Reviews: Synthesis and Applications of Biomedical and Pharmaceutical Polymers Via Click Chemistry Methodologies

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 16

NOVEMBER 2009

Volume 20, Number 11


 Copyright 2009 by the American Chemical Society

REVIEWS

Synthesis and Applications of Biomedical and Pharmaceutical Polymers via


Click Chemistry Methodologies
Maarten van Dijk,†,‡ Dirk T. S. Rijkers,‡ Rob M. J. Liskamp,‡ Cornelus F. van Nostrum,† and Wim E. Hennink*,†
Department of Pharmaceutics and Department of Medicinal Chemistry and Chemical Biology, Utrecht Institute for
Pharmaceutical Sciences, Utrecht University, P.O. Box 80082, 3508 TB Utrecht, The Netherlands. Received February 26, 2009;
Revised Manuscript Received June 19, 2009

In this review, the synthesis and application of biomedical and pharmaceutical polymers synthesized via the
copper(I)-catalyzed alkyne-azide cycloaddition, the thiol-ene reaction, or a combination of both click reactions
are discussed. Since the introduction of both “click” methods, numerous articles have disclosed new approaches
for the synthesis of polymers with different architectures, e.g., block and graft copolymers, dendrimers, and
hydrogels, for pharmaceutical and biomedical applications. By describing selected examples, an overview is given
of the possibilities and limitations that these two “click” methods may offer.

INTRODUCTION focus on the different click chemistry strategies to synthesize


this class of polymers and will describe some selected examples
The use of polymers for biomedical and pharmaceutical
in more detail to highlight their applications in the biomedical
applications has gained an enormous impact during the past
and pharmaceutical field.
decades. Polymers can be applied in drug delivery systems, as
scaffolds for tissue-engineering and -repair, and as novel In 2001, Sharpless and co-workers coined the concept of
biomaterials (1). These applications have led to an increasing “click chemistry” (2) to classify a particular set of nearly perfect
demand of well-defined polymers with tailorable properties. The reactions, among others, the Cu(I)-catalyzed 1,3-dipolar cy-
emergence of the “click chemistry” concept greatly facilitated cloaddition, vide infra. By means of the “click reaction” concept,
the synthesis of these polymers. Click chemistry provides very large (bio)macromolecules can be synthesized by coupling small
attractive possibilities for (bio)conjugation reactions, because building blocks via heteroatom-containing linkages. Such a
it can be performed at ambient conditions with readily available coupling reaction should meet several criteria: it should be
starting materials. In recent years, click chemistry has also been modular, high-yielding, and generate only harmless side prod-
applied for the synthesis of polymers with different architectures ucts, and it should be carried out under mild reaction conditions,
(such as block and graft copolymers), including polymers with preferentially in the presence of many other functional groups
pharmaceutical and biomedical applications. This review will using readily available starting materials and reagents.
There are several well-known reactions that comply with the
* Author to whom correspondence should be addressed. Tel.: +31 “click chemistry” concept, including the hetero-Diels-Alder
302536964, fax: +31 302517839. E-mail: W.E.Hennink@uu.nl. reaction (3), the thiol-ene coupling (4), the Staudinger

Department of Pharmaceutics. ligation (5-7), native chemical ligation (8, 9), the amidation

Department of Medicinal Chemistry and Chemical Biology. reaction between thio acids and sulfonyl azides (sulfo-click) (10-
10.1021/bc900087a CCC: $40.75  2009 American Chemical Society
Published on Web 07/17/2009
2002 Bioconjugate Chem., Vol. 20, No. 11, 2009 van Dijk et al.

Table 1. List of Reviews Dealing with the Synthesis of Large can be especially troublesome when it is difficult to remove
(Multivalent) Systems and Macromolecules by the CuAAC the copper catalyst from the synthesized polymers. Therefore,
An overview of polymer synthesis and Binder et al. (19, 20) the use of ring strain has been investigated as an alternative for
modification via the CuAAC. Cu(I) as catalyst to activate acetylenes. Wittig and Krebs (43)
Review about click polymerization Meldal (21) had already reported in 1961 that the reaction between a
reactions as well as ligation and
cyclooctyne and phenyl azide resulted in the formation of a
functionalization of polymers.
Reviews dealing with the CuAAC in Nandivada et al. (22), triazole functionality (Scheme 1C), and in 2004, Bertozzi and
material science. Lutz (23), and Johnson co-workers (44) demonstrated that the [3 + 2] cycloaddition of
et al. (24) azides and cyclooctyne derivatives occurs under physiological
Review about synthesis of peptidomimetics Angell et al. (25) conditions. However, the first generation of cyclooctynes was
via CuAAC. hampered with a relatively low reactivity toward azides, as
Overview of carbohydrates and Pieters et al. (26)
compared to the CuAAC, resulting in long reaction times and
peptide-based dendrimers/polymers.
Reviews about the construction of Dirks et al. (27) Le lower coupling efficiencies. To improve their reactivity, Boons
biohybrid materials. Droumaguet et al. (28) and co-workers (45) synthesized several 4-dibenzocyclooctynols
and Lutz (29) (Scheme 1D). By introducing aromatic moieties to the cyclooc-
Review with emphasis on the synthesis of Fournier et al. (30) tyne, additional ring strain was created, and also, better
well-defined polymer architectures. conjugation was achieved as an additional factor to increase
Review about various cross-linking Read et al. (31)
the reactivity of the alkyne. However, the first generation of
methods of micelles including CuAAC.
An overview of the synthesis of Lecomte et al. (32) these cyclooctynes and 4-dibenzocyclooctynols was rather
biodegradable polyesters by ring-opening insoluble in water, and in an attempt to improve the reactivity
polymerization and CuAAC. and water solubility, Bertozzi and co-workers designed and
A review about copper-free azide-alkyne Lutz (33) synthesized a second generation of difluorinated cyclooctynes
cycloaddition reactions. (46, 47) (Scheme 1E). These difluorinated cyclooctynes possess
Other important reviews dealing with Meldal and Tornøe. (34),
similar reaction kinetics as the Cu(I)-catalyzed cycloaddition
CuAAC. Bock et al. (35), Moses
et al. (36), Moorhouse reaction. Unfortunately, these difluorinated cyclooctynes are
et al. (37) rather difficult to synthesize (48). A slightly different approach
toward Cu-free click reactions was explored by Rutjes and co-
15), and presently the most popular and already mentioned copper(I)- workers (49). They used a tandem [3 + 2] cycloaddition-retro-
catalyzed alkyne-azide cycloaddition (CuAAC) (16-18). Diels-Alder ligation method in which trifluoromethyl-substi-
Since a large number of recent original publications and tuted oxanorbornadiene derivatives react with an azide to form
(specialized) reviews have been published during the past years a triazole linkage (Scheme 1F). This approach was successfully
on CuAAC (for a detailed list, see Table 1), we will focus in applied for labeling proteins and to the synthesis of pegylated
this review on the synthesis of polymers and/or their postsyn- oligopeptides in various media at ambient temperature. Although
thetic modification/functionalization by the CuAAC and the the trifluoromethyl-substituted oxanorbornadiene derivatives are
thiol-ene reaction. In addition, we will describe some selected much easier to synthesize than difluorinated cyclooctynes, they
examples of biomedically and pharmaceutically relevant poly- have a lower reactivity toward azides compared to the diflu-
mers that have been synthesized by these two click reactions. orinated cyclooctynes. Recently, an overview of the development
This review covers the literature until April, 2009. and perspective of copper-free azide-alkyne cycloadditions was
given by Lutz (33).
GENERAL CONSIDERATIONS While the CuAAC generally yields a 1,4-disubstituted 1,2,3-
triazole, in some cases, however, (e.g., for the synthesis of
The Cu(I) catalysis of the well-known 1,3-dipolar cycload-
certain enzyme inhibitors) the 1,5-disubstituted triazole is
dition reaction (CuAAC) between an azide and an alkyne (38)
preferred (50). In order to obtain exclusively the 1,5-disubsti-
was discovered in 2002 independently by the groups of
tuted 1,2,3-triazoles from organic azides and alkynes, the groups
Meldal (16, 17) and Sharpless (18). In this cycloaddition
of Fokin and Jia (51) used a ruthenium(II) catalyst (Scheme
reaction, an organic azide reacts with an alkyne to form a triazole
1B). This Ru-catalyzed process, RuAAC, exhibits a good scope
ring, similar to the classical Huisgen cycloaddition reaction (38).
with respect to both azides and terminal or internal alkynes (52)
However, in the presence of copper(I) the reaction proceeds
and functional group tolerance. In contrast to CuAAC, RuAAC
faster even under ambient reaction conditions (Scheme 1A).
requires more stringent reaction conditions with respect to
Moreover, in the presence of copper(I) only the 1,4-disubstituted
temperature and solvent (53, 54), which hampers a broad
triazole ring is formed in contrast to the classical Huisgen
cycloaddition reaction in which both 1,4- and 1,5-disubstituted application as a bioconjugation reaction for biologically relevant
triazole regioisomers are formed (39). Since this Cu(I)-catalyzed molecules.
1,3-dipolar cycloaddition reaction is very selective, it is highly Another reaction that is becoming increasingly popular as
compatible with almost all other functional groups present in an attractive click reaction is the Michael addition between thiols
biological macromolecules such as proteins, polysaccharides, and acrylates (55), acrylamides (56), vinyl sulfones (57, 58),
and DNA/RNA. Additionally, the CuAAC can be performed or maleimides (59) to form thioethers (Scheme 2). This Michael
in aqueous media with a high reaction rate and generally leads addition reaction, which is also called thiol-ene coupling (TEC)
to good product yields. Moreover, recent studies showed that or “thio-click” (60, 61), is highly efficient and orthogonal to a
the 1,2,3-triazole moiety, which is formed during the Cu(I)- wide variety of functional groups. The thiol-ene coupling
catalyzed 1,3-dipolar cycloaddition, is similar to the peptide- makes use of the high nucleophilicity of the sulfhydryl moiety
amide bond in terms of geometry. Therefore, the triazole moiety and proceeds under physiological conditions. The formed
has been suggested as a mimic of a peptide amide bond and thioether linkage is very stable under physiological conditions
has been used for example as a dipeptide isostere in β-strands and resists a strong basic or acidic environment and is also stable
and R-helical coiled coils (40-42). toward reducing agents; however, it is susceptible toward
A potential drawback of the Cu(I)-catalyzed cycloaddition oxidizing agents.
reaction for the synthesis of polymers aimed for biomedical and A drawback of the thiol-ene coupling reaction is the
pharmaceutical applications is the cytotoxicity of Cu(I). This sensitivity of the free thiol functionality toward oxidation into
Reviews Bioconjugate Chem., Vol. 20, No. 11, 2009 2003

Scheme 1. Cycloaddition Reactions between Azides and Alkynesa

a
(A) Cu(I)-catalyzed 1,3-dipolar cycloaddition reaction (16-18), (B) ruthenium(II)-catalyzed cycloaddition (51), (C) ring strain promoted
cycloaddition (44), (D) ring strain promoted cycloaddition with 4-dibenzocyclooctynol (45), (E) second generation of ring strain promoted
cycloaddition (46, 47), (F) the tandem [3 + 2] cycloaddition-retro-Diels-Alder ligation method (49).

Scheme 2. Variations of the Thiol-Ene Reactiona co-workers successfully protected the thiol functionality as a
thio acetate (65, 66). Thio esters are highly sensitive toward
mild base and can be cleaved at pH 8.
A detailed discussion on the reaction rate of the thiol-ene
coupling reaction was provided by Hoyle et al. (67). On the
basis of extensive studies by Morgan et al. (68), the influence
of both thiol and ene structure on the thiol-ene free radical
reaction rate is directly related to the electron density on the
ene moiety. Electron-rich enes react more rapidly than electron-
poor enes, e.g., alkene ≈ vinyl ester > allyl ether > acrylate >
N-substituted maleimide > methacrylate (67).

POLYPEPTIDE-BASED POLYMERS
There is great interest in the use of functional peptides as
building blocks for the synthesis of peptide-based polymers,
since these polymers can be applied for the design of drug
delivery systems, scaffolds for tissue engineering and repair,
and as novel biomaterials (1, 69). However, the synthesis of
peptide-based polymers imposes several major synthetic chal-
lenges. Current methods to synthesize such polymers require
a
(G) Michael addition between thiols and acrylates (55), (H) thiols the elaborated use of protection groups and/or unstable preac-
and vinyl sulfones (57, 58), (I) thiols and maleimides (59), (J) thiols tivated building blocks (70-72), or the polymers have to be
and alkenes (R2 and R3 have to be electron withdrawing groups) (62-64). synthesized via protein engineering, which can be very chal-
lenging (73). Recently, new methods have been developed to
a disulfide. This unwanted side reaction can be overcome by synthesize peptide-based polymers by chemoselective ligation,
excluding oxygen from the reaction medium or by adding where unprotected functional peptides are coupled by an
reducing agents, such as TCEP, or by using protected thiol orthogonal conjugation method to form polypeptide products.
derivatives. Such protecting groups must meet several require- In particular, the CuAAC is very suitable, because the formed
ments: they have to be removed quantitatively under mild 1,2,3-triazole moiety is a mimic of a native peptide bond (40, 41).
conditions and the byproduct should be nontoxic. Goessl and In view of the scope of this review, native chemical ligation
2004 Bioconjugate Chem., Vol. 20, No. 11, 2009 van Dijk et al.

Figure 1. Examples of peptidetriazole sequences synthesized by Fan and co-workers (75).

Scheme 3. Synthesis of Ditriazole via “Click-Click” Strategy (76)

Scheme 4. Schematic Representation of the Synthesis of the presence of CuI/ascorbic acid/DIPEA/30% 2,6-lutidine/
Peptide-Based Polymers as Described in Ref 79a DMF, followed by the removal of the Fmoc group with 20%
piperidine/DMF, and the cycle was repeated. After four cycles,
the peptidotriazole was cleaved from the resin to yield the
desired product 2 in high purity. To demonstrate the versatility
of this approach and its compatibility with the common
protecting groups used in peptide chemistry, the synthesis was
repeated, replacing Fmoc-proline azide by side chain-protected
tyrosine, aspartic acid, and lysine building blocks (Figure 1).
Also, in this case the desired product 3 was obtained in high
yield and purity.
Leigh and co-workers (76) developed synthetic methodologies
in which the copper(I)-mediated alkyne-azide cycloaddition
reaction was combined with a silver(I)-catalyzed TMS-alkyne
deprotection (77, 78), and this so-called “click-click” strategy
was used to synthesize step-by-step oligomers of peptide
triazoles (Scheme 3). The authors started with the tripeptide
phenylalanyl-glycyl-glycinate hydrochloride and functionalized
a
Approach A: using CuOAc and high concentration of monomer the C-terminus of this peptide with an alkyne and the N-terminus
led to predominantly linear polymers. Approach B: using CuOAc and with a TMS-protected alkyne (compound 4, Scheme 3). In the
low monomer concentration led to predominantly cyclic oligomers. first reaction, they performed a click reaction with azide-
containing pseudodipeptides, based on the amino acids pheny-
(NCL) for the synthesis of peptide-based polymers is not lalanine 5, leucine, proline, or Nε-Boc-protected lysine. In the
discussed. A recent review on protein synthesis by NCL is given presence of CuSO4/Na-ascorbate, an almost quantitative yield
by Kent (74). of monoclicked product was obtained after 18 h. After the first
In 2005, Fan and co-workers (75) developed a protocol for click reaction, the TMS group was removed by treatment with
the solid-phase synthesis of peptidotriazoles with alternating AgBF4. The addition of the second azido pseudodipeptide 7
triazole and amide linkages in the backbone. The first pepti- resulted in the formation of the ditriazole 8 (Scheme 3).
dotriazole they synthesized consisted of four repeating units of Additional CuSO4/Na-ascorbate was needed to drive the reaction
4-pentynoic acid and Fmoc-proline azide 1 (Figure 1). First, to completion. This one-pot method can be a promising tool
4-pentynoic acid was grafted onto the resin with PyBOP/HOBt for peptide ligation reactions and the synthesis of peptide-based
as coupling reagents. Next, Fmoc-proline azide was clicked in polymers.
Reviews Bioconjugate Chem., Vol. 20, No. 11, 2009 2005

Scheme 5. Synthesis of Diblock Copolymers Containing Poly(γ-benzyl-L-glutamate) and Poly(Nε-trifluoroacetyl-L-lysine) (82)

Scheme 6. Synthesis of Alkyne Functionalized Polyesters (83)

Scheme 7. Synthesis of Glycopolymers via Click Reactionsa

a
(A) Reagents and conditions: (a) N-(ethyl)-2-pyridylmethanimide/CuBr, toluene, 70 °C; (b) TBAF, acetic acid, THF, -20 to +25 °C; (c) RN3,
[CuBr(Ph3P)2], DIPEA. (B) (d) 4-Pentenoic anhydride, DMP, pyridine, DMF; (e) UV, glucothiose, DMPA, DMF.

Rather than a step-by-step method, Liskamp and co-workers.


explored the one-step polymerization of azido-phenylalanyl-
alanyl-propargyl amide as a model peptide under different
reaction conditions (Scheme 4) (79), aiming to develop an
efficient method for the synthesis of mimics of natural biopoly-
mers with repetitive peptide sequences. It was demonstrated that
the degree of polymerization was strongly dependent on the
polymerization conditions (e.g., type of catalyst, monomer
concentration, and reaction temperature). When the monomer
was polymerized at room temperature with CuSO4/Na-ascorbate
as the catalyst, only relatively low molecular weight (Mn: 6.9
kDa) and mostly linear polypeptides were formed. When the Figure 2. Structure of trehalose polymers synthesized by Reineke and
same monomer was polymerized in the presence of CuOAc as co-workers (89, 90).
2006 Bioconjugate Chem., Vol. 20, No. 11, 2009 van Dijk et al.

Scheme 8. Synthesis of Degradable-Brushed hydrolysis of the ester bond. In a similar approach, Guan and
pHEMA-pDMAEMA (102) co-workers (81) recently reported on the synthesis of peptide-
based polymers that can fold into well-defined β-sheets followed
by a self-assembly into hierarchical nanostructures.
Taton and co-workers (82) developed a methodology for the
synthesis of diblock copolypeptides by combining ring-opening
polymerization (ROP) of N-carboxyanhydrides (NCA) combined
with CuAAC-catalyzed click-chemistry. In their approach, azide-
and alkyne-terminated poly(γ-benzyl-L-glutamate) (9 and 12,
Scheme 5) and poly(Nε-trifluoroacetyl-L-lysine) (10 and 13,
Scheme 5) were synthesized by ROP of the corresponding NCA
with an azide- or alkyne-containing initiator. The azide and
alkyne polymers were subsequently conjugated in DMF at 50
°C with CuBr complexed by N,N,N′,N′′,N′′-pentamethyldieth-
catalyst using microwave irradiation at 100 °C, long linear ylenetriamine (PMDETA) as catalyst. After 36 h, the block
dipeptide-based polymers were formed with an Mn up to 45 copolymers (11 and 14) were obtained in almost quantitative
kDa (300 amino acid residues) as shown in Scheme 4 (approach yield (Scheme 5).
A). Interestingly, when the monomer was polymerized at low Emrick and co-workers (83) developed a method for the
monomer concentration, using CuOAc as catalyst and under synthesis of biocompatible amphiphilic graft polyesters. In their
microwave irradiation at 100 °C, cyclic oligomeric structures approach, aliphatic polyesters with pendant acetylene groups
were formed (approach B, Scheme 4). In a followup study, these were synthesized by ROP of R-propargyl-δ-valerolactone 15
optimized polymerization conditions were used to synthesize with ε-caprolactone 16 (Scheme 6). These acetylene-function-
biodegradable peptide-based polymers with functional groups, alized polymers 17 were subsequently grafted with PEG and
such as the ε-NH2 group of lysine (80). The polymers based on oligopeptides, using CuAAC and azide-terminated PEG or
the monomers azido-phenylalanyl-alanyl-lysyl-propargyl amide oligopeptides in the presence of CuSO4/Na-ascorbate for 16 h
or azido-phenylalanyl-alanyl-glycoloyl-lysyl-propargyl amide at 80 °C. The amphiphilic graft polyesters were shown to be
were designed to contain recognition sites for proteases like biocompatible by in Vitro cytotoxicity evaluation, suggesting
trypsin and chymotrypsin. Moreover, the latter polymer also their suitability for a range of biomaterial applications.
contains ester bonds that can be cleaved by chemical hydrolysis.
It was shown that the number average molecular weight (Mn)
GLYCOPOLYMERS
of the polymers could be tailored between 4500 and 14 000 Da Carbohydrates are involved in a number of important biologi-
(33 to 100 amino acid residues) depending on the monomer cal processes such as cell-cell recognition and cell-protein
concentration used during the polymerization. As anticipated, interactions and play a role in the targeting of hormones.
the synthesized polymers were cleaved by both proteases Moreover, antibodies and toxins make use of carbohydrates for
yielding low molecular weight degradation products. The selective recognition of their antigens and target receptors (84-86).
polymer with the ester bond (polymers based on azido- However, individual protein-carbohydrate interactions are
phenylalanyl-alanyl-glycoloyl-lysyl-propargyl) was also de- generally weak; therefore, protein-binding carbohydrates exist
graded under physiological conditions (t1/2 ) 5 h) due to in higher-order oligomeric structures, presenting multiple bind-

Figure 3. Structures of alkyne-bearing dendrimers used by Pieters, Liskamp, and co-workers (106, 107).
Reviews Bioconjugate Chem., Vol. 20, No. 11, 2009 2007

Scheme 9. Synthesis of Dendrimers via Thiol-Ene Reaction (109)a

a
Reaction conditions: (a) 2,2-dimethoxy-2-phenylacetophenone, hυ (30 min), solvent free; (b) DMAP, pyridine.

ing sites, known as the “cluster glycoside effect” (87, 88). These via CuAAC in the presence of [CuBr(Ph3P)2] and DIPEA
multiple binding sites help to improve the weak individual (Scheme 7A).
interactions to achieve strong and selective interactions. Sci- Alternatively, Stenzel and co-workers (93) used the thiol-ene
entists have attempted to mimic this “cluster glycoside effect” coupling reaction for the postsynthesis modification of glyco-
by synthesizing various well-defined synthetic glycopolymers. polymers. First, a block copolymer containing di(ethylene
Depending on the method applied, such glycopolymers can glycol) methyl ether methacrylate (DEGMA) and 2-hydroxy-
be divided into two different classes. They can be synthesized ethyl methacrylate (HEMA) was synthesized via RAFT polym-
either by polymerization of carbohydrate containing monomers erization. Then, the blockcopolymer and a commercially
(vide infra, Reineke et al. 89, 90) or by postfunctionalization available HEMA polymer were grafted with glucothiose in the
of polymers with carbohydrates. Especially for the latter presence of the photoinitiator 2,2-dimethoxy-2-phenyl-acetophe-
approach, click chemistry turned out to be a valuable tool. none (DMPA) under UV irradiation for 2 h (shown for
Several examples to obtain glycopolymers employing both homopolymer 24 in Scheme 7B). The resulting block copolymer
CuAAC and the thiol-ene coupling reaction have been men- was used to form thermo-responsive micelles that can be used
tioned in the literature. Haddleton and co-workers have syn- as a potential drug carrier.
thesized various glycopolymers from alkyne backbone-func- In a recent article by Diehl and Schlaad (94), the thiol-ene
tionalized polymers and azide-functionalized carbohydrates via coupling reaction has been used to functionalize poly[2-
the CuAAC as shown in Scheme 7A (91, 92), such as various (isopropyl/3-butenyl)-2-oxazoline] copolymers with various
mannose- and galactose-containing multidentate ligands for thiols (e.g., octanethiol, 3-mercaptopropionic acid, and 2-mer-
studying lectin binding (91). In this approach, the alkyne- captoethanol). Via this approach, these authors synthesized a
functionalized polymers were synthesized by transition-metal- series of thermosensitive glycopolymers with cloud points
mediated living radical polymerization (TMM-LRP) of meth- tunable over a wide temperature range.
acrylate 19. After polymerization, the TMS group was removed In a recent article by Haddleton and co-workers (95), the
with TBAF and acetic acid, followed by the coupling of various synthesis of alkyne-functionalized polymers via chain transfer
protected and unprotected azide-functionalized carbohydrates polymerization (CCTP) is described. The polymers obtained by
2008 Bioconjugate Chem., Vol. 20, No. 11, 2009 van Dijk et al.

alized trehalose monomers and dialkyne-amide comonomers in


the presence of CuSO4/Na-ascorbate. The polymers contained
a trehalose unit for increased biocompatibility, an oligoamine
unit for electrostatic interactions with DNA, and a triazole
functionality for hydrophobic, van der Waals, and hydrogen-
bonding interactions with nucleic acids (Figure 2).
All three newly synthesized polymers were effective as
nucleic acid carriers in the absence as well as in the presence
of serum. The polymer with the highest secondary amine density
gave polyplexes with low toxicity and high cellular delivery.
The transfection efficiency of the trehalose polymers was an
order of magnitude higher than Jet-PEI, one of the most efficient
in Vitro gene delivery polymers, in serum-free conditions. In a
second study (90), the same group investigated the influence
of the molecular weight of the trehalose click polymer on
polyplex stability and pDNA cellular delivery efficiency. In this
study, it was shown that a higher degree of polymerization
resulted in a higher polyplex stability, although no effect was
observed in pDNA binding affinity, cellular uptake, and DNase
protection in relation to the Mw.
Ideally, suitable polymeric transfectants should be nontoxic,
nonimmunogenic, and preferably biodegradable in a controlled
Figure 4. Schematic representation of the approach to immobilize manner. Furthermore, biodegradable polymers should yield
proteins on a solid support (110). (A) activation of the surface; (B)
attachment of dendrimers (aminocaproic acid linker, dendrimers); (C) degradation products with a molecular weight lower than 30
functionalization with thiols; (D) drop casting of terminal-olefin- kDa, because these degradation products can be excreted by
functionalized proteins and immediate coverage with a photomask; (E) the kidneys (100, 101). To reduce the cytotoxicity of cationic
removal of the mask. (Copyright Wiley-VCH Verlag GmbH & Co. polymers, Hennink and co-workers (102) designed a high-
KGaA. Reproduced with permission (110).) molecular-weight polymer composed of a low-molecular-weight
cationic poly(2-dimetylamino)ethyl methacrylate) (pDMAEMA)
this approach can be end-functionalized via the thiol-ene that was grafted onto a polymer backbone of an uncharged
coupling reaction as well as the CuAAC reaction. In this article, hydrophilic polymer, poly(hydroxyethyl methacrylate) (pHE-
several azide-functionalized carbohydrates were grafted via the MA), via biodegradable linkages. Both pDMAEMA and pHE-
CuAAC reaction onto the alkyne side chain of the polymers. MA were synthesized by atom transfer radical polymerization
The grafted polymers on their turn were end-functionalized using (ATRP). For this goal, pDMAEMA was end-functionalized with
the thiol-ene coupling reaction with different thiols. The an azide (27), while pHEMA was randomly functionalized with
synthesized mannose-and galactose-containing glycopolymers acetylene moieties (28, Scheme 8). The polymers were “clicked”
were able to function as multivalent ligands for the recognition together via the CuAAC in DMF at 50 °C with CuBr as catalyst
of lectins. The CCTP in combination with the thiol-ene (Scheme 8). The molecular weight of the polymer as well as
coupling reaction and the CuAAC reaction is a powerful tool the number of grafts could easily be varied. Upon incubation
for the convenient synthesis of many different types of functional at physiological conditions (pH 7.4, 37 °C), the carbonate ester
polymers and conjugates. bonds were readily hydrolyzed (t1/2: 96 h). The molecular weight
of the final main degradation product was very close to that of
POLYMERS FOR NONVIRAL GENE DELIVERY the starting pDMAEMA, indicating that the carbonate esters
were quantitatively hydrolyzed. Furthermore, the synthesized
Gene therapy holds a great promise for the treatment of polymers were able to condense DNA into small particles, which
diseases with a genetic origin that are currently incurable. The were able to transfect cells efficiently in the presence of
success of gene therapy largely depends on the availability of endosome-disruptive INF/7 peptide. Finally, the polymers had
suitable delivery vehicles. Although viral vectors display rather a lower toxicity compared to high molecular weight pD-
good transfection properties, both in Vitro and in ViVo, there MAEMA, making this an effective approach to reduce the
are a large number of problems associated with the use of these toxicity of high-molecular-weight cationic polymers (102).
vectors (96, 97). Therefore, there is an increasing interest in
the development of so-called nonviral gene delivery systems
DENDRIMERS
(98). Two major classes of nonviral systems can be distin-
guished, namely, those based on cationic lipids or those based Dendrimers are highly branched well-defined polymers and
on cationic polymers. However, generally speaking, the cationic among others used for the simultaneous presentation of biologi-
polymers developed to date often display cytotoxicity effects cally relevant but individually weakly interacting ligands (103).
likely due to interactions with proteins and phospholipid Multivalent ligands often bind much more strongly to the
membranes. Therefore, new polymers with a significantly interacting protein than their monovalent counterparts (103-105).
reduced cytotoxicity have to be developed. The click reaction Pieters and co-workers (106) developed a versatile microwave-
can be a valuable tool to synthesize these new polymers for assisted CuAAC approach that allowed the conjugation of azido
nonviral drug delivery systems. carbohydrates with different kinds of alkyne-bearing dendrimers
Reineke and co-workers (89) exploited the CuAAC to (Figure 3). With this procedure, they were able to synthesize
synthesize a family of trehalose-based glycopolymers, e.g., 26 triazole glycodendrimers in high yields, up to the nonavalent
(Figure 2). They were inspired by the work of Dervan et al., level. These glycodendrimers can be used to increase affinities
who showed that macromolecules that contain various hetero- in various applications such as the binding with bacteria,
cyclic residues, such as derivatives of pyrrole and imidazole, bacterial toxins, and lectins.
are able to bind nucleic acids (99). Three polymers with different In a similar approach, Liskamp and co-workers (107) used
amine stoichiometry were synthesized from diazide function- the microwave-assisted 1,3-dipolar cycloaddition reaction to
Reviews Bioconjugate Chem., Vol. 20, No. 11, 2009 2009

Scheme 10. Schematic Representation of the Approach Described by Caruso and Co-Workers for the Synthesis of Ultrathin
pH-Responsive Polymer Films (115, 116)

Scheme 11. Synthesis and Hydrolysis of Dextran-Based biomolecules onto solid surfaces in a patterned way. In detail,
Multilayers (118) they first attached polyamidoamine dendrimers covalently to
silicon oxide surfaces (Figure 4B). The dendrimers were
extended with an aminocaproic acid spacer, and cystamine was
coupled to this spacer and subsequently reduced to obtain a free
thiol functionality (Figure 4C). Next, the silicon oxide wafers
were incubated with terminal-olefin-functionalized proteins
dissolved in ethylene glycol and covered with a photomask
(Figure 4D). Subsequent irradiation led to patterning with protein
adducts covalently attached via a thioether moiety. The amount
of protein that was immobilized could be tailored by varying
the irradiation time. With this approach, various biomolecules,
like biotin, calf-intestine-alkaline phosphatase, the small GTPase
Ras, and a phosphopeptide, were photochemically attached to
the dendrimer-coated wafers. All biomolecules were successfully
attached and retained their structure and activity. The authors
anticipate that this method can be used for the controlled
fabrication of structured assemblies of proteins on artificial
surfaces.

PREPARATION AND MODIFICATION OF MICRO-


AND NANOPARTICLES
Nanoparticles have great potential in the biomedical and
synthesize peptide-presenting multivalent dendrimers. The azido- pharmaceutical field (111). Nanoparticles can, for example, serve
peptides were mixed with the acetylene-functionalized den- as delivery vehicles for drugs (112, 113). However, current
drimers (e.g., 31, Figure 3) in the presence of CuSO4/Na- methods to encapsulate drugs in polymeric nanoparticles require
ascorbate in aqueous DMF and heated at 100 °C in a microwave the use of organic solvents or harsh conditions, which may result
reactor. The products were obtained in nearly quantitative yield in significant loss of activity, especially if complex biomolecules
in a relatively short reaction time (5-10 min). With this method, are used as the active compound. Consequently, novel and mild
di-, tetra-, octa-, and hexadecavalent dendrimeric peptides were approaches to obtain nanoparticles are highly desirable.
successfully synthesized. In a recent study, Liskamp and co- With the aim to prepare functionalized polymersomes from
workers (108) used the same strategy as described above to amphiphilic polystyrene-block-poly(acrylic acid), van Hest and
synthesize 111In-labeled DOTA-conjugated dendrimers with co-workers (114) used ATRP to synthesize the block copolymers
multivalent cyclic RGD peptides. The tetravalent 111In-labeled followed by the substitution of the bromine functional end group
RGD dendrimers had increased affinity toward the Rvβ3 integrin of the hydrophilic block by an azide. Upon slow addition of
receptor and had better tumor targeting properties than their water to a solution of block copolymers in dioxane, the
monovalent congeners. amphiphilic block copolymers self-assemble into polymersomes
Recently, Hawker and co-workers (109) developed a facile with surface-exposed azide functionalities. These azide-func-
and efficient method for the synthesis and end-functional- tionalized vesicles were used as scaffold for further conjugation
ization of poly(thioether) dendrimers using the thio-ene with alkyne-functionalized dansyl, biotin, and enhanced green
coupling reaction. The synthesis of the dendrimers was started fluorescent protein using CuSO4/Na-ascorbate in the presence
from the tris-alkene core 2,4,6-triallyloxy-1,3,5-triazine (33, of the ligand tris-(benzyltriazolylmethyl)amine (TBTA).
Scheme 9). To this core, 1-thioglycerol (34) in the presence Caruso and co-workers developed a general approach for the
of catalytic amounts of photoinitiator 2,2-dimethoxy-2- layer-by-layer assembly of ultrathin polymer films on planar
phenylacetophenone was coupled by means of UV irradiation substrates (115) and particles (116) for the preparation of pH-
for 30 min at room temperature to yield the first-generation responsive nanocapsules. In their approach, poly(acrylic acid)
dendrimer 35. Subsequent esterification of the hexa-hydroxy with either alkyne or azide functionalities were synthesized by
dendrimers with 4-pentenoic anhydride 36 yielded the ene- living radical polymerization. Subsequently, the azide and alkyne
functionalized dendrimers 37. Repeating the final two reaction functionalized poly(acrylic acid)s were alternately assembled
steps resulted in the synthesis of dendrimers functionalized on alkyne-functionalized silica particles, with CuSO4/Na-
with 48 alkene functionalities. End-functionalization was also ascorbate as catalyst (Scheme 10). The silica template was
performed with the thiol-ene coupling reaction with suitable removed by treatment with NH4F-buffered HF at pH 5. The
thiol-containing reagents, e.g., thioglycolic acid 38, 4-(pyren- “click capsules” showed pH reversible responsive behavior in
1-yl)butyl 2-mercaptoacetate 39, and Fmoc-Cys-OH (40) as a reversible manner, i.e., the capsule diameter varied between
shown in Scheme 9. about 5 µm in acidic and 8 µm in basic conditions. This pH-
Waldmann and co-workers (110) used dendrimers and the responsiveness could be exploited to load and concentrate drugs
thiol-ene coupling reaction to immobilize proteins and other inside the capsules. Moreover, the presence of azide or alkyne
2010 Bioconjugate Chem., Vol. 20, No. 11, 2009 van Dijk et al.

Scheme 12. The Two Different Approaches to Synthesize PVA Hydrogels (123)

Scheme 13. Formation of Hyaluronic Acid Based Click Gels (Scheme 11). In a subsequent article (118), the potential of the
(125) biodegradable dextran multilayers for drug delivery was exam-
ined. The microcapsules were loaded with fluorescein isothio-
cyanate-labeled dextran as a model drug compound and
incubated at physiological conditions. It was shown that the drug
release rate could be tailored by the cross-linking density of
the dextran multilayers.
In a recent paper by Caruso and co-workers (119), the
development of a thiol-ene version of the layer-by-layer
assembly of polymer films on silica particles is described. In
this approach, poly(methacrylic acid), containing either thiol
groups or ene functionalities, was alternatively deposited with
poly(vinylpyrrolidone) on silica particles under UV irradiation.
Schlaad and co-workers (62) used a thiol-ene coupling
reaction for the synthesis of self-assembling peptide hybrid
amphiphiles. They grafted cysteine and cysteine-containing
dipeptides onto the hydrophobic block of polybutadiene-block-
poly(ethylene oxide) polymer. The grafting of cysteine had little
effect on micelle formation. However, when the dipeptide
cysteinyl-phenylalanine was coupled to the amphiphilic polymer,
vesicles and worm-like micelles were formed. More recently,
moieties on the outer shell of the capsules allows for the Schlaad and co-workers (63, 64) used the same method to
chemoselective postfunctionalization of the nanoparticles. synthesize polybutadienes modified with hydrophilic function-
In a similar approach, De Geest and co-workers (117) alities. The thiol-ene coupling reaction was used to incorporate
synthesized dextran-based multilayer films and hollow capsules. thiol-containing molecules, in which the carboxylic acid or
By introducing carbonate ester bonds in the azide and alkyne amine functional groups and amino acid residues were unpro-
linkages, the multilayers can be degraded by chemical hydrolysis tected. With the incorporation of such titratable groups, the
polymers were able to self-assemble into pH-responsive unila-
mellar or multilamellar vesicles.
Riguera and co-workers (120) reported an efficient method
to synthesize polyion complex (PIC) micelles with the CuAAC
catalyzed click reaction that are stable under physiological
conditions. In their approach, azide-functionalized gallic acid-
triethylene glycoside (GATG) dendrimers and PEG-GATG
block copolymers were coupled to different alkynes function-
alized with sulfate, sulfonate, or carboxylate groups in the
Figure 5. Structures of doxorubicin 55 and benzidamine 56. presence of CuSO4/Na-ascorbate. When the sulfated PEG-
Reviews Bioconjugate Chem., Vol. 20, No. 11, 2009 2011

Figure 6. Structure of tetraazide-functionalized four-armed star-shaped PEG 57 and diacetylene-functionalized peptide 58 (126).

solubility of the azide- or alkyne-functionalized PVA, only low


degrees (1-5%) of modification were possible. In a second
approach to synthesize PVA-based hydrogels, acetylene-func-
tionalized polyvinyl alcohol 49 was cross-linked with telechelic
bifunctional poly(ethylene glycol) diazide 50 (Scheme 12). This
approach was superior in gel formation and gave higher values
of gel fraction.
Hawker and co-workers used PEG as the main structural
component for model “click” hydrogels. In their approach,
diacetylene-functionalized telechelic PEG derivatives and tet-
raazide-functionalized PEG were coupled by “click chemistry”
at room temperature using 2 equiv of PEG diacetylene, 0.4 equiv
CuSO4, and 1 equiv of Na-ascorbate. Under these conditions,
the hydrogels were formed within 30 min, and fluorescence
analysis revealed the presence of maximal 0.2% unreacted azide/
acetylene moieties. By varying the length of the diacetylene
PEG chain, they showed that the properties (swelling degree
and max true stress) of the hydrogels could be tailored (124).
Lamanna and co-workers (125) developed a procedure for
the formation of hydrogels based on hyaluronan derivatives
suitable for tissue engineering applications. They synthesized
Figure 7. Two different approaches to synthesize photocleavable hyaluronic acid bearing either azide or alkyne groups (52 and
hydrogels (reproduced with permission from ref 128).
53, Scheme 13). Hydrogels were formed within a few minutes
by dissolving both components in H2O in the presence of 1%
GATG block copolymer was mixed in a stoichiometric ratio w/v of CuCl (Scheme 13). Residual copper entrapped in the
with poly(L-lysine), spherical micelles with a narrow size hydrogel was removed by dialyzing the hydrogel against an
distribution were obtained. aqueous buffer containing EDTA as metal chelator. The
In a recent article authored by Pine and co-workers (121), a synthesized hydrogels were evaluated for drug-release capabili-
new generic method for the covalent linkage of a wide variety ties with doxorubicin 55 and benzidamine 56 as model drugs
of molecules to the surface of colloidal polymer microspheres (Figure 5). Benzidamine was quantitatively released within
via the CuAAC reaction has been described. In their approach, hours, whereas the release of doxorubicin was prolonged over
polystyrene particles containing 4-vinylbenzyl chloride moieties several weeks. The release rate could be tailored by varying
were converted into the corresponding azides and subsequently the degree of cross-linking. The slower release of doxorubicin
reacted with various alkynes to obtain triazole-functionalized was explained by strong electrostatic interactions between the
microparticles. As a typical example, the authors functionalized protonated amino group of doxorubicin and carboxylate groups
polystyrene microspheres with two polyethylene oxide-based
of the hydrogels. The authors also argued that π-π stacking
polymers. However, it has been stated that this approach is
interactions between the triazole ring and the aromatic moiety
sufficiently general since it can readily be adapted to colloids
likely contributed to the slower release of doxorubicin.
that consist of other (polymeric) materials.
Lamanna and co-workers also examined the suitability of their
hyaluronan hydrogels as scaffolds for tissue engineering.
HYDROGELS Therefore, a hydrogel was formed in a YPD (yeast peptone
Hydrogels are three-dimensional, hydrophilic polymeric dextrose) cell suspension and the gels were stored for 2 days at
networks capable of absorbing large amounts of water (122). 28 °C. Subsequently, the gels were mechanically broken and
Hydrogels have received great interest over the past decades, the number of cells was counted, and 24 h later, 80% of the
since they can be used for a wide range of applications including cells exhibited proliferating activity.
drug delivery systems and scaffolds for tissue engineering and Anseth and co-workers (126) developed a procedure to
repair. In 2006, Hilborn and co-workers were the first to apply integrate multifunctional photoreactive polypeptides into a
click chemistry to synthesize hydrogels (123). They synthesized hydrogel network. In their approach, they combined both the
poly(vinyl alcohols) (PVA) functionalized with either acetylene CuAAC and thiol-ene coupling reaction. First, they synthesized
or azide groups (46 and 47, Scheme 12). Upon addition of hydrogels by clicking tetraazide-functionalized four-arm star-
CuSO4/Na-ascorbate, a hydrogel was formed within a few shaped PEG 57 with diacetylene-functionalized peptide 58
minutes (48, Scheme 12). However, in order to retain the water (Figure 6). The hydrogels were formed within a few minutes
2012 Bioconjugate Chem., Vol. 20, No. 11, 2009 van Dijk et al.

Scheme 14. Dextran Hydrogel Formation by Thiol-Ene Coupling (135, 136)

in the presence of 0.25 equiv CuSO4 (based on azide functional- cycloaddition, thereby abolishing the need of a (cytotoxic)
ity) and Na-ascorbate. Next, a cysteine containing fluorescently copper catalyst.
labeled peptide was incorporated by thiol-ene photocoupling Zhou and co-workers (130) developed a strategy for the in
to the Alloc protecting group present in the hydrogel. situ gelation of poly(N-isopropylacrylamide-co-hydroxyethyl
Dubois and co-workers (127) synthesized adaptative and methacrylate) (p(NIPAAm-co-HEMA))-based polymers by the
amphiphilic polymer conetworks based on hydrophilic poly(N,N- CuAAC. Two p(NIPAAm-co-HEMA)-based polymers, with
dimethylamino-2-ethyl methacrylate) (pDMAEMA) and hydro- either pendant azide or alkyne groups, were synthesized.
phobic poly(ε-caprolactone) (pCL) by a combination of ATRP, Hydrogels were formed by dissolving the two polymers and
ROP, and CuAAC. In their approach, the azido-containing 2-(2- incubating them with CuSO4/Na-ascorbate as catalyst and
azidoethoxy)ethyl methacrylate (AEEMA) was copolymerized PMDETA as ligand for 24 h at room temperature. The obtained
with DMAEMA by ATRP. The p(DMAEMA-co-AEEMA) CuAAC hydrogels had faster shrinking/swelling kinetics com-
copolymer was subsequently cross-linked with diacetylene- pared with traditionally synthesized p(NIPAAm) hydrogels.
functionalized pCL, in the presence of CuBr complexed with Yang et al. (131) synthesized PEG-RGD peptide hydrogels
2,2′-bipyridine as a ligand at room temperature for 24 h to obtain for cell entrapment using the CuAAC. PEG is an ideal polymer
a highly porous amphiphilic co-network. for hydrogel formation, because the formed gels generally have
Turro and co-workers (128) utilized ATRP and CuAAC for a high water absorbing capacity and are resistant to protein
the synthesis of photodegradable polymeric model networks. adsorption. However, the viability of encapsulated cells in PEG
For the synthesis of the networks, the authors employed two hydrogels is generally low due to the nonadherent properties
different approaches (Figure 7). In their first approach, linear for the cells. By virtue of introducing RGD peptides into the
azido-telechelic macromonomers (MAC) possessing a photo- PEG hydrogels, cell adhesion properties were increased, result-
cleavable group (nitrobenzyloxycarbonyl) at the center were ing in higher cell viability. For the PEG precursor of the
synthesized by ATRP. Subsequently, both end groups of the hydrogel, the authors prepared tetra-acetylene PEG by esteri-
linear polymers were reacted with NaN3. The obtained azide- fication of tetra-hydroxy-terminated four-armed PEG with
functionalized macromonomers were cross-linked with a small 4-pentynoic acid. As a cross-linking agent, they synthesized a
tetrafunctional alkyne in the presence of CuBr as catalyst and series of diazide-functionalized RGD peptides by reacting RGD
N,N,N′,N′′,N′′-pentamethyldiethylenetriamine (PMDETA) as peptides with 6-azidohexanoic acid. Hydrogels were prepared
ligand. In a second approach, the authors started from a by combining the acetylene-functionalized PEG and diazide-
tetrafunctional ATRP initiator containing four photocleavable functionalized RGD peptides in the presence of CuSO4/Na-
groups to yield four-armed star-shaped poly(tert-butyl acrylate) ascorbate at ambient temperatures. The gelation time could be
polymers. Next, the four end groups were converted into azide tailored between 2 and 30 min by variation of the catalyst or
functionalities, and the star-shaped polymers were cross-linked precursor concentration and by variation of the temperature. To
with a linear bifunctional alkyne. Both model networks were investigate if the hydrogels could be used as cell delivery
degraded upon exposure to UV light of 350 nm to yield soluble vehicle, the hydrogels were seeded with primary human dermal
well-defined polymer fragments. In a second paper, Turro and fibroblasts cells. The hydrogels without or at low RGD peptide
co-workers (129) optimized this procedure by substituting the loading were unable to adhere cells after 18 h. When the RGD
CuAAC cross-linking with a strain-promoted azide-alkyne peptide loading in the precursor solution was increased, the
Reviews Bioconjugate Chem., Vol. 20, No. 11, 2009 2013

attachment and proliferation of cells was also enhanced. These ceutical polymers synthesized by the CuAAC, the thiol-ene
data hold promise to use these types of hydrogels as carriers reaction, or a combination of both reactions were highlighted.
for cell entrapment and tissue engineering. In recent years, the CuAAC and thiol-ene reactions have been
In 2003, Stein and co-workers (132) reported a method, based used to synthesize polymers with different architectures. e.g.,
on the thiol-ene coupling, to synthesize PEG-based hydrogels block and graft polymers, dendrimers, and hydrogels, for
in situ. They used PEG-based copolymers containing multiple pharmaceutical and biomedical polymers. Although the CuAAC
thiol groups synthesized from diamino-PEG and 2-mercapto- and thiol-ene reactions have many advantages, they also have
succinic acid. As cross-linking agent, PEG functionalized with some drawbacks. The CuAAC reaction requires the need of a
vinylsulfone groups was used. Upon combining both PEG cytotoxic copper catalyst, which is sometimes difficult to
polymers, hydrogels with a water content higher than 90% were remove, while the free thiol needed for the thiol-ene reaction
rapidly formed under mild conditions (pH 6.0-8.0 at room is susceptible to oxidation. The development of new variations
temperature). Furthermore, it has been shown that increasing of these “click reactions”, such as the strain-promoted click
the pH gave rapid gel formation, which is in line with the reaction, can address the problem of cytotoxic copper catalyst.
increasing nucleophilicity of the thiol functionality at higher Nevertheless, both the CuAAC and thiol-ene reactions have
pH values. The authors have demonstrated that the reaction acquired central positions in the synthesis of polymers for
between thiol and vinylsulfone is compatible with other biomedical and pharmaceutical applications, and it is expected
functional groups present in, e.g., proteins, by carrying out a that their importance for the design and synthesis of such
release study of fluorescein-labeled BSA, which showed quan- polymers will increase rapidly in the coming years.
titative release after 25 days.
Hubbell and co-workers (133) examined the formation of
hydrolytically degradable PEG hydrogels via thiol-ene cou- LITERATURE CITED
pling. They used a four-armed and eight-armed PEG spacer (1) Deming, T. J. (1997) Polypeptide materials: New synthetic
functionalized with acrylate end groups and either dithiothreitol methods and applications. AdV. Mater. 9, 299–311.
or PEG-dithiol as cross-linking agents. The hydrogels were (2) Kolb, H. C., Finn, M. G., and Sharpless, K. B. (2001) Click
prepared by mixing the acrylate-functionalized star-shaped PEG chemistry: diverse chemical function from a few good reactions.
and dithiol in 1:1 stoichiometric ratio in PBS (pH 7.8) and Angew. Chem., Int. Ed. 40, 2004–21.
incubated overnight at 37 °C to ensure complete conversion. (3) Gacal, B., Durmaz, H., Tasdeln, M. A., Hizal, G., Tunca, U.,
The PEG-acrylate contains an ester bond that can be cleaved Yagci, Y., and Demirel, A. L. (2006) Anthracene-maleimide-
by chemical hydrolysis. The presence of the thioether group in based Diels-Alder “click chemistry” as a novel route to graft
the proximity of the PEG-acrylate ester greatly increases the copolymers. Macromolecules 39, 5330–6.
sensitivity of the ester toward hydrolysis by several orders of (4) Posner, T. (1905) Beiträge zur Kenntniss der ungesättigten
magnitude (134). By altering the molecular weight of the PEG Verbindungen. II. Ueber die Addition von Mercaptanen an
derivative, or by varying the concentration of either the cross- ungesättigte Kohlenwasserstoffe. Chem. Ber. 38, 646–57.
linker or the PEG-acrylate molecules, resulted in hydrogels with (5) Saxon, E., Armstrong, J. I., and Bertozzi, C. R. (2000) A
swelling ratios. The different swelling ratios also resulted in “traceless” Staudinger ligation for the chemoselective synthesis
different degradation kinetics. of amide bonds. Org. Lett. 2, 2141–3.
Feijen and co-workers (135, 136) synthesized biodegradable (6) Nilsson, B. L., Kiessling, L. L., and Raines, R. T. (2000)
dextran hydrogels by thiol-ene coupling. The dextran hydrogels Staudinger ligation: A peptide from a thioester and azide. Org.
were formed in situ by mixing solutions of vinylsulfone dextran Lett. 2, 1939–41.
59 (Scheme 14) and tetramercapto four-armed star poly(ethylene (7) Saxon, E., and Bertozzi, C. R. (2000) Cell surface engineering
glycol) 60. The degradation time of hydrogels 61 could be tailored by a modified Staudinger reaction. Science 287, 2007–10.
by varying the length of the spacer of the dextran polymers and (8) Dawson, P. E., Muir, T. W., Clark-Lewis, I., and Kent, S. B. H.
the degree of vinylsulfone substitution. To study the versatility of (1994) Synthesis of proteins by native chemical ligation. Science
these gels as a drug delivery system, four model proteins (immu- 266, 776–9.
noglobulin, bovine serum albumin, lysozyme, and fibroblast growth (9) Ruttekolk, I. R., Duchardt, F., Fischer, R., Wiesmuller, K. H.,
factor) were entrapped inside the gel. The release of the proteins Rademann, J., and Brock, R. (2008) HPMA as a scaffold for
did not show a burst effect and depended on the degree of the modular assembly of functional peptide polymers by native
chemical ligation. Bioconjugate Chem. 19, 2081–7.
vinylsulfone substitution and dextran molecular weight.
(10) Shangguan, N., Katukojvala, S., Greenburg, R., and Williams,
Hoffman and co-workers (137) used the thiol-ene coupling
L. J. (2003) The reaction of thio acids with azides: A new
method to synthesize heparin-based hydrogels. In their approach,
mechanism and new synthetic applications. J. Am. Chem. Soc.
heparin was first reacted with EDC/HOBt and an excess of 125, 7754–5.
cysteamine to obtain thiol-functionalized heparin. PEG-diacrylate (11) Merkx, R., Brouwer, A. J., Rijkers, D. T. S., and Liskamp,
was used as the cross-linking agent. The mechanical properties R. M. J. (2005) Highly efficient coupling of beta-substituted
and gelation kinetics of the hydrogels could be tailored by the aminoethane sulfonyl azides with thio acids, toward a new
degree of thiol group substitution of heparin. The heparin-based chemical ligation reaction. Org. Lett. 7, 1125–8.
hydrogels were degraded at physiological pH by chemical hy- (12) Kolakowski, R. V., Shangguan, N., Sauers, R. R., and
drolysis of the ester bond present in the PEG diacrylate cross-linker. Williams, L. J. (2006) Mechanism of thio acid/azide amidation.
To prove that the hydrogels are compatible with cells, the hydrogels J. Am. Chem. Soc. 128, 5695–702.
were gelated in the presence of fibroblasts. After gelation, most of (13) Barlett, K. N., Kolakowski, R. V., Katukojvala, S., and
the cells (95%) retained their viability. Williams, L. J. (2006) Thio acid/azide amidation: An improved
route to N-acyl sulfonamides. Org. Lett. 8, 823–6.
CONCLUSIONS (14) Merkx, R., van Haren, M. J., Rijkers, D. T. S., and Liskamp,
R. M. (2007) Resin-bound sulfonyl azides: efficient loading and
The CuAAC reaction as well as thiol-ene coupling reaction activation strategy for the preparation of the N-acyl sulfonamide
as discussed in this review allow the synthesis of well-defined linker. J. Org. Chem. 72, 4574–7.
polymers with tailored properties and provide very attractive (15) Zhang, X., Li, F., Lu, X. W., and Liu, C. F. (2009) Protein
possibilities for (bio)conjugation reactions. Several examples C-terminal modification through thioacid/azide amidation. Bio-
of the synthesis and applications of biomedical and pharma- conjugate Chem. 20, 197–200.
2014 Bioconjugate Chem., Vol. 20, No. 11, 2009 van Dijk et al.

(16) Tornøe, C. W., Meldal, M. (2001) Peptidotriazoles: Copper(I)- (38) Huisgen, R. (1968) Cycloadditions - definition, classification,
catalyzed 1,3-dipolar cycloadditions on solid phase. In Peptides: and characterization. Angew. Chem., Int. Ed. Engl. 7, 321–8.
The WaVe of the Future: Proceedings of the 2nd International (39) Huisgen, R., Szeimies, G., and Mobius, L. (1967) 1.3-Dipolare
and the 17th American Peptide Symposium; Lebl, M., and Cycloadditionen, XXXII. Kinetik der Additionen organischer
Houghten, R. A. Eds.; American Peptide Society and Kluwer Azide an CC-Mehrfachbindungen. Chem. Ber. 100, 2494–507.
Academic Publishers: San Diego, 263-4. (40) Horne, W. S., Yadav, M. K., Stout, C. D., and Ghadiri, M. R.
(17) Tornøe, C. W., Christensen, C., and Meldal, M. (2002) (2004) Heterocyclic peptide backbone modifications in an alpha-
Peptidotriazoles on solid phase: [1,2,3]-triazoles by regiospecific helical coiled coil. J. Am. Chem. Soc. 126, 15366–7.
copper(I)-catalyzed 1,3-dipolar cycloadditions of terminal alkynes (41) Angelo, N. G., and Arora, P. S. (2005) Nonpeptidic foldamers
to azides. J. Org. Chem. 67, 3057–64. from amino acids: Synthesis and characterization of 1,3-substituted
(18) Rostovtsev, V. V., Green, L. G., Fokin, V. V., and Sharpless, triazole oligomers. J. Am. Chem. Soc. 127, 17134–5.
K. B. (2002) A stepwise Huisgen cycloaddition process: Cop- (42) van Maarseveen, J. H., Horne, W. S., and Ghadiri, M. R.
per(I)-catalyzed regioselective “ligation” of azides and terminal (2005) Efficient route to C2 symmetric heterocyclic backbone
alkynes. Angew. Chem., Int. Ed. 41, 2596–9. modified cyclic peptides. Org. Lett. 7, 4503–6.
(19) Binder, W. H., and Sachsenhofer, R. (2007) ‘Click’ chemistry in (43) Wittig, G., and Krebs, A. (1961) Zur Existenz niedergliedriger
polymer and materials science. Macromol. Rapid Commun. 28, 15– Cycloalkine, I. Chem. Ber. 94, 3260–75.
54. (44) Agard, N. J., Prescher, J. A., and Bertozzi, C. R. (2004) A
(20) Binder, W. H., and Sachsenhofer, R. (2008) ‘Click’ chemistry strain-promoted [3 + 2] azide-alkyne cycloaddition for covalent
in polymer and material science: An update. Macromol. Rapid modification of biomolecules in living systems. J. Am. Chem.
Commun. 29, 952–81. Soc. 126, 15046–7.
(21) Meldal, M. (2008) Polymer “Clicking” by CuAAC reactions. (45) Ning, X. H., Guo, J., Wolfert, M. A., and Boons, G. J. (2008)
Macromol. Rapid Commun. 29, 1016–51. Visualizing metabolically labeled glycoconjugates of living cells
(22) Nandivada, H., Jiang, X. W., and Lahann, J. (2007) Click by copper-free and fast Huisgen cycloadditions. Angew. Chem.,
chemistry: Versatility and control in the hands of materials Int. Ed. 47, 2253–5.
scientists. AdV. Mater. 19, 2197–208. (46) Codelli, J. A., Baskin, J. M., Agard, N. J., and Bertozzi, C. R.
(23) Lutz, J. F. (2007) 1,3-dipolar cycloadditions of azides and (2008) Second-generation difluorinated cyclooctynes for copper-
alkynes: A universal ligation tool in polymer and materials free click chemistry. J. Am. Chem. Soc. 130, 11486–93.
science. Angew. Chem., Int. Ed. 46, 1018–25.
(47) Sletten, E. M., and Bertozzi, C. R. (2008) A hydrophilic
(24) Johnson, J. A., Finn, M. G., Koberstein, J. T., and Turro, N. J. (2008)
azacyclooctyne for Cu-free click chemistry. Org. Lett. 10, 3097–
Construction of linear polymers, dendrimers, networks, and other
9.
polymeric architectures by copper-catalyzed azide-alkyne cycloaddition
(48) Baskin, J. M., Prescher, J. A., Laughlin, S. T., Agard, N. J.,
“Click” chemistry. Macromol. Rapid Commun. 29, 1052–72.
Chang, P. V., Miller, I. A., Lo, A., Codelli, J. A., and Bertozzi,
(25) Angell, Y. L., and Burgess, K. (2007) Peptidomimetics via
C. R. (2007) Copper-free click chemistry for dynamic in vivo
copper-catalyzed azide-alkyne cycloadditions. Chem. Soc. ReV.
imaging. Proc. Natl. Acad. Sci. U.S.A. 104, 16793–7.
36, 1674–89.
(26) Pieters, R. J., Rijkers, D. T. S., and Liskamp, R. M. J. (2007) (49) van Berkel, S. S., Dirks, A. T., Debets, M. F., van Delft, F. L.,
Application of the 1,3-dipolar cycloaddition reaction in chemical Cornelissen, J. J. L. M., Nolte, R. J. M., and Rutjes, F. P. J. T.
biology: Approaches toward multivalent carbohydrates and peptides (2007) Metal-free triazole formation as a tool for bioconjugation.
and peptide-based polymers. QSAR Comb. Sci. 26, 1181–90. ChemBioChem 8, 1504–8.
(27) Dirks, A. J., Cornelissen, J. J. L. M., van Delft, F. L., van Hest, (50) Mocharla, V. P., Colasson, B., Lee, L. V., Roper, S., Sharpless,
J. C. M., Nolte, R. J. M., Rowan, A. E., and Rutjes, F. P. J. T. K. B., Wong, C. H., and Kolb, H. C. (2004) In situ click
(2007) From (bio)molecules to biohybrid materials with the click chemistry: enzyme-generated inhibitors of carbonic anhydrase
chemistry approach. QSAR Comb. Sci. 26, 1200–10. II. Angew. Chem., Int. Ed. 44, 116–20.
(28) Le Droumaguet, B., and Velonia, K. (2008) Click chemistry: (51) Zhang, L., Chen, X. G., Xue, P., Sun, H. H. Y., Williams,
A powerful tool to create polymer-based macromolecular I. D., Sharpless, K. B., Fokin, V. V., and Jia, G. C. (2005)
chimeras. Macromol. Rapid Commun. 29, 1073–89. Ruthenium-catalyzed cycloaddition of alkynes and organic
(29) Lutz, J. F., and Zarafshani, Z. (2008) Efficient construction of azides. J. Am. Chem. Soc. 127, 15998–9.
therapeutics, bioconjugates, biomaterials and bioactive surfaces using (52) Majireck, M. M., and Weinreb, S. M. (2006) A study of the
azide-alkyne “click” chemistry. AdV. Drug DeliVery ReV. 60, 958– scope and regioselectivity of the ruthenium-catalyzed [3 +
70. 2]-cycloaddition of azides with internal alkynes. J. Org. Chem.
(30) Fournier, D., Hoogenboom, R., and Schubert, U. S. (2007) 71, 8680–3.
Clicking polymers: a straightforward approach to novel macro- (53) Rasmussen, L. K., Boren, B. C., and Fokin, V. V. (2007)
molecular architectures. Chem. Soc. ReV. 36, 1369–80. Ruthenium-catalyzed cycloaddition of aryl azides and alkynes.
(31) Read, E. S., and Armes, S. P. (2007) Recent advances in shell Org. Lett. 9, 5337–9.
cross-linked micelles. Chem. Commun. 3021–35. (54) Boren, B. C., Narayan, S., Rasmussen, L. K., Zhang, L., Zhao,
(32) Lecomte, P., Riva, R., Jerome, C., and Jerome, R. (2008) H. T., Lin, Z. Y., Jia, G. C., and Fokin, V. V. (2008) Ruthenium-
Macromolecular engineering of biodegradable polyesters by ring- catalyzed azide-alkyne cycloaddition: Scope and mechanism. J. Am.
opening polymerization and ‘Click’ chemistry. Macromol. Rapid Chem. Soc. 130, 8923–30 (correction: J. Am. Chem. Soc., 14900).
Commun. 29, 982–97. (55) Jemal, M., and Hawthorne, D. J. (1997) Quantitative deter-
(33) Lutz, J. F. (2008) Copper-free azide-alkyne cycloadditions: New mination of BMS186716, a thiol compound, in dog plasma by
insights and perspectives. Angew. Chem., Int. Ed. 47, 2182–4. high-performance liquid chromatography-positive ion electro-
(34) Meldal, M., and Tornøe, C. W. (2008) Cu-catalyzed azide- spray mass spectrometry after formation of the methyl acrylate
alkyne cycloaddition. Chem. ReV. 108, 2952–3015. adduct. J. Chromatogr., B: Biomed. Sci. Appl. 693, 109–16.
(35) Bock, V. D., Hiemstra, H., and van Maarseveen, J. H. (2005) (56) Romanowska, A., Meunier, S. J., Tropper, F. D., Laferriere,
Cu(I)-catalyzed alkyne-azide “click” cycloadditions from a mecha- C. A., and Roy, R. (1994) Michael additions for syntheses of
nistic and synthetic perspective. Eur. J. Org. Chem. 51–68. neoglycoproteins. Methods Enzymol. 242, 90–101.
(36) Moses, J. E., and Moorhouse, A. D. (2007) The growing (57) Masri, M. S., and Friedman, M. (1988) Protein reactions with
applications of click chemistry. Chem. Soc. ReV. 36, 1249–62. methyl and ethyl vinyl sulfones. J. Protein Chem. 7, 49–54.
(37) Moorhouse, A. D., and Moses, J. E. (2008) Click chemistry (58) Morpurgo, M., Veronese, F. M., Kachensky, D., and Harris,
and medicinal chemistry: a case of “cyclo-addiction”. ChemMed- J. M. (1996) Preparation of characterization of poly(ethylene
Chem 3, 715–23. glycol) vinyl sulfone. Bioconjugate Chem. 7, 363–8.
Reviews Bioconjugate Chem., Vol. 20, No. 11, 2009 2015

(59) Pounder, R. J., Stanford, M. J., Brooks, P., Richards, S. P., (81) Yu, T.-B., Bai, J. Z., and Guan, Z. (2009) Cycloaddition-
and Dove, A. P. (2008) Metal free thiol-maleimide ‘Click’ promoted self-assembly of a polymer into well-defined β sheets and
reaction as a mild functionalisation strategy for degradable hierarchical nanofibrils. Angew. Chem., Int. Ed. 48, 1097–101.
polymers. Chem. Commun. 5158–60. (82) Agut, W., Agnaou, R., Lecommandoux, S., and Taton, D.
(60) Gress, A., Volkel, A., and Schlaad, H. (2007) Thio-click (2008) Synthesis of block copolypeptides by click chemistry.
modification of poly[2-(3-butenyl)-2-oxazoline]. Macromolecules Macromol. Rapid Commun. 29, 1147–55.
40, 7928–33. (83) Parrish, B., Breitenkamp, R. B., and Emrick, T. (2005) PEG-
(61) Dondoni, A. (2008) The emergence of thiol-ene coupling as and peptide-grafted aliphatic polyesters by click chemistry. J. Am.
a click process for materials and bioorganic chemistry. Angew. Chem. Soc. 127, 7404–10.
Chem., Int. Ed. 47, 8995–7. (84) Tirrell, D. A. (2004) Chemical biology - Hitting the sweet
(62) Geng, Y., Discher, D. E., Justynska, J., and Schlaad, H. (2006) spot. Nature 430, 837.
Grafting short peptides onto polybutadiene-block-poly(ethylene (85) Dwek, R. A. (1996) Glycobiology: Toward understanding the
oxide): A platform for self-assembling hybrid amphiphiles. function of sugars. Chem. ReV. 96, 683–720.
Angew. Chem., Int. Ed. 45, 7578–81. (86) Kiessling, L. L., and Cairo, C. W. (2002) Hitting the sweet
(63) Hordyjewicz-Baran, Z., You, L. C., Smarsly, B., Sigel, R., spot. Nat. Biotechnol. 20, 234–5.
and Schlaad, H. (2007) Bioinspired polymer vesicles based on (87) Ambrosi, M., Cameron, N. R., and Davis, B. G. (2005)
hydrophilically modified polybutadienes. Macromolecules 40, Lectins: tools for the molecular understanding of the glycocode.
3901–3. Org. Biomol. Chem. 3, 1593–608.
(64) Justynska, J., Hordyjewicz, Z., and Schlaad, H. (2005) Toward (88) Lundquist, J. J., and Toone, E. J. (2002) The cluster glycoside
a toolbox of functional block copolymers via free-radical addition effect. Chem. ReV. 102, 555–78.
of mercaptans. Polymer 46, 12057–64. (89) Srinivasachari, S., Liu, Y. M., Zhang, G. D., Prevette, L., and
(65) Woghiren, C., Sharma, B., and Stein, S. (1993) Protected thiol- Reineke, T. M. (2006) Trehalose click polymers inhibit nano-
polyethylene glycol: a new activated polymer for reversible particle aggregation and promote pDNA delivery in serum. J. Am.
protein modification. Bioconjugate Chem. 4, 314–8. Chem. Soc. 128, 8176–84.
(66) Goessl, A., Tirelli, N., and Hubbell, J. A. (2004) A hydrogel (90) Srinivasachari, S., Liu, Y. M., Prevette, L. E., and Reineke, T. M.
system for stimulus-responsive, oxygen-sensitive in situ gelation. (2007) Effects of trehalose click polymer length on pDNA complex
J. Biomater. Sci. Polym. Ed. 15, 895–904. stability and delivery efficacy. Biomaterials 28, 2885–98.
(67) Hoyle, C. E., Lee, T. Y., and Roper, T. (2004) Thiol-enes: (91) Ladmiral, V., Mantovani, G., Clarkson, G. J., Cauet, S., Irwin,
Chemistry of the past with promise for the future. J. Polym. Sci., J. L., and Haddleton, D. M. (2006) Synthesis of neoglycopoly-
Part A: Polym. Chem. 42, 5301–38. mers by a combination of “click chemistry” and living radical
(68) Morgan, C. R., Magnotta, F., and Ketley, A. D. (1977) Thiol/ polymerization. J. Am. Chem. Soc. 128, 4823–30.
ene photocurable polymers. J. Polym. Sci., Polym. Chem. Ed. (92) Geng, J., Mantovani, G., Tao, L., Nicolas, J., Chen, G., Wallis,
15, 627–45. R., Mitchell, D. A., Johnson, B. R., Evans, S. D., and Haddleton,
(69) van Hest, J. C. M., and Tirrell, D. A. (2001) Protein-based D. M. (2007) Site-directed conjugation of “clicked” glycopoly-
materials, toward a new level of structural control. Chem. mers to form glycoprotein mimics: binding to mammalian lectin
Commun. 1897–904. and induction of immunological function. J. Am. Chem. Soc. 129,
(70) Kricheldorf, H. R. (1987) Alpha-Aminoacid-N-Carboxyanhy- 15156–63.
drides and Related Heterocycles, Springer, Berlin. (93) Chen, G. J., Amajjahe, S., and Stenzel, M. H. (2009) Synthesis
(71) Chiu, H. C., Kopeckova, P., Deshmane, S. S., and Kopecek, of thiol-linked neoglycopolymers and thermo-responsive gly-
J. (1997) Lysosomal degradability of poly(alpha-amino acids). comicelles as potential drug carrier. Chem. Commun. 1198–200.
J. Biomed. Mater. Res. 34, 381–92. (94) Diehl, C., and Schlaad, H. (2009) Thermo-responsive poly-
(72) Metzke, M., O’Connor, N., Maiti, S., Nelson, E., and Guan, oxazolines with widely tuneable LCST. Macromol. Biosci. 9,
Z. B. (2005) Saccharide-peptide hybrid copolymers as bioma- 157–61.
terials. Angew. Chem., Int. Ed. 44, 6529–33. (95) Nurmi, L., Lindqvist, J., Randev, R., Syrett, J., and Haddleton,
(73) Haider, M., Megeed, Z., and Ghandehari, H. (2004) Geneti- D. M. (2009) Glycopolymers via ctalytic chain transfer poly-
cally engineered polymers: status and prospects for controlled merisation (CCTP), huisgens cycloaddition and thiol-ene double
release. J. Controlled Release 95, 1–26. click reactions. Chem. Commun. 2727–9.
(74) Kent, S. B. H. (2009) Total chemical synthesis of proteins. (96) El-Aneed, A. (2004) Current strategies in cancer gene therapy.
Chem. Soc. ReV. 38, 338–51. Eur. J. Pharmacol. 498, 1–8.
(75) Zhang, Z. S., and Fan, E. (2006) Solid phase synthesis of (97) Zhou, H. S., Liu, D. P., and Liang, C. C. (2004) Challenges
peptidotriazoles with multiple cycles of triazole formation. and strategies: the immune responses in gene therapy. Med. Res.
Tetrahedron Lett. 47, 665–9. ReV. 24, 748–61.
(76) Aucagne, V., and Leigh, D. A. (2006) Chemoselective (98) Mintzer, M. A., and Simanek, E. E. (2009) Nonviral vectors
formation of successive triazole linkages in one pot: “ Click- for gene delivery. Chem. ReV. 109, 259–302.
click” chemistry. Org. Lett. 8, 4505–7. (99) White, S., Szewczyk, J. W., Turner, J. M., Baird, E. E., and
(77) Orsini, A., Viterisi, A., Bodlenner, A., Weibel, J.-M., and Pale, Dervan, P. B. (1998) Recognition of the four Watson-Crick base
P. (2005) A chemoselective deprotection of trimethylsilyl acetylenes pairs in the DNA minor groove by synthetic ligands. Nature 391,
catalyzed by silver salts. Tetrahedron Lett. 46, 2259–62. 468–71.
(78) Carpita, A., Mannocci, L., and, and Rossi, R. (2005) Silver(I)- (100) Petersen, H., Merdan, T., Kunath, K., Fischer, D., and Kissel,
catalysed protiodesilylation of 1-(trimethylsilyl)-1-alkynes. Eur. T. (2002) Poly(ethylenimine-co-L-lactamide-co-succinamide): a
J. Org. Chem. 1859–64. biodegradable polyethylenimine derivative with an advantageous
(79) van Dijk, M., Mustafa, K., Dechesne, A. C., van Nostrum, pH-dependent hydrolytic degradation for gene delivery. Biocon-
C. F., Hennink, W. E., Rijkers, D. T. S., and Liskamp, R. M. J. jugate Chem. 13, 812–21.
(2007) Synthesis of peptide-based polymers by microwave- (101) Luten, J., van Nostrum, C. F., De Smedt, S. C., and Hennink,
assisted cycloaddition backbone polymerization. Biomacromol- W. E. (2008) Biodegradable polymers as non-viral carriers for
ecules 8, 327–30. plasmid DNA delivery. J. Controlled Release 126, 97–110.
(80) van Dijk, M., Nollet, M. L., Weijers, P., Dechesne, A. C., (102) Jiang, X., Lok, M. C., and Hennink, W. E. (2007) Degradable-
van Nostrum, C. F., Hennink, W. E., Rijkers, D. T. S., and brushed pHEMA-pDMAEMA synthesized via ATRP and click
Liskamp, R. M. J. (2008) Synthesis and characterization of chemistry for gene delivery. Bioconjugate Chem. 18, 2077–84.
biodegradable peptide-based polymers prepared by microwave- (103) Varki, A. (1993) Biological roles of oligosaccharides: all of
assisted click chemistry. Biomacromolecules 9, 2834–43. the theories are correct. Glycobiology 3, 97–130.
2016 Bioconjugate Chem., Vol. 20, No. 11, 2009 van Dijk et al.

(104) Kramer, R. H., and Karpen, J. W. (1998) Spanning binding (122) Peppas, N. A., Bures, P., Leobandung, W., and Ichikawa,
sites on allosteric proteins with polymer-linked ligand dimers. H. (2000) Hydrogels in pharmaceutical formulations. Eur.
Nature 395, 710–3. J. Pharm. Biopharm. 50, 27–46.
(105) Mammen, M., Choi, S.-K., and Whitesides, G. M. (1998) (123) Ossipov, D. A., and Hilborn, J. (2006) Poly(vinyl alcohol)-
Polyvalent interactions in biological systems: implications for based hydrogels formed by “click chemistry”. Macromolecules
design and use of multivalent ligands and inhibitors. Angew. 39, 1709–18.
Chem., Int. Ed. 37, 2754–94. (124) Malkoch, M., Vestberg, R., Gupta, N., Mespouille, L.,
(106) Joosten, J. A. F., Tholen, N. T. H., El Maate, F. A., Brouwer, Dubois, P., Mason, A. F., Hedrick, J. L., Liao, Q., Frank, C. W.,
A. J., van Esse, G. W., Rijkers, D. T. S., Liskamp, R. M. J., and Kingsbury, K., and Hawker, C. J. (2006) Synthesis of well-
Pieters, R. J. (2005) High-yielding microwave-assisted synthesis defined hydrogel networks using Click chemistry. Chem. Com-
of triazole-linked glycodendrimers by copper-catalyzed [3 + 2] mun. 2774–6.
cycloaddition. Eur. J. Org. Chem. 3182–5. (125) Crescenzi, V., Cornelio, L., Di Meo, C., Nardecchia, S., and
(107) Rijkers, D. T. S., van Esse, G. W., Merkx, R., Brouwer, A. J., Lamanna, R. (2007) Novel hydrogels via click chemistry:
Jacobs, H. J. J., Pieters, R. J., and Liskamp, R. M. J. (2005) Synthesis and potential biomedical applications. Biomacromol-
Efficient microwave-assisted synthesis of multivalent dendrimeric ecules 8, 1844–50.
peptides using cycloaddition reaction (click) chemistry. Chem. (126) Polizzotti, B. D., Fairbanks, B. D., and Anseth, K. S. (2008)
Commun. 4581–3. Three-dimensional biochemical patterning of click-based com-
(108) Dijkgraaf, I., Rijnders, A. Y., Soede, A., Dechesne, A. C., posite hydrogels via thiolene photopolymerization. Biomacro-
van Esse, G. W., Brouwer, A. J., Corstens, F. H., Boerman, O. C., molecules 9, 1084–7.
Rijkers, D. T. S., and Liskamp, R. M. J. (2007) Synthesis of
(127) Mespouille, L., Coulembier, O., Paneva, D., Degée, E.,
DOTA-conjugated multivalent cyclic-RGD peptide dendrimers
Rashkov, I., and Dubois, P. (2008) Synthesis of adaptative and
via 1,3-dipolar cycloaddition and their biological evaluation:
amphiphilic polymer model conetworks by versatile combination
implications for tumor targeting and tumor imaging purposes.
of ATRP, ROP and “click chemistry”. J. Polym. Sci., Part A:
Org. Biomol. Chem. 5, 935–44.
Polym. Chem. 46, 4997–5013.
(109) Killops, K. L., Campos, L. M., and Hawker, C. J. (2008)
Robust, efficient, and orthogonal synthesis of dendrimers via (128) Johnson, J. A., Finn, M. G., Koberstein, J. T., and Turro,
thiol-ene “click” chemistry. J. Am. Chem. Soc. 130, 5062–4. N. J. (2007) Synthesis of photocleavable linear macromonomers
(110) Jonkheijm, P., Weinrich, D., Kohn, M., Engelkamp, H., by ATRP and star macromonomers by a tandem ATRP-click
Christianen, P. C., Kuhlmann, J., Maan, J. C., Nusse, D., reaction: precursors to photodegradable model networks. Mac-
Schroeder, H., Wacker, R., Breinbauer, R., Niemeyer, C. M., romolecules 40, 3589–98.
and Waldmann, H. (2008) Photochemical surface patterning by (129) Johnson, J. A., Baskin, J. M., Bertozzi, C. R., Koberstein,
the thiol-ene reaction. Angew. Chem., Int. Ed. 47, 4421–4. J. T., and Turro, N. J. (2008) Copper-free click chemistry for
(111) Xia, Y. N. (2008) Nanomaterials at work in biomedical the in situ crosslinking of photodegradable star polymers. Chem.
research. Nat. Mater. 7, 758–60. Commun. 3064–6.
(112) Schiffelers, R. M., Ansari, A., Xu, J., Zhou, Q., Tang, Q., (130) Xu, X., Chen, C., Wang, Z., Wang, G., Cheng, S., Zhang,
Storm, G., Molema, G., Lu, P. Y., Scaria, P. V., and Woodle, X., and Zhou, R. (2008) “Click” chemistry for in situ formation
M. C. (2004) Cancer siRNA therapy by tumor selective delivery of thermoresponsive p(NIPAAm-co-HEMA)-based hydrogels. J.
with ligand-targeted sterically stabilized nanoparticle. Nucleic Polym. Sci., Part A: Polym. Chem. 46, 5263–77.
Acids Res. 32, e149. (131) Lui, S. Q., Ee, P. L. R., Ke, C. Y., Hendrick, J. L., and Yang,
(113) Panyam, J., and Labhasetwar, V. (2003) Biodegradable Y. Y. (2009) Biodegradable poly(ethylene glycol)-peptide hy-
nanoparticles for drug and gene delivery to cells and tissue. AdV. drogels with well-defined structure and properties for cell
Drug DeliVery ReV. 55, 329–47. delivery. Biomaterials 30, 1453–61.
(114) Opsteen, J. A., Brinkhuis, R. P., Teeuwen, R. L. M., Löwik, (132) Qiu, B., Stefanos, S., Ma, J., Lalloo, A., Perry, B. A.,
D. W. P. M., and van Hest, J. C. M. (2007) “Clickable” Leibowitz, M. J., Sinko, P. J., and Stein, S. (2003) A hydrogel
polymersomes. Chem. Commun. 3136–8. prepared by in situ cross-linking of a thiol-containing poly(eth-
(115) Such, G. K., Quinn, J. F., Quinn, A., Tjipto, E., and Caruso, ylene glycol)-based copolymer: a new biomaterial for protein
F. (2006) Assembly of ultrathin polymer multilayer films by click drug delivery. Biomaterials 24, 11–8.
chemistry. J. Am. Chem. Soc. 128, 9318–9. (133) Metters, A., and Hubbell, J. (2005) Network formation and
(116) Such, G. K., Tjipto, E., Postma, A., Johnston, A. P., and degradation behavior of hydrogels formed by Michael-type
Caruso, F. (2007) Ultrathin, responsive polymer click capsules. addition reactions. Biomacromolecules 6, 290–301.
Nano Lett. 7, 1706–10. (134) Schoenmakers, R. G., van de Wetering, P., Elbert, D. L.,
(117) De Geest, B. G., Van Camp, W., Du Prez, F. E., De Smedt, and Hubbell, J. A. (2004) The effect of the linker on the
S. C., Demeester, J., and Hennink, W. E. (2008) Degradable hydrolysis rate of drug-linked ester bonds. J. Controlled Release
multilayer films and hollow capsules via a ‘click’ strategy. 95, 291–300.
Macromol. Rapid Commun. 29, 1111–8. (135) Hiemstra, C., Zhong, Z., Van Tomme, S. R., van Steenber-
(118) De Geest, B. G., Van Camp, W., Du Prez, F. E., De Smedt, S. C., gen, M. J., Jacobs, J. J. L., Den Otter, W., Hennink, W. E., and
Demeester, J., and Hennink, W. E. (2008) Biodegradable microcapsules Feijen, J. (2007) In vitro and in vivo protein delivery from in
designed via ‘click’ chemistry. Chem. Commun. 190–2. situ forming poly(ethylene glycol)-poly(lactide) hydrogels. J.
(119) Connal, L. A., Kinnane, C. R., Zelikin, A. N., and Caruso, Controlled Release 119, 320–7.
F. (2009) Stabilization and functionalization of polymer multi- (136) Hiemstra, C., van der Aa, L. J., Zhong, Z. Y., Dijkstra, P. J.,
layers and capsules via thiol-ene click chemistry. Chem. Mater. and Feijen, J. (2007) Novel in situ forming, degradable dextran
21, 576–8. hydrogels by Michael addition chemistry: Synthesis, rheology,
(120) Sousa-Herves, A., Fernandez-Megia, E., and Riguera, R. and degradation. Macromolecules 40, 1165–73.
(2008) Synthesis and supramolecular assembly of clicked anionic (137) Tae, G., Kim, Y. J., Choi, W. I., Kim, M., Stayton, P. S.,
dendritic polymers into polyion complex micelles. Chem. Com- and Hoffman, A. S. (2007) Formation of a novel heparin-based
mun. 3136–8. hydrogel in the presence of heparin-binding biomolecules.
(121) Breed, D. R., Thibault, R., Xie, F., Wang, Q., Hawker, C. J., Biomacromolecules 8, 1979–86.
and Pine, D. J. (2009) Functionalization of polymer microspheres
using click chemistry. Langmuir 25, 4370–6. BC900087A

You might also like