications [2, 4, 6–9].
In both algae and bacteria, the polymer is first synthesized as poly-mannuronate
(polyM) chain, and then certain M residues are converted to G by mannuronan
C5-epimerases [10]. Alginate from Pseudomonas spp. does not contain G-blocks
10 Alginate Oligomers and Their Use as Active Pharmaceutical Drugs 239
due to the expression of a single alginate epimerase, the intracellular AlgG, which
introduces only single G residues into the alginate chain [1]. Conversely, Azotobacter
species such as Azotobacter vinelandii produce and secrete several epimerases
(AlgE1–AlgE7) that lead to alginates with a variety of physical properties and dif-
ferent structures including G-blocks [11, 12]. Bacterial mannuronan C5-epimerases
have been isolated and extensively characterized, providing powerful in vitro tools
to manipulate the structure of alginates, enabling the synthesis of new alginates with
defined chemical composition [10]. These enzymes have also been further engi-
neered for optimizing their properties in relation to the tailoring of alginates [13].
Unlike the alginates of algal origin, the bacterial alginates contain various degrees
of O-acetyl groups associated with the M residues. Acetylated M residues cannot be
epimerised, and thus acetylation is thought to be involved in controlling the level of
G residues in the alginate.
The G-content and composition of algal alginates differ greatly depending on the
species, the part of the algae used, the harvesting season and growth conditions,
implying great diversity in the algal epimerases. Indeed, genomic analyses of the
model brown algae Ectocarpus siliculosus and Saccharina japonica indicate the
presence of 45 and 105 mannuronan C5-epimerases, respectively [14, 15]. Algal
epimerases are apparently difficult to express in recombinant systems, and so far
only the successful expression of one of the S. japonica enzymes has been reported
[16]. This enzyme was produced in an insect-cell expression system and found to
introduce alternating MG structures when polyM polymer was epimerised in vitro,
thus resembling the A. vinelandii AlgE4.
To understand the chemical-physical properties, and thereby facilitate optimal
tailoring of alginate structures, it is important to characterize the alginates at a com-
positional level. By combining the action of sequence-specific alginate lyases with
analytical techniques like NMR, high-performance anion-exchange chromatogra-
phy with pulsed amperometric detection (HPAEC-PAD) and size exclusion chroma-
tography with multi-angle laser light scattering (SEC-MALLS), it is possible to
obtain information on the length and distribution of the different alginate block
structures (M, MG and G-blocks) (Fig. 10.1). This approach has been used success-
fully to determine the block structure and more specifically the length of G-blocks
in both natural and in vitro epimerised alginates [17, 18].
Alginate oligomers are prepared by controlled acid hydrolysis of high molecular
weight alginates [19, 20], followed by purification and characterization.
Characterization includes a determination of the G-content by NMR, protein and
endotoxin content and inorganic elements by inductively coupled plasma mass
spectroscopy (ICP-MS) and a determination of the molecular weight distribution by
HPAEC-PAD and SEC-MALLS. HPAEC-PAD is a powerful tool for determining
the distribution of oligosaccharides after acid hydrolysis of homopolysaccharides
(Fig. 10.2) [21].
240
Fig. 10.1 Diagram outlining the combined role of sequence-specific alginate lyases with analytical techniques to derive alginate oligomers with defined length
and distribution of block structures (M, MG and G-blocks). The lyase enzymes cleave the alginate polymer through β-elimination, resulting in unchanged satu-
rated uronate at the reducing end and an unsaturated (4-deoxy-L-erythro-hex-4-enepyranosyluronate) residue (Δ) at the non-reducing end [17]
P.D. Rye et al.
10 Alginate Oligomers and Their Use as Active Pharmaceutical Drugs 241
Alginate oligomers retain most of the basic chemical and physical properties of
commercial alginates of higher molecular weight, with some exceptions. Alginate
oligomers retain affinity towards monovalent and divalent ions, such affinity being
dependent on the chemical composition of the oligomer. Notably, oligoguluronates
can be considered as free G-blocks that readily compete for and bind to multivalent
cations. However, due to their low molecular weight, they will normally not gel or
introduce network connectivities [22]. This property can be exploited in biological
systems by using oligoguluronates to chelate multivalent ions like Ca2+, thereby
disrupting Ca2+-mediated cross-linking in the environment. Studies have also identi-
fied potential electrostatic interactions between alginate oligomers and complex
mucin polymer systems [23, 24], bacteria [25, 26] and extracellular polymeric sub-
stances (EPS) suggesting that these interactions are not simply the result of cationic
chelation. By virtue of their low molecular weight, alginate oligomers can stay in
solution at high concentration without significant increase in viscosity. While algi-
nates of higher molecular weights are usually polydisperse and contain a mixture of
molecules with different chemical composition [27], alginate oligomers can be
tailor-made to precisely defined chemical composition and molecular weight. This
opens up the possibility of designing highly efficient formulations with precisely
defined properties.
242 P.D. Rye et al.
10.2.2 Biocompatibility
The biocompatibility and low toxicity of alginates have been extensively reviewed.
Due to their advantageous properties and their safety profile, alginates are widely
employed in the food and pharmaceutical industry. Alginate oligomers have demon-
strated similar safety profiles with no safety issues reported. Alginate oligomers
have been proved to be safe for inhalation, as supported by clinical safety and toler-
ability studies (www.clinicaltrials.gov, Identifier: NCT00970346; NCT01465529).
There are currently no known mammalian enzymes that degrade alginates, and pre-
clinical toxicity studies have shown alginate oligomers are rapidly cleared through
the kidneys. The US FDA recognizes alginic acid and alginates as GRAS (generally
recognized as safe) for use in foodstuff (reference no. 21CFR184.1724).
A number of biological effects have been demonstrated with alginate oligomers that
highlight their potential use in a range of clinical applications. These essentially
address the challenges of mucus-biopolymer interactions such as those seen in cys-
tic fibrosis (CF) and other related respiratory diseases and the treatment of multidrug-
resistant (MDR) biofilm-related infections. The primary mechanisms of action
driving these effects and their potential clinical applications are outlined in the fol-
lowing section.
Mucins are a large family of glycoproteins which line the aerodigestive, urinary and
reproductive tracks (along with DNA, lipids, proteins and cellular debris) and
represent the major macromolecular component of mucus. Mucins are secreted by
specialized goblet cells and rapidly assemble and “unfold” at the cell surface (in a
partly Ca2+-dependent process) into a hydrated, cross-linked network. Mucin pro-
duction represents an effective barrier to pathogens and facilitates their removal/
clearance in the respiratory tract by the action of cilia. Mucus secretion serves to
protect the surface of the lung, intestine and urogenitary tract [28]. The overproduc-
tion, increased viscosity of mucus and/or failure to secrete mucus impairs this
mucosal barrier function and in a range of diseases favours microbial colonization,
chronic inflammation and luminal obstruction.
In cystic fibrosis (CF), a genetic disorder, the absence of a functional CF trans-
membrane conductance regulator (CFTR) protein leads to altered anion flow and a
dehydrated luminal surface. The loss of bicarbonate secretion through CFTR further
impedes normal mucus formation, which requires bicarbonate to sequester calcium
10 Alginate Oligomers and Their Use as Active Pharmaceutical Drugs 243
from packed condensed mucins enabling expansion and release from glands [29–
31]. In CF, this altered mucus results in bacterial overgrowth and an inflammatory
response with subsequent overproduction of additional viscous mucus [28] and a
concomitant decrease in lung function. Dependency on chronic antibiotic therapy in
CF eventually leads to the acquisition of MDR pathogens within the CF lung, with
associated morbidity and mortality. The accumulation of stagnant mucus in CF
patients also leads to complications observed in the CF intestine such as distal intes-
tinal obstruction syndrome (DIOS) and small intestinal bacterial overgrowth (SIBO)
[32]. The latter condition is not limited to CF but can be associated in a wide variety
of conditions where an abnormal intestinal transit may result in dysbiosis such as
inflammatory bowel disease (IBD). Osmotic laxatives and stool softeners are widely
prescribed to treat intestinal obstructions such as those found in DIOS and constipa-
tion. However, these treatments often requiring frequent repetition are not without
side-effects, including diarrhoea, vomiting and dehydration [33].
Initial ex vivo studies using CF sputum samples demonstrated the impressive
ability of alginate oligomers to modify the shear rheology of the viscous sputum
from patients with CF [26]. Subsequent studies using molecular dynamic modelling
and infrared spectroscopy have shown that this is the result of specific interactions
between the alginate oligomers and the mucin. Moreover, further studies of sputum
from CF patients receiving rhDNase-I (a mucolytic therapy that specifically breaks
up extracellular DNA in sputum) demonstrated that these alginate oligomers were
also able to potentiate the activity of this mucolytic DNase. It is thought that these
potentiating effects result from the direct interaction of alginate oligomers with the
mucin biopolymer network in CF mucus. The respiratory effects of the alginate
oligomer OligoG CF-5/20, a dry powder for inhalation developed by AlgiPharma
AS, are currently being investigated in phase 2b clinical trials.
A recent study performed at Case Western Reserve University [34] used a CF
mouse model to investigate the effects of an alginate oligomer in disrupting the
intestinal mucus accumulation in vivo. While no significant changes were observed
in the treated and non-treated wild-type (WT) mice, the administration of the algi-
nate oligomer to CF mice had a significant impact on the CF intestinal phenotype.
Intestinal transit times were normalized in the treated CF mice suggesting that intes-
tinal contents could more easily move through the small intestine and consequently
shorten intestinal transit time. Both short- (7 days) and long-term (25 days) treat-
ment of CF mice resulted in a significant decrease in intestinal obstruction and
dramatically improved survival to near WT levels (Fig. 10.3). The improved sur-
vival was most likely due to the reduction in accumulated mucus, which was the
direct result of the alginate oligomer enabling the normal unpacking of the mucus.
Although these effects have not yet been tested in other non-CF models of intestinal
disease, the dramatic improvement in intestinal transit time shows promise for other
intestinal complications. Indeed, evidence for the potential in other non-CF intesti-
nal conditions is provided by the ability to (1) chelate or sequester calcium and
(2) bind to mucin. Calcium is a major factor in pre-secreted mucin, and removal of
calcium leads to mucin release, unfolding and expansion [35]. Several studies have
shown that mucus aggregates dissolve in calcium chelators [30, 35]. In a study
244 P.D. Rye et al.
Fig. 10.3 Survival curve showing the effect of alginate oligomer treatment on the survival of CF
mice [34]. Alginate oligomer administered in drinking water improved the survival of CF mice
compared to untreated controls (Reprinted from Ref. [34], Copyright 2016 with permission from
Elsevier)
specific to CF, the combination of bicarbonate with the calcium chelator EDTA
increased the detachment of preformed CF mucus in ex vivo explants [36]. Although
alginates in general do have calcium-chelating properties, guluronate oligomers
specifically have an increased ability to bind calcium in vitro compared to M oligo-
mers [37]. Several studies have demonstrated that low molecular weight guluronate
oligomers are particularly proficient at sequestering calcium [22, 38] including oli-
goG's [39]. OligoG CF-5/20 has also been shown to act as a calcium-chelating agent
which can compensate for the impaired secretion of bicarbonate in CF intestinal
environment [40]. These studies showed that OligoG CF-5/20 worked by sequester-
ing calcium away from the packed mucins that are then secreted into the intestine
allowing the mucus to expand, mature and disperse [40].
Low molecular weight guluronic acid oligomers have also been shown to directly
bind mucins leading to the disruption of intermolecular interactions between mucins
in complex mucus polymer networks. This interaction is independent of the cation
chelating ability of these G-rich oligomers, and results in an increased pore size of
the mucin matrix, and altered mucus rheology [23, 24, 41]. Similarly, direct binding
of OligoG CF-5/20 to mucins has been demonstrated which modifies the mucin
surface charge and may explain its ability to reduce the viscoelastic properties of CF
sputum [42]. Further studies are needed to examine the effect of alginate oligomers
on additional intestinal alterations such as SIBO or changes in the intestinal micro-
biota and intestinal inflammation [43, 44].
CF is not the only disease where mucociliary clearance is impaired due to abnor-
mal mucus viscosity. Other chronic respiratory conditions such as chronic obstruc-
tive pulmonary disease (COPD) and sinusitis are also potential application areas
where the mucus rheology-modifying properties of alginate oligomers could have
clinical benefit.
10 Alginate Oligomers and Their Use as Active Pharmaceutical Drugs 245
Fig. 10.4 Microbial burden in rat soft tissue animal model after treatment with antibiotic (azithro-
mycin) alone or in combination with alginate oligomers. The infection challenge was a clinical
isolate of A. baumannii strain 307-0294 (Unpublished data from a study performed at Prof. Russo’s
lab at University of Buffalo (SUNY) on behalf of AlgiPharma AS)
246 P.D. Rye et al.
Fig. 10.5 Motility testing of Proteus mirabilis (NSM6) grown on ISO agar containing different
concentrations of alginate oligomers (Reprinted from [47], Copyright 2012 with permission from
American Society for Microbiology)
10 Alginate Oligomers and Their Use as Active Pharmaceutical Drugs 247
Fig. 10.6 Confocal laser scanning microscopy images of P. aeruginosa (NH57388A) biofilms
grown for 24 h (37oC) in Mueller-Hinton broth. Cells are visualized by live/dead staining (live cells
stained green, dead cells stained red). Untreated control biofilm (a); biofilm treated with alginate
oligomer (b). Scale bar = 0.05 μm
Fig. 10.8 Scanning electron microscopy (SEM) images of mixed biofilm-infected wounds,
untreated (a) or treated with alginate oligomers (b). Treatment with alginate oligomers shows a
marked reduction in bacteria and biofilm. SEM images were of biopsies taken at days 7, 14 and 35
postinoculation from multispecies biofilm-infected wounds. Scale bar = 5 μm. [62]
removal of the device (which is not always possible) and/or an increased use of
antibiotics. Depending on the medical device, this can have life-threatening impli-
cations for the patient. Since alginate oligomers have already been shown to effec-
tively disrupt established biofilm infections, their use in preventing biofilm
contamination of medical devices represents a significant area of clinical value.
Studies using alginate oligomers and their effects on oral pathogens appear to sup-
port this. Roberts et al. [61] showed that alginate oligomers coated onto dental
materials such as titanium could inhibit the attachment of Streptococcus mutans and
Porphyromonas gingivalis two important oral pathogens associated with dental
carious lesions and periodontitis. This highlights the potential value of using algi-
nate oligomers either as a preventative coating for dental implants/prostheses or
simply as part of a decontamination strategy prior to device implantation.
Studies investigating a wound care application demonstrated that alginate oligo-
mers formulated as a salve and applied to an infected wound could prevent the for-
mation of mixed species biofilms in a pig skin burn wound model (Fig. 10.8) [62].
Clearly there are several opportunities to further exploit the properties of alginate
oligomers in the coating of other medical devices, such as catheters, etc. Ventilator-
associated pneumonia (VAP) is a particularly serious lung infection, and the patho-
physiology of the infection is thought to be due to the endotracheal or tracheostomy
10 Alginate Oligomers and Their Use as Active Pharmaceutical Drugs 251
tube allowing free passage of bacteria into the lungs. P. aeruginosa is one of the
more common organisms that are associated with ventilator-associated pneumonia
as they are known to colonize and form biofilms within the endotracheal tube. Since
the effects of alginate oligomers on these organisms are well established, it is not
unreasonable to postulate that coating the surface of these tubes with alginate oligo-
mers could inhibit the accumulation of biofilm on the internal surface of these medi-
cal devices, thereby reducing the risk of potentially lethal lung infections.
10.3.5 Anticoagulation
10.4 Conclusions
Based on the studies to date, it is clear that alginate oligomers, with their excellent
safety profiles for oral and inhaled delivery, have considerable potential in a variety
of medical applications (Fig. 10.9). The indication that is most advanced in terms of
clinical studies is in the symptomatic treatment of CF. It is anticipated that the com-
bined properties of alginate oligomers will not only facilitate the clearance of the
viscous sputum characteristic of the disease but also potentiate the treatment of
chronic MDR lung infections that affect CF patients. The potential use of the algi-
nate oligomers in the treatment of infections is considerable, particularly in the
context of addressing MDR bacterial and fungal infections, where there are limited
therapeutic options. Moreover, in contrast to antibiotic treatments that induce resis-
tance, long-term studies with alginate oligomers have not shown any development
of resistance. This application of alginate oligomer technology is of particular
importance in the treatment of biofilm-related infections, such as chronic venous leg
ulcers, diabetic foot ulcers and periodontal disease, where the persistence of biofilm
contributes to the chronic inflammation and infection in these patients. The safety
profile and chemical characteristics of alginate oligomers facilitate their applica-
tion, formulation and delivery for these multiple therapeutic applications.
252 P.D. Rye et al.
Fig. 10.9 Summary diagram of biological properties associated with alginate oligomers
References
9. Lien SK, Sletta H, Ellingsen TE, Valla S, Correa E, Goodacre R, Vernstad K, Borgos SEF,
Bruheim P (2013) Investigating alginate production and carbon utilization in Pseudomonas
fluorescens SBW25 using mass spectrometry-based metabolic profiling. Metabolomics
9(2):403–417
10. Ertesvåg H (2015) Alginate-modifying enzymes: biological roles and biotechnological uses.
Front Microbiol 6:10
11. Ertesvåg H, Høidal HK, Hals IK, Rian A, Doseth B, Valla S (1995) A family of modular type
mannuronan C-5 epimerase genes controls alginate structure in Azotobacter vinelandii. Mol
Microbiol 16(4):719–731
12. Høidal HK, Svanem BIG, Gimmestad M, Valla S (2000) Mannuronan C-5 epimerases and
cellular differentiation of Azotobacter vinelandii. Environ Microbiol 2(1):27–38
13. Tøndervik A, Klinkenberg G, Aachmann FL, Svanem BIG, Ertesvåg H, Ellingsen TE, Valla
S, Skjåk-Bræk G, Sletta H (2013) Mannuronan C-5 epimerases suited for tailoring of spe-
cific alginate structures obtained by high-throughput screening of an epimerase mutant library.
Biomacromolecules 14(8):2657–2666
14. Fischl R, Bertelsen K, Gaillard F, Coelho S, Michel G, Klinger M, Boyen C, Czjzek M, Herve
C (2016) The cell-wall active mannuronan C5-epimerases in the model brown alga Ectocarpus:
from gene context to recombinant protein. Glycobiology 26(9):973–983
15. Ye NH, Zhang XW, Miao M, Fan X, Zheng Y, Xu D, Wang JF, Zhou L, Wang DS, Gao Y, Wang
YT, Shi WY, Ji PF, Li DM, Guan Z, Shao CW, Zhuang ZM, Gao ZW, Qi J, Zhao FQ (2015)
Saccharina genomes provide novel insight into kelp biology. Nat Commun 6
16. Inoue A, Satoh A, Morishita M, Tokunaga Y, Miyakawa T, Tanokura M, Ojima T (2016)
Functional heterologous expression and characterization of mannuronan C5-epimerase from
the brown alga Saccharina japonica. Algal Res 16:282–291
17. Aarstad OA, Tøndervik A, Sletta H, Skjåk-Bræk G (2012) Alginate sequencing: an analysis of
block distribution in alginates using specific alginate degrading enzymes. Biomacromolecules
13(1):106–116
18. Aarstad O, Strand BL, Klepp-Andersen LM, Skjåk-Bræk G (2013) Analysis of G-block
distributions and their impact on gel properties of in vitro epimerized mannuronan.
Biomacromolecules 14(10):3409–3416
19. Haug A, Larsen B, Smidsrød O (1967) Studeis on the sequence of uronic acid residues in
alginic acid. Acta Chem Scand 21:691–704
20. Haug A, Larsen B, Smidsrød O (1974) Uronic acid sequence in alginate from different sources.
Carbohydr Res 32:217–225
21. Campa C, Oust A, Skjåk-Bræk G, Paulsen BS, Paoletti S, Christensen BE, Ballance S (2004)
Determination of average degree of polymerisation and distribution of oligosaccharides in a
partially acid-hydrolysed homopolysaccharide: a comparison of four experimental methods
applied to mannuronan. J Chromatogr A 1026(1-2):271–281
22. Padol AM, Draget KI, Stokke BT (2016) Effects of added oligoguluronate on mechanical
properties of Ca – alginate – oligoguluronate hydrogels depend on chain length of the alginate.
Carbohydr Polym 147:234–242
23. Nordgård CT, Draget KI (2011) Oligosaccharides as modulators of rheology in complex
mucous systems. Biomacromolecules 12(8):3084–3090
24. Nordgård CT, Nonstad U, Olderøy MO, Espevik T, Draget KI (2014) Alterations in mucus
barrier function and matrix structure induced by guluronate oligomers. Biomacromolecules
15(6):2294–2300
25. Powell LC, Pritchard MF, Emanuel C, Onsøyen E, Rye PD, Wright CJ, Hill KE, Thomas DW
(2014) A nanoscale characterization of the interaction of a novel alginate oligomer with the cell
surface and motility of Pseudomonas aeruginosa. Am J Respir Cell Mol Biol 50(3):483–492
26. Powell LC, Sowedan A, Khan S, Wright CJ, Hawkins K, Onsøyen E, Myrvold R, Hill KE,
Thomas DW (2013) The effect of alginate oligosaccharides on the mechanical properties of
Gram-negative biofilms. Biofouling 29(4):413–421
254 P.D. Rye et al.
27. Draget KI (2016) Alginates: fundamental properties and food applications. Reference module
in food. Science:1–9
28. Kreda SM, Davis CW, Rose MC (2012) CFTR, mucins, and mucus obstruction in cystic fibro-
sis. Cold Spring Harb Perspect Med 2(9):32
29. Garcia MAS, Yang N, Quinton PM (2009) Normal mouse intestinal mucus release requires
cystic fibrosis transmembrane regulator-dependent bicarbonate secretion. J Clin Invest
119(9):2613–2622
30. Gustafsson JK, Ermund A, Ambort D, Johansson MEV, Nilsson HE, Thorell K, Hebert
H, Sjovall H, Hansson GC (2012) Bicarbonate and functional CFTR channel are required
for proper mucin secretion and link cystic fibrosis with its mucus phenotype. J Exp Med
209(7):1263–1272
31. Yang N, Garcia MA, Quinton PM (2013) Normal mucus formation requires cAMP-dependent
HCO3- secretion and Ca2+-mediated mucin exocytosis. J Physiol 591(18):4581–4593
32. De Lisle RC, Borowitz D (2013) The cystic fibrosis intestine. Cold Spring Harb Perspect Med
3(9)
33. Mascarenhas MR (2003) Treatment of gastrointestinal problems in cystic fibrosis. Curr Treat
Options Gastroenterol 6(5):427–441
34. Vitko M, Valerio DM, Rye PD, Onsøyen E, Myrset AH, Dessen A, Drumm ML, Hodges CA
(2016) A novel guluronate oligomer improves intestinal transit and survival in cystic fibrosis
mice. J Cyst Fibros
35. Ambort D, Johansson ME, Gustafsson JK, Nilsson HE, Ermund A, Johansson BR, Koeck PJ,
Hebert H, Hansson GC (2012) Calcium and pH-dependent packing and release of the gel-
forming MUC2 mucin. Proc Natl Acad Sci U S A 109(15):5645–5650
36. Ermund A, Meiss LN, Gustafsson JK, Hansson GC (2015) Hyper-osmolarity and calcium
chelation: effects on cystic fibrosis mucus. Eur J Pharmacol 764:109–117
37. Braccini I, Grasso RP, Perez S (1999) Conformational and configurational features of acidic
polysaccharides and their interactions with calcium ions: a molecular modeling investigation.
Carbohydr Res 317(1-4):119–130
38. Jørgensen TE, Sletmoen M, Draget KI, Stokke BT (2007) Influence of oligoguluronates on
alginate gelation, kinetics, and polymer organization. Biomacromolecules 8(8):2388–2397
39. Bowman KA, Aarstad OA, Nakamura M, Stokke BT, Skjåk-Bræk G, Round AN (2016) Single
molecule investigation of the onset and minimum size of the calcium-mediated junction zone
in alginate. Carbohydr Polym 148:52–60
40. Ermund A, Recktenwald CV, Skjåk-Bræk G, Meiss LN, Onsøyen E, Rye PD, Dessen A,
Myrset AH, Gustafsson JK, Hansson GC (2017) OligoG CF 5/20 normalizes cystic fibrosis
mucus by chelating calcium
41. Sletmoen M, Maurstad G, Nordgård CT, Draget KI, Stokke BT (2012) Oligoguluronate
induced competitive displacement of mucin-alginate interactions: relevance for mucolytic
function. Soft Matter 8(32):8413–8421
42. Pritchard MF, Powell LC, Menzies GE, Lewis PD, Hawkins K, Wright C, Doull I, Walsh TR,
Onsøyen E, Dessen A, Myrvold R, Rye PD, Myrset AH, Stevens HN, Hodges LA, MacGregor
G, Neilly JB, Hill KE, Thomas DW (2016) A new class of safe oligosaccharide polymer ther-
apy to modify the mucus barrier of chronic respiratory disease. Mol Pharm 13(3):863–872
43. Bazett M, Honeyman L, Stefanov AN, Pope CE, Hoffman LR, Haston CK (2015) Cystic
fibrosis mouse model-dependent intestinal structure and gut microbiome. Mamm Genome
26(5-6):222–234
44. De Lisle RC (2007) Altered transit and bacterial overgrowth in the cystic fibrosis mouse small
intestine. Am J Physiol Gastrointest Liver Physiol 293(1):G104–G111
45. Laxminarayan R, Duse A, Wattal C, Zaidi AKM, Wertheim HFL, Sumpradit N, Vlieghe E,
Hara GL, Gould IM, Goossens H, Greko C, So AD, Bigdeli M, Tomson G, Woodhouse W,
Ombaka E, Peralta AQ, Qamar FN, Mir F, Kariuki S, Bhutta ZA, Coates A, Bergstrom R,
Wright GD, Brown ED, Cars O (2013) Antibiotic resistance-the need for global solutions.
Lancet Infect Dis 13(12):1057–1098
10 Alginate Oligomers and Their Use as Active Pharmaceutical Drugs 255
46. Hengzhuang W, Song Z, Ciofu O, Onsøyen E, Rye PD, Hoiby N (2016) OligoG CF-5/20
disruption of mucoid pseudomonas aeruginosa biofilm in a murine lung infection model.
Antimicrob Agents Chemother 60(5):2620–2626
47. Khan S, Tøndervik A, Sletta H, Klinkenberg G, Emanuel C, Onsøyen E, Myrvold R, Howe
RA, Walsh TR, Hill KE, Thomas DW (2012) Overcoming drug resistance with alginate oligo-
saccharides able to potentiate the action of selected antibiotics. Antimicrob Agents Chemother
56(10):5134–5141
48. Russo TA, Beanan JM, Olson R, MacDonald U, Luke NR, Gill SR, Campagnari AA (2008)
Rat pneumonia and soft-tissue infection models for the study of Acinetobacter baumannii biol-
ogy. Infect Immun 76(8):3577–3586
49. Powell LC, Pritchard MF, Emanuel C, Khan S, Sletta H, Tøndervik A, Klinkenberg G, Onsøyen
ER, Myrvold R, Rye P, Hill KE, Thomas DW (2013) Characterization of the effect of a novel
antifungal alginate oligomer on fungal hyphae formation. Pediatr Pulmonol 48:329–329
50. Pritchard MF, Powell LC, Onsøyen E, Rye PD, Hill KE, Thomas DW (2014) Utilization of a
recombinant in vitro epithelial model to study the effect of novel therapies on microbial colo-
nization and invasion of the epidermis. Wound Repair Regen 22(5):A95
51. Tøndervik A, Sletta H, Klinkenberg G, Emanuel C, Powell LC, Pritchard MF, Khan S, Craine
KM, Onsøyen E, Rye PD, Wright C, Thomas DW, Hill KE (2014) Alginate oligosaccharides
inhibit fungal cell growth and potentiate the activity of antifungals against candida and asper-
gillus spp. PLoS One 9(11)
52. Kohler T, Curty LK, Barja F, van Delden C, Pechere JC (2000) Swarming of Pseudomonas
aeruginosa is dependent on cell-to-cell signaling and requires flagella and pili. J Bacteriol
182(21):5990–5996
53. Pritchard MF, Ferguson E, Powell L, Onsøyen E, Rye P, Hill K, Thomas DW (2015)
Characterization of the in vitro interaction of an alginate oligosaccharide (OligoG CF-5/20)
with Pseudomonal biofilms using fluorescent labelling and quantitative image analysis. Pediatr
Pulmonol 50(S41):S295
54. Pritchard MF, Powell L, Jack AA, Powell K, Onsøyen E, Rye PD, Beck PD, Hill KE, Thomas
DW (2016) OligoG CF-5/20 induces microcolony disruption and potentiates the activity of
colistin against multidrug resistant Pseudomonal biofilms. Pediatr Pulmonol 51(S45):S285
55. Sherbrock-Cox V, Russell NJ, Gacesa P (1984) The purification and chemical characterisation
of the alginate present in extracellular material produced by mucoid strains of Pseudomonas
aeruginosa. Carbohydr Res 135(1):147–154
56. Pritchard MF, Powell L, Khan S, Griffiths PC, Mansour OT, Schweins R, Beck K, Buurma NJ,
Dempsey CE, Wright CJ, Rye PD, Hill KE, Thomas DW, Ferguson EL (2017) The antimicro-
bial effects of the alginate oligomer OligoG CF-5/20 are independent of direct bacterial cell
membrane disruption. Sci Rep. (in press). https://doi.org/10.1038/srep44731
57. Jack AA, Khan S, Pritchard MF, Beck K, Onsøyen E, Rye PD, Thomas DW, Hill KE (2016)
OligoG CF-5/20 modifies the Las and Rhl signalling pathways in a time dependent manner in
Pseudomonas aeruginosa PAO1. Pediatr Pulmonol 51(S45):S333–S334
58. Ryall B, Carrara M, Zlosnik JE, Behrends V, Lee X, Wong Z, Lougheed KE, Williams HD
(2014) The mucoid switch in Pseudomonas aeruginosa represses quorum sensing systems
and leads to complex changes to stationary phase virulence factor regulation. PLoS One
9(5):e96166
59. Hall S, McDermott C, Anoopkumar-Dukie S, McFarland AJ, Forbes A, Perkins AV, Davey
AK, Chess-Williams R, Kiefel MJ, Arora D, Grant GD (2016) Cellular effects of pyocyanin, a
secreted virulence factor of pseudomonas aeruginosa. Toxins (Basel) 8(8)
60. Chang CY, Krishnan T, Wang H, Chen Y, Yin WF, Chong YM, Tan LY, Chong TM, Chan KG
(2014) Non-antibiotic quorum sensing inhibitors acting against N-acyl homoserine lactone
synthase as druggable target. Sci Rep 4(4275)
61. Roberts JL, Khan S, Emanuel C, Powell LC, Pritchard MF, Onsøyen E, Myrvold R, Thomas
DW, Hill KE (2013) An in vitro study of alginate oligomer therapies on oral biofilms. J Dent
41(10):892–899
256 P.D. Rye et al.
Rosalia Crupi and Salvatore Cuzzocrea
R. Crupi
Department of Chemical, Biological, Pharmacological and Environmental Sciences,
University of Messina, Viale Ferdinando Stagno D’Alcontres 31, Messina 98166, Italy
S. Cuzzocrea (*)
Department of Chemical, Biological, Pharmacological and Environmental Sciences,
University of Messina, Viale Ferdinando Stagno D’Alcontres 31, Messina 98166, Italy
Department of Pharmacological and Physiological Science, Saint Louis University,
Saint Louis, MO, USA
e-mail: salvator@unime.it
11.1 Introduction
11.2 D
-Mannuronic Acid (M2000): Anti-inflammatory
Properties
Fig. 11.1 Chemical
structure of β-D-
mannuronic acid
published in near future). It should be noted that cardiopathy could be one of the
most important NSAID side effects.
In RA, the immune system is misdirected and attacks the joints. This misdirect-
ing could promote inflammation in joints and also in other various organs and body
tissues, occasionally. Generally, the conventional disease-modifying antirheumatic
drugs (DMARDs) and TNF-α blockers were used to decrease both disease activity
and progression in patients. In this clinical trial, patients were allowed to use their
routine medications, which included methotrexate (15–20 mg weekly), hydroxy-
chloroquine (400 mg daily), and steroids (5–15 mg daily), and also 7 out of 12
patients received a subcutaneous injection of etanercept (25 mg twice weekly).
However, patients were forbidden from using NSAIDs or other pharmacologic
treatment during this 12-week follow-up. Based on the preclinical assessment, a
minimum dosage (18 mg/kg/d) of M2000 was provided in a gelatinized capsule
(500 mg of M2000) for oral administration. Finally, the M2000 capsule (500 mg)
was prescribed twice daily for 12 weeks. The patient status showed an improvement
after 2 weeks and also continued during the treatment course. The mean of disease
activity (DAS28), morning stiffness, rheumatoid factor (RF), and C-reactive protein
(CRP) had met a significant reduction after 12 weeks of treatment. In addition,
improvement was observed in other clinical laboratory tests, including anti-CCP,
anti-dsDNA, and ESR levels, which returned to the normal range. After 12 weeks of
therapy with M2000, IL-17 and RORγt gene expressions in patients’ PBMCs were
decreased by 22.39- and 2.36-fold, respectively, when compared with the gene
expressions of the patients before therapy. However, several patients do not respond
to these treatments and present a diminished response to them over time. Furthermore,
RA patients always suffer from the adverse effects of these chemicals and biological
drugs. Therefore, M2000 as a new and natural anti-inflammatory agent was used
during a 3-month clinical trial and showed a suitable response in the proposed gene
expression and clinical as well as paraclinical results. Findings indicated that IL-17
and its transcription factor RORγt displayed a significant diminution in PBMCs.
Previous evidence indicated the critical role of IL-17 in RA and that suitable thera-
pies were able to decrease its level. The increase of IL-17 production and upregula-
tion of Th17 cells which was characterized by expression of the RORγt are the most
common features of RA [71]. Regarding IL-17-producing CD4+ T (Th17) cells as
unique T-helper cells associated with T-cell-mediated tissue injury, they could pro-
mote inflammation by producing pro-inflammatory cytokines and chemokines in
autoimmune disease [72, 73]. The IL-17 level after treatment with M2000 signifi-
cantly decreased 22.39-fold the level of these cytokines as compared to before treat-
ment. The results of this trial according with the other studies reported the potential
role of IL-17 in mediating joint damage and the ability of IL-17 to induce collagen
release from cartilage [74]. In this study, there were significant correlations between
the gene expression results and clinical and paraclinical assessments. The levels of
these cytokines were in conformity with the disease activity, tender joint, and swell-
ing, all of which showed a reduction in mean after treatment. Moreover, the range
of ESR and CRP, which was quite high before M2000, faced a significant decrease
and was back to the normal range.
264 R. Crupi and S. Cuzzocrea
11.3 Conclusion
Several literature data revealed that M2000 therapy represents an important tool in
the management of many inflammatory diseases; moreover M2000 is the first novel
designed NSAID with the lowest molecular weight and therapeutic effects, and for
this reason it could be recommended in an extensive scale as the safest drug for
decreasing anti-inflammatory reactions.
References
1. Hay ID, Ur Rehman Z, Moradali MF, Wang Y, Rehm BH (2013) Microbial alginate pro-
duction, modification and its applications. Microb Biotechnol 6(6):637–650. https://doi.
org/10.1111/1751-7915.12076
2. Fabich HT, Vogt SJ, Sherick ML, Seymour JD, Brown JR, Franklin MJ, Codd SL (2012)
Microbial and algal alginate gelation characterized by magnetic resonance. J Biotechnol
161(3):320–327. https://doi.org/10.1016/j.jbiotec.2012.04.016
11 Mannuronic Acid as an Anti-inflammatory Drug 265
3. Lee KY, Mooney DJ (2012) Alginate: properties and biomedical applications. Prog Polym Sci
37(1):106–126. https://doi.org/10.1016/j.progpolymsci.2011.06.003
4. Kaushik AY, Tiwari AK, Gaur A (2015) Role of excipients and polymeric advancements in
preparation of floating drug delivery systems. International journal of pharmaceutical investi-
gation 5(1):1–12. https://doi.org/10.4103/2230-973X.147219
5. Barati A, Jamshidi AR, Ahmadi H, Aghazadeh Z, Mirshafiey A (2017) Effects of beta-d-
mannuronic acid, as a novel non-steroidal anti-inflammatory medication within immunosup-
pressive properties, on IL17, RORgammat, IL4 and GATA3 gene expressions in rheumatoid
arthritis patients. Drug Des Devel Ther 11:1027–1033. https://doi.org/10.2147/DDDT.S129419
6. Rehm BH, Valla S (1997) Bacterial alginates: biosynthesis and applications. Appl Microbiol
Biotechnol 48(3):281–288
7. Skjak-Braek G, Grasdalen H, Larsen B (1986) Monomer sequence and acetylation pattern in
some bacterial alginates. Carbohydr Res 154:239–250
8. Rhein-Knudsen N, Ale MT, Meyer AS (2015) Seaweed hydrocolloid production: an update
on enzyme assisted extraction and modification technologies. Mar Drugs 13(6):3340–3359.
https://doi.org/10.3390/md13063340
9. Sidiropoulos PI, Hatemi G, Song IH, Avouac J, Collantes E, Hamuryudan V, Herold M, Kvien
TK, Mielants H, Mendoza JM, Olivieri I, Ostergaard M, Schachna L, Sieper J, Boumpas DT,
Dougados M (2008) Evidence-based recommendations for the management of ankylosing
spondylitis: systematic literature search of the 3E Initiative in Rheumatology involving a broad
panel of experts and practising rheumatologists. Rheumatology 47(3):355–361. https://doi.
org/10.1093/rheumatology/kem348
10. Fattahi MJ, Abdollahi M, Agha Mohammadi A, Rastkari N, Khorasani R, Ahmadi H, Tofighi
Zavareh F, Sedaghat R, Tabrizian N, Mirshafiey A (2015) Preclinical assessment of beta-d-
mannuronic acid (M2000) as a non-steroidal anti-inflammatory drug. Immunopharmacol
Immunotoxicol 37(6):535–540. https://doi.org/10.3109/08923973.2015.1113296
11. Mirshafiey A, Rehm B, Sotoude M, Razavi A, Abhari RS, Borzooy Z (2007) Therapeutic
approach by a novel designed anti-inflammatory drug, M2000, in experimental immune
complex glomerulonephritis. Immunopharmacol Immunotoxicol 29(1):49–61. https://doi.
org/10.1080/08923970701282387
12. Lodish HF, Zhou B, Liu G, Chen CZ (2008) Micromanagement of the immune system by
microRNAs. Nat Rev Immunol 8(2):120–130. https://doi.org/10.1038/nri2252
13. Bartel DP (2004) MicroRNAs: genomics, biogenesis, mechanism, and function. Cell
116(2):281–297
14. Ambros V (2004) The functions of animal microRNAs. Nature 431(7006):350–355. https://
doi.org/10.1038/nature02871
15. Martinelli-Boneschi F, Fenoglio C, Brambilla P, Sorosina M, Giacalone G, Esposito F,
Serpente M, Cantoni C, Ridolfi E, Rodegher M, Moiola L, Colombo B, De Riz M, Martinelli
V, Scarpini E, Comi G, Galimberti D (2012) MicroRNA and mRNA expression profile screen-
ing in multiple sclerosis patients to unravel novel pathogenic steps and identify potential bio-
markers. Neurosci Lett 508(1):4–8. https://doi.org/10.1016/j.neulet.2011.11.006
16. Dai R, Ahmed SA (2011) MicroRNA, a new paradigm for understanding immunoregulation,
inflammation, and autoimmune diseases. Transl Res J Lab Clin Med 157(4):163–179. https://
doi.org/10.1016/j.trsl.2011.01.007
17. Tufekci KU, Oner MG, Genc S, Genc K (2010) MicroRNAs and multiple sclerosis. Autoimmun
Dis 2011:807426. https://doi.org/10.4061/2011/807426
18. Lofgren SE, Frostegard J, Truedsson L, Pons-Estel BA, D’Alfonso S, Witte T, Lauwerys BR,
Endreffy E, Kovacs L, Vasconcelos C, Martins da Silva B, Kozyrev SV, Alarcon-Riquelme
ME (2012) Genetic association of miRNA-146a with systemic lupus erythematosus in
Europeans through decreased expression of the gene. Genes Immun 13(3):268–274. https://
doi.org/10.1038/gene.2011.84
266 R. Crupi and S. Cuzzocrea
19. Nahid MA, Pauley KM, Satoh M, Chan EK (2009) miR-146a is critical for endotoxin-induced
tolerance: Implication in innate immunity. J Biol Chem 284(50):34590–34599. https://doi.
org/10.1074/jbc.M109.056317
20. Nahid MA, Satoh M, Chan EK (2011) Mechanistic role of microRNA-146a in endotoxin-
induced differential cross-regulation of TLR signaling. J Immunol 186(3):1723–1734. https://
doi.org/10.4049/jimmunol.1002311
21. Rom S, Rom I, Passiatore G, Pacifici M, Radhakrishnan S, Del Valle L, Pina-Oviedo S, Khalili
K, Eletto D, Peruzzi F (2010) CCL8/MCP-2 is a target for mir-146a in HIV-1-infected human
microglial cells. FASEB J 24(7):2292–2300. https://doi.org/10.1096/fj.09-143503
22. Curtale G, Citarella F, Carissimi C, Goldoni M, Carucci N, Fulci V, Franceschini D, Meloni
F, Barnaba V, Macino G (2010) An emerging player in the adaptive immune response:
microRNA-146a is a modulator of IL-2 expression and activation-induced cell death in T lym-
phocytes. Blood 115(2):265–273. https://doi.org/10.1182/blood-2009-06-225987
23. Atarod S, Ahmed MM, Lendrem C, Pearce KF, Cope W, Norden J, Wang XN, Collin M,
Dickinson AM (2016) miR-146a and miR-155 expression levels in acute graft-versus-host
disease incidence. Front Immunol 7:56. https://doi.org/10.3389/fimmu.2016.00056
24. Pauley KM, Satoh M, Chan AL, Bubb MR, Reeves WH, Chan EK (2008) Upregulated miR-
146a expression in peripheral blood mononuclear cells from rheumatoid arthritis patients.
Arthritis Res Ther 10(4):R101. https://doi.org/10.1186/ar2493
25. Sonkoly E, Stahle M, Pivarcsi A (2008) MicroRNAs: novel regulators in skin inflammation.
Clin Exp Dermatol 33(3):312–315. https://doi.org/10.1111/j.1365-2230.2008.02804.x
26. Jensen LE, Muzio M, Mantovani A, Whitehead AS (2000) IL-1 signaling cascade in liver
cells and the involvement of a soluble form of the IL-1 receptor accessory protein. J Immunol
164(10):5277–5286
27. Ye H, Arron JR, Lamothe B, Cirilli M, Kobayashi T, Shevde NK, Segal D, Dzivenu OK,
Vologodskaia M, Yim M, Du K, Singh S, Pike JW, Darnay BG, Choi Y, Wu H (2002) Distinct
molecular mechanism for initiating TRAF6 signalling. Nature 418(6896):443–447. https://doi.
org/10.1038/nature00888
28. Rehm BH (1998) Alginate lyase from Pseudomonas aeruginosa CF1/M1 prefers the hexameric
oligomannuronate as substrate. FEMS Microbiol Lett 165(1):175–180
29. Jackson CJ, Nguyen M (1997) Human microvascular endothelial cells differ from macrovas-
cular endothelial cells in their expression of matrix metalloproteinases. Int J Biochem Cell
Biol 29(10):1167–1177
30. Mirshafiey A, Rehm B, Abhari RS, Borzooy Z, Sotoude M, Razavi A (2007) Production of
M2000 (beta-d-mannuronic acid) and its therapeutic effect on experimental nephritis. Environ
Toxicol Pharmacol 24(1):60–66. https://doi.org/10.1016/j.etap.2007.02.002
31. Saba R, Sorensen DL, Booth SA (2014) MicroRNA-146a: a dominant, negative regulator of
the innate immune response. Front Immunol 5:578. https://doi.org/10.3389/fimmu.2014.00578
32. Sabroe I, Parker LC, Dower SK, Whyte MK (2008) The role of TLR activation in inflamma-
tion. J Pathol 214(2):126–135. https://doi.org/10.1002/path.2264
33. O’Connell RM, Rao DS, Baltimore D (2012) microRNA regulation of inflammatory responses.
Annu Rev Immunol 30:295–312. https://doi.org/10.1146/annurev-immunol-020711-075013
34. Stanczyk J, Pedrioli DM, Brentano F, Sanchez-Pernaute O, Kolling C, Gay RE, Detmar M,
Gay S, Kyburz D (2008) Altered expression of MicroRNA in synovial fibroblasts and syno-
vial tissue in rheumatoid arthritis. Arthritis Rheum 58(4):1001–1009. https://doi.org/10.1002/
art.23386
35. Tang Y, Luo X, Cui H, Ni X, Yuan M, Guo Y, Huang X, Zhou H, de Vries N, Tak PP, Chen S,
Shen N (2009) MicroRNA-146A contributes to abnormal activation of the type I interferon
pathway in human lupus by targeting the key signaling proteins. Arthritis Rheum 60(4):1065–
1075. https://doi.org/10.1002/art.24436
36. Maitra U, Davis S, Reilly CM, Li L (2009) Differential regulation of Foxp3 and IL-17 expres-
sion in CD4 T helper cells by IRAK-1. J Immunol 182(9):5763–5769. https://doi.org/10.4049/
jimmunol.0900124
11 Mannuronic Acid as an Anti-inflammatory Drug 267
37. Deng C, Radu C, Diab A, Tsen MF, Hussain R, Cowdery JS, Racke MK, Thomas JA (2003)
IL-1 receptor-associated kinase 1 regulates susceptibility to organ-specific autoimmunity.
J Immunol 170(6):2833–2842
38. Wu H, Arron JR (2003) TRAF6, a molecular bridge spanning adaptive immunity, innate immu-
nity and osteoimmunology. BioEssays 25(11):1096–1105. https://doi.org/10.1002/bies.10352
39. Mirshafiey A, Matsuo H, Nakane S, Rehm BH, Koh CS, Miyoshi S (2005) Novel immu-
nosuppressive therapy by M2000 in experimental multiple sclerosis. Immunopharmacol
Immunotoxicol 27(2):255–265. https://doi.org/10.1081/IPH-200067751
40. Mirshafiey A, Cuzzocrea S, Rehm B, Mazzon E, Saadat F, Sotoude M (2005) Treatment of
experimental arthritis with M2000, a novel designed non-steroidal anti-inflammatory drug.
Scand J Immunol 61(5):435–441. https://doi.org/10.1111/j.1365-3083.2005.01594.x
41. Mirshafiey A, Cuzzocrea S, Rehm BH, Matsuo H (2005) M2000: a revolution in pharmacol-
ogy. Med Sci Monit Int Med J Exp Clin Res 11(8):PI53–PI63
42. Pourgholi F, Hajivalili M, Razavi R, Esmaeili S, Baradaran B, Movasaghpour AA, Sadreddini S,
Goodarzynejad H, Mirshafiey A, Yousefi M (2017) The role of M2000 as an anti-inflammatory
agent in toll-like receptor 2/microRNA-155 pathway. Avicenna J Med Biotechnol 9(1):8–12
43. Aletaha S, Haddad L, Roozbehkia M, Bigdeli R, Asgary V, Mahmoudi M, Mirshafiey A (2017)
M2000 (beta-D-Mannuronic Acid) as a novel antagonist for blocking the TLR2 and TLR4
downstream signalling pathway. Scand J Immunol 85(2):122–129. https://doi.org/10.1111/
sji.12519
44. O’Callaghan CA (2004) Renal manifestations of systemic autoimmune disease: diagno-
sis and therapy. Best Pract Res Clin Rheumatol 18(3):411–427. https://doi.org/10.1016/j.
berh.2004.03.002
45. Fogo AB (2003) Quiz page. Acute interstitial nephritis and minimal change disease lesion,
caused by NSAID injury. Am J Kidney Dis 42(2):A41–E41
46. Reinhold SW, Fischereder M, Riegger GA, Kramer BK (2003) Acute renal failure after admin-
istration of a single dose of a highly selective COX-2 inhibitor. Clin Nephrol 60(4):295–296
47. Basivireddy J, Jacob M, Pulimood AB, Balasubramanian KA (2004) Indomethacin-induced
renal damage: role of oxygen free radicals. Biochem Pharmacol 67(3):587–599. https://doi.
org/10.1016/j.bcp.2003.09.023
48. Stollberger C, Finsterer J (2004) Side effects of conventional nonsteroidal anti-inflammatory
drugs and celecoxib: more similarities than differences. South Med J 97(2):209. https://doi.
org/10.1097/01.SMJ.0000093569.26036.27
49. Perazella MA (2003) Drug-induced renal failure: update on new medications and unique
mechanisms of nephrotoxicity. Am J Med Sci 325(6):349–362
50. Kaiser A (2003) Diclofenac caused renal insufficiency. A case illustrating the necessity of
pharmaceutical intervention and care. Med Monatsschr Pharm 26(11):384–388
51. Lenz O, Elliot SJ, Stetler-Stevenson WG (2000) Matrix metalloproteinases in renal develop-
ment and disease. J Am Soc Nephrol 11(3):574–581
52. Chadban S (2001) Glomerulonephritis recurrence in the renal graft. J Am Soc Nephrol
12(2):394–402
53. Marti HP (2002) The role of matrix metalloproteinases in the activation of mesangial cells.
Transpl Immunol 9(2–4):97–100
54. Lovett DH, Johnson RJ, Marti HP, Martin J, Davies M, Couser WG (1992) Structural char-
acterization of the mesangial cell type IV collagenase and enhanced expression in a model of
immune complex-mediated glomerulonephritis. Am J Pathol 141(1):85–98
55. Harendza S, Schneider A, Helmchen U, Stahl RA (1999) Extracellular matrix deposition and
cell proliferation in a model of chronic glomerulonephritis in the rat. Nephrol Dial Transplant
14(12):2873–2879
56. Mirshafiey A, Rehm BH, Sahmani AA, Naji A, Razavi A (2004) M-2000, as a new anti-
inflammatory molecule in treatment of experimental nephrosis. Immunopharmacol
Immunotoxicol 26(4):611–619. https://doi.org/10.1081/IPH-200042362
268 R. Crupi and S. Cuzzocrea