Chapter 5

Noncovalent Functionalization of Cell Surface

Debjit Dutta, Srujan K. Marepally and Praveen Kumar Vemula,    Institute for Stem Cell Biology and Regenerative Medicine, National Centre for Biological Sciences-Campus, UAS-GKVK, Bangalore, India

The cell membrane is a highly heterogenous and dynamic environment that dictates many biological functions like cell–cell and cell–niche communication, cell adhesion, migration, and differentiation. Thus, cell surface is a potential target for introducing different functional groups and molecular and therapeutic probes for diverse biomedical applications. In this chapter, we have discussed different noncovalent approaches such as membrane fusion, electrostatic and hydrophobic interactions using cationic and polymeric amphiphiles, and use of metabolic engineering and bioorthogonal click chemistry for the modification of cell surface. The advantages of noncovalent methods over covalent and genetic methods of modification and their applications in cell-based therapy, bioimaging, and tissue engineering have also been described. Future directions in terms of developing dynamic approaches to modify cell surfaces have been discussed in brief.

Keywords

Noncovalent methods; self-assembly; cell surface; electrostatic interactions; hydrophobic interactions; bioorthogonal chemistry; amphiphiles; cell therapy

5.1 Introduction

Surface of the cell plays a pivotal role in defining the cell fate. A wide range of functional biomolecules that anchored on the cell surface perform different roles. A cell communicates with other cells, environment, pathogens, and nutrients through its surface. In addition, cell migration is highly dependent on the cell membrane. Thus, cell fate could be modulated through engineering cell surface (Figure 5.1). Cell surface is a fertile ground to manipulate cell phenotypes and biological fates. While these manipulations are key tools to answer some of the basic biological questions [1], they could be instrumental in developing novel diagnostic and therapeutic systems [2] through engineering of cell surface. Applications of genetic tools for manipulating proteins and probing cellular functions have revolutionized biology and medicine. These tools were proven to be successful, but their applications are restricted due to their inability to introduce novel functionality into biomolecules or interrogate the distinct chemical entities that control cellular responses. To address these limitations, multiple research teams have developed a set of chemoselective reactions, known as bioorthogonal reactions, with minimal cross-reactivity and toxicity in biological systems.

image
Figure 5.1 Schematic of cell surface that plays a vital role in numerous cell functions.

A number of synergistic approaches combining chemistry, material science, and biology have been applied to engineer and manipulate living cells. Desirable chemical functionalities and materials can be introduced onto the cell surface by covalent and noncovalent methodologies, as well as through specific biological recognition events, including antibody–antigen and ligand–receptor interactions. These strategies have been used in a wide range of applications that include designing the biomaterial scaffolds [3,4] to control cell fate, labeling of cells with molecular and nanoparticle probes to visualize cellular processes and molecular pathways [57], delivering diverse species into cells [8,9], and patterning cells for drug discovery [10,11]. To date, cell surface engineering has been primarily achieved through the use of molecular biology and genetic engineering [12]. However, perturbations of inherent cellular physiology with genetic engineering may significantly interfere with cellular functions. Thus, there is a need to develop tools that will provide simple alternatives to genetic and biosynthetic pathways.

Nongenetic cell surface modifications have been generally achieved using the following methods: (i) covalent conjugation of desired molecules to surface functional groups like amines [11,1320], (ii) incorporation of amphiphilic polymers into the lipid bilayer membrane of cells by harnessing hydrophobic interactions [21], (iii) electrostatic interactions between cationic polymers and a negatively charged surface [15], and (iv) metabolic engineering of ketone- or azide-modified carbohydrates to display desired functional groups on the cell surface, followed by bioorthogonal click reactions to tether molecule of choice [2224]. Schematic representation of some examples of engineering of cell surface is shown in Figures 5.2 and 5.3.

image
Figure 5.2 (A) and (B) Schematic of engineering the cell surface using noncovalent methods. Typically, hydrophobic and ionic interactions have been utilized to immobilize the molecules of interest on the surface of the cell.
image
Figure 5.3 Schematics of metabolic approach of cell surface engineering. (A) The per-O-acetylated azido sugars passively diffuse across the cell membrane. In the cytosol, the acetyl groups are removed by intracellular esterases, the azido sugar is activated to a nucleotide sugar donor (X=uridine diphosphate or cytidine monophosphate) and exported to the cell surface through glycosylation. (B) Different chemical structures of azido sugars and the glycan subtypes labeled. (C) The Staudinger ligation: triaryl phosphines react selectively with alkyl azides to form an amide bond to the alkyl substituent, such as an azido sugar. Source: Reprinted with permission from Ref. [24] © NPG 2007.

This chapter focuses on nanotechnology approaches that could be applied for the application of cell surface engineering and their limitations.

5.2 Methods of Cell Surface Engineering—Applications and Recent Developments

5.2.1 Targeting Ability of Surface Engineered Cells

Cell therapy involving systemic infusion and targeted delivery of progenitor/mesenchymal stem cells (MSCs) and other cell types to damage and diseased tissues hold tremendous promise for the treatment of a number of diseases such as cardiomyopathy, cancer, and fibrosis [25]. MSCs are capable of differentiating into multiple lineages and produce extracellular matrix (bone, cartilage, fat) and can also secrete immunomodulatory cytokines to reduce inflammation. Although they are currently being tested in more than 100 clinical trials for treatment of multiple diseases, including graft versus host disease, myocardial infarction, multiple sclerosis, and skeletal tissue repair, the major challenge in MSC therapy is the delivery of MSCs efficiently to targeted location. Systemically administered therapeutic MSCs have poor expression of homing ligands or lose them during the culture expansion and only less than 1% of culture-expanded MSCs can actually home to the target tissue [26]. Therefore, efforts have been made to introduce cell homing ligands onto cell membranes. Cell homing ligands (on the homing cell) and receptors (on the endothelium) allow homing cells to tether, roll, adhere, and then transmigrate on endothelium as part of the cell homing cascade. A great effort has been made to achieve targeted delivery of cells, and a wide range of methods such as genetic and enzymatic engineering [2729] treatment with cytokines [30] and chemical approaches have been developed [31,32].

5.2.2 Noncovalent Immobilization of Agents on the Surface of Cells

Although a covalent approach offered a method to modify the cell surface for longer duration, often, covalent functionalization might interfere with the function of native proteins or other molecules that are present on the cell surface. Thus, to overcome these potential limitations, a robust noncovalent approach to transiently engineer the cell surface has been developed [33]. In this method, unilamellar vesicles composed of biotinylated lipid were fused with the cell membrane to anchor biotin on the cell membrane of MSCs (Figure 5.4). Intercalation of the lipid vesicles with lipid bilayer of the cell membrane is a robust phenomenon; it offers numerous opportunities to immobilize any desired molecules on the surface of the cell membrane. The biotinylated surface of the cell was further hybridized with multivalent streptavidin and followed by biotinylated SLeX (Figure 5.4). As cell membrane undergoes a flip–flop process, lipids in the membrane are rearranged and recycled. Thus, modification on the cell surface through noncovalent liposomal hybridization offers a transient cell surface medication and avoids nonspecific modification of cell surface molecules. In addition, transient immobilization of adhesion ligands offered a great advantage through allowing for initial adhesion and cytokine activation, not impeding subsequent extravasation.

image
Figure 5.4 (A) Schematic showing formation of vesicles from biotinylated lipid and modification of MSCs with biotinylated lipid vesicles. (B) Biotinylated cell surface was complexed with streptavidin followed by biotinylated SLeX. (C) Induced rolling response of SLeX-modified MSCs on a P-selectin-coated surface. hMSC, human mesenchymal stem cell. Source: Reprinted with permission from Ref. [33] © Elsevier 2010.

Furthermore, the rolling response of the SLeX-modified MSCs with P-selectin-coated substrates has been examined using a flow chamber assay. MSCs modified with SLeX showed considerably lower velocities on immobilized P-selectin substrates compared to phosphate buffer saline (PBS) treated cells (Figure 5.5). SLeX-modified MSCs exhibited ~eightfold less velocity compared to unmodified MSCs at a shear stress of 0.5 dyn/cm2 which suggested the effect of noncovalently modified surface with SLeX.

image
Figure 5.5 Flow chamber assay on P-selectin-coated surface. Velocity of SLeX-modified and PBS-treated MSCs at 0.5 dyn/cm2. Source: Reprinted with permission from Ref. [33] © Elsevier 2010.

Inherent properties of stem cells are sensitive to minor modifications of cell machinery and microenvironment. Thus, it is important to investigate the effect of cell surface modification through liposomal fusion on the cell fate. Therefore, Sarkar et al. [33] have thoroughly characterized the phenotype of MSCs after modification through liposomal fusion (Figure 5.6). Viability and proliferative rates of the modified MSCs were comparable to unmodified MSCs. In addition, the adhesion kinetics of modified MSCs and untreated MSCs are similar on tissue culture polystyrene plates, which indicated that inherent properties of the cell surface did not compromise due to noncovalent liposomal fusion. Another important characteristic of MSCs is the ability to differentiate into multiple lineages that did not compromise modification of the cell surface with biotinylated lipids and subsequent complexation of streptavidin and biotinylated SLeX. Modified MSCs were efficiently differentiated into osteogenic and adipogenic lineages as shown by alkaline phosphatase activity and Oil Red O staining, respectively (Figure 5.6). Those results clearly suggested that modifying the cells with rolling ligands through a noncovalent lipid vesicle approach did not affect the cell phenotype and thus cells may maintain their normal functional characteristics that will be important for cell therapy. This is a compelling example for the broad scope of cell surface modification through noncovalent approaches.

image
Figure 5.6 (A) Viability of the SLeX-modified MSCs at 0 h and after 48 h. (B) Adherence of SLeX-modified cells compared to the PBS-treated cells. (C) Doubling time of the SLeX-modified cells compared to PBS-treated cells. (D) Alkaline phosphatase and Oil Red O staining, 23 days after addition of osteogenic and adipogenic differentiation media to SLeX-modified MSCs. VB, biotinylated vesicles; BSLeX, biotinylated SLeX; S, streptavidine (−Control: SLeX-modified MSCs in cell expansion media; +Control: MSCs in differentiation media; Experimental group: SLeX-modified MSCs in differentiation media). Source: Reprinted with permission from Ref. [33] © Elsevier 2010.

5.2.3 Noncovalent Cell Surface Modification Followed by Covalent Chemistry to Enhance the Cell Interactions

As described in previous sections, where cell surface molecules were covalently reacted with biotinylated-N-hydroxyl-succinimidyl (NHS) followed by immobilization of a molecule of interest through noncovalent complexation, often it may lead to nonspecific functionalization of cell surface molecules. To overcome this limitation, Dutta et al. [34] have developed an alternative method where low-molecular-weight amphiphile with a desired functional group was noncovalently fused with the cell membrane; subsequently, they have covalently conjugated those immobilized functional groups selectively with complementary reactive groups (Figure 5.7). In this approach, they have eliminated the nonspecific functionalization of biomolecules that are present on the cell surface.

image
Figure 5.7 (A) Formation of oxyamine-based liposomes followed by dye functionalization. Subsequently, liposomes were noncovalently fused with cells to immobilize dye-based lipids on the cell surface. (B) Ketone-based lipids were immobilized on the cell surface followed by covalent attachment of dyes on ketones selectively. (C) Schematics and fluorescent micrographs of rewired cells adhered to pattern on self-assembled monolayers (SAMs) of alkanethiolates on gold substrates. Source: Reprinted with permission from Ref. [34] © ACS 2011.

Dutta et al. [34] have presented a method to tether chemoselective ketone and oxyamine groups from mouse fibroblast surfaces by liposome fusion toward the goal of rewiring the cell surface. The modified cells were patterned on solid surfaces using bioorthogonal oxime bond formation and without the use of any cell-adhesive ligand (Figure 5.7) and could be quantified using flow cytometry (Figure 5.8). The ketone and oxyamine groups act as synthetic cell surface receptors and can immobilize a wide range of ligands and probes for subsequent use in cell targeting and tracking. Thus, rewiring cell adhesion was achieved using chemoselective oxime bond formation and without using any cell-adhesive ligands such as Arg-Gly-Asp (RGD) and fibronectin. Through avoiding the use of any additional biomolecules in this strategy, one can eliminate the potential long-term stability and degradation limitation in complex cell culture media or cell targeting in vivo. This methodology also bypasses any metabolic or genetic approach to modify the cell surface, which holds the risk of perturbing any cellular processes by altering cellular physiology.

image
Figure 5.8 (A) Immobilization of dye through liposome fusion with cell membrane. (B) and (C) Cell surface molecule quantification using flow cytometry. Source: Reprinted with permission from Ref. [34] © ACS 2011.

5.2.4 Programmed Cell–Substrate and Cell–Cell Assembly

Cellular communication in a three-dimensional (3D) environment often requires physical contact between neighboring cells that is mediated by the surrounding extracellular matrix [35]. Efforts to mimic the 3D cellular environment and assembly have been mostly limited to using artificial solid scaffolds as platforms to seed cells and manipulating them by external forces and tools [3639]. Recently, chemists and material scientists have begun to develop minimal invasive ways to engineer cell surfaces by biorecognition moieties that act as synthetic cell surface receptors to control cell–cell and cell–substrate interactions. Several approaches, including the use of dielectric forces [4042], laser-guided writing [43,44], surface manipulation [45], and a number of lithographic printing techniques [11,46,47], have been integrated with 3D scaffold designs to produce multitype cellular arrays or 3D cell clusters or spheroids.

A DNA-based approach has been used for the programmable assembly of 3D cellular structure, which may find applications in synthesizing artificial tissues [38]. Recently, Gartner and Bertozzi [38] have demonstrated the assembly of DNA-modified cells into microtissues in a well-defined manner (i.e., structures were controlled by varying the DNA site density on the cell surface, cell concentration, and DNA sequence) as demonstrated in Figures 5.9 and 5.10. Briefly, by adjusting the stoichiometry (i.e., 1:50) of two Jurkat cell populations that were modified with complementary DNA strands, single cell clusters were produced where the limiting cell type was surrounded by cells added in excess. DNA strands were introduced to the cell surface by Staudinger ligation between phosphine-modified single-stranded DNA (ssDNA) and azide-modified cell surfaces where the azide functionality was introduced by metabolic engineering of peracetylated N-α-azidoacetylmannosamine (Ac4ManNAz) and subsequent display of the corresponding sialic acid in the cell membrane. Interestingly, using fluorescently labeled DNA, they showed that DNA strands are clustered at the cell–cell interface, confirming their specific role in cell–cell assembly. Significantly, such assembled cell clusters could be isolated and purified using flow cytometry and subsequently used as the core for further iterative cell assembly (i.e., adding another layer of DNA-modified cells).

image
Figure 5.9 Oligonucleotides direct the synthesis of 3D multicellular structures with defined patterns of connectivity. (A) Cells bearing complementary cell surface oligonucleotides react to form stable cell–cell contacts. (B) Nonadherent Jurkat cells with CMFDA (fluorescein, green) or CMTPX (Texas Red, red) cytosolic stains were combined at a 1:1 ratio. (C) Green- and red-stained Jurkat cells labeled with the mismatched oligos 1-Phos and 5′-Phos were also mixed at a 1:1 ratio. (D) Green- and red-stained Jurkat cells labeled with the complementary sequences 1-Phos and 1′-Phos and combined at a 1:50 ratio. (E) Discrete structures from (D) at higher magnification. (F) Cells labeled with fluorescein-conjugated DNA (single cell, top right) were assembled with cells bearing a nonfluorescent complementary strand (single cell, top left). (G) 3D reconstruction of a single multicellular structure encapsulated in agarose by using deconvolution fluorescence microscopy. Red, Texas Red; green, fluorescein; blue, DAPI. (Scale bars: (B)–(D), 50 μm; (E)–(G), 10 μm.). Source: Reprinted with permission from Ref. [38] © PNAS 2009.
image
Figure 5.10 Fluorescence-activated cell-sorting (FACS) separates cell assemblies based on their fluorescence properties. (A) Schematic diagram of multistep assembly of cellular building blocks. (B) Jurkat cell-based structures were isolated by FACS based on red and green fluorescence proportional to the number of each cell type in the cluster. (C) Fluorescence micrograph of the isolated structures. (D) Higher order multicellular structure generated by incubation of the structures in (C) with a third population of unstained Jurkat cells bearing cell surface oligonucleotides complementary to those on the red-stained cells. Red, Texas Red; green, fluorescein. (Scale bars: 10 μm.). Source: Reprinted with permission from Ref. [38] © PNAS 2009.

More importantly, they have demonstrated that cells within the assembled microtissues communicate via a paracrine mechanism, which is a prerequisite for the development of therapeutically functional tissues. Specifically, in a proof-of-concept model, a CHO cell line engineered to express growth factor interleukin-3 (IL-3) was used as a first building block and the second was an untransformed hematopoietic progenitor cell line (FL5.12) whose survival and replication depend on the presence of IL-3. Cell composites were assembled using DNA hybridization at the interface and then embedded within a 3D agarose matrix. After 16 h, the cell structures underwent accelerated growth due to the presence of IL-3. By contrast, when CHO cells lacking the gene that codes IL-3 were used in the microtissue, FL5.12 cells showed no growth and instead underwent a morphological change corresponding to the absence of IL-3-induced apoptosis. Additional advantages of using DNA assembly in such systems include their reversibility and versatility with respect to engineering multiple orthogonal cell–cell interactions. In addition to ssDNA, self-assembled cargo-carrying DNA arrays can also be attached to the surface of cells. Reiche and colleagues attached hexagonal DNA arrays on cells using streptavidin and antibody as bridges [48]. In their study, it was observed that one or two DNA array patches often attached to one cell and that single DNA arrays often to bridged cells. Occasionally, micron-sized DNA arrays seemed capable of binding to and enveloping multiple cells into larger cellular structures. These larger structures may have the potential to form the foundation for the construction of tissues or organs for transplantation in tissue engineering.

Engineered cell–cell interactions in a bottom-up tissue engineering approach can also be achieved by presenting bioorthogonal functional groups on the cell surface. Dutta et al. [49] have developed a simple liposome fusion method to display ketone or oxyamine functional groups from cell surfaces to generate 3D spheroid assemblies and multilayered tissue-like structures (Figures 5.11 and 5.12). Tissue-like networks were also formed on the surface with geometric control. Surfaces were patterned with cell-adhesive and nonadhesive regions to generate multilayered sheets and patterned tissue structures, respectively. Ketone- and oxyamine-tailored liposomes were cultured with separate fibroblast populations, resulting in membrane fusion and subsequent presentation of chemoselective sites for oxime conjugation from the surface. Culturing these groups on a solid support (~105 cells/ml) and in a layer-by-layer deposition manner gave rise to multilayered, tissue-like cell sheets, which were characterized by confocal microscopy. Cells cultured without the ketone and oxyamine groups did not form any multilayered structures. Spatial control on tissue formation was demonstrated using a micro-contact printing method where a variety of patterns and geometries were produced on a gold substrate. Employing SAMs and microfabrication techniques, hexadecanethiol (1 mM in ethanol) was printed on a gold surface. The surface was then backfilled with tetraethylene glycol (1 mM in ethanol, 16 h) to render the remaining regions inert to nonspecific protein absorption. Fibronectin, a cell-adhesive protein, was then added (10 mg/ml in calf bovine serum (CBS) 2 h), adhering only to the hydrophobic, patterned areas. When liposome fusion occurs to display complementary ketone and oxyamine groups from cell surfaces, multilayered 3D cell patterns were formed. The assembled tissue-like structures were generated in different shapes such as circular, bar, and squares. The strategy was also extended to form 3D cocultures of hMSCs and fibroblasts. The 3D cocultures were generated with two different cell types, in this instance, hMSCs and fibroblasts. Using this method, they retained individual cell phenotype properties that were demonstrated through differentiating hMSCs into adipocytes [49].

image
Figure 5.11 Fluorescent, phase contrast, and scanning electron micrographs (SEMs) describing 3D spheroid formation via liposome fusion and chemoselective cell surface tailoring. (A) Two fibroblast populations were cultured separately with ketone- or oxyamine-containing liposomes. Upon mixing these cell populations, clustering and tissue-like formation, based on chemoselective oxime conjugation, occurred. (B) Control experiments (overlay image) demonstrate no spheroid formation for cells that did not contain either ketone or oxyamine groups. (C) and (D) Two cell populations with chemoselective recognition groups were mixed to form interconnected spheroidal assemblies (overlay images). (E)–(G) Representative SEM images. Source: Reprinted with permission from Ref. [49] © ACS 2011.
image
Figure 5.12 Schematic and images of oxime-mediated, 3D tissue-like structure formation with controlled interconnectivity. (A) Ketone- and oxyamine-containing liposomes were added to two separate fb populations, resulting in membrane fusion. Culturing these cells on substrates, alternating cell population seeding layer-by-layer, gave rise to multilayered, tissue-like cell sheets through stable oxime chemistry. (B) A 3D reconstruction and (C) confocal micrograph showing only a monolayer of cells after oxyamine-presenting cells were cultured with adhered nonfunctionalized cells. (E) A 3D reconstruction and (F) confocal micrograph of multiple cell layers. (D) and (G) Brightfield and fluorescent images of deadhered cell sheets from the surface. Cells were stained with DAPI (blue, nucleus) and phalloidin (red, actin). Source: Reprinted with permission from Ref. [49] © ACS 2011.

5.2.5 Anchoring Biomolecules on the Cell Surface Through Insertion of Hydrophobic Chain into the Cell Membrane

Although hydrophobic interactions are inherently weak in nature (in the order of 2–5 kBT in vacuum), they play a prominent role in many biological processes like lipid bilayer formation, protein folding, and protein–protein recognition [50]. In a seminal work by Dennis and coworkers [51] using a lipid bilayer system, MSCs were treated with palmitated protein G where the palmitate chain was incorporated into the lipid bilayer via hydrophobic interactions and protein G provides generic binding sites for antibodies (Figure 5.13). In a proof-of-concept work, intercellular cell adhesion molecule-1 (ICAM-1) was conjugated onto MSCs, which enabled the cells to bind to ICAM-1, a critical adhesion molecule expressed on activated endothelium. These chemical approaches to tailor cell surfaces with functional ligands are broad and generic and are not limited to only MSCs and the ligands mentioned herein.

image
Figure 5.13 Diagrammatic representation of endothelial cell targeting using antibody-coated MSCs for tissue regeneration. (A) A schematic representation of the membrane structure of AbICAM-1–MSC. (B) Targeting activated ECs expressing ICAM-1 abundantly in ischemic sites with AbICAM-1–MSCs. Source: Reprinted with permission from Ref. [51] © Elsevier 2010.

Cell surface hydrophobicity and cell surface charge have been recognized as measurable physicochemical variables for evaluating cellular adhesion to surfaces. Amphiphilic polymers, such as polyethylene glycol (PEG) conjugated phospholipids (PEG-lipid) and poly(vinyl alcohol) bearing hydrophobic alkyl side chains (PVA-alkyl), have been used in cell surface modification [5256]. In these methods, the hydrophobic alkyl chains of amphiphilic polymers are spontaneously anchored into the lipid bilayer membranes through hydrophobic interactions (Figure 5.14). These spontaneous insertions can be monitored by a surface plasmon resonance (SPR) when a PEG-lipid solution is applied to supported lipid membranes that are formed on an SPR sensor. The spontaneous incorporation is greatly affected by the length of phospholipid alkyl chain length.

image
Figure 5.14 Schematic illustration of immobilization of urokinase on the islet surface. SH-PVA-alkyl was introduced to the cell surface by the hydrophobic interaction between alkyl side chains and the lipid bilayer of the cell membrane. Then, urokinase–Mal was immobilized on the cell surface via thiol–maleimide bonding. Source: Reprinted with permission from Ref. [56] © Elsevier 2008.

The other method is the formation of an intermediary polymer layer through hydrophobic interactions using amphiphilic polymers, such as PEG-lipid and PVA-alkyl, that bear a wide range of functional groups. The functional groups that are introduced on the cell surface through amphiphilic polymers can be used for the immobilization of bioactive substances (Figure 5.14). In one example, PVA-alkyl presenting SH groups were used for the immobilization of bioactive substances displaying maleimide groups. In a similar manner, PEG-lipids presenting maleimide at the end of the PEG chain was also used for the immobilization of bioactive substances tethering SH groups without cytotoxicity.

By utilizing this approach, Totani et al. have developed a promising system that is urokinase-immobilized islets for the potential treatment of diabetes. Although transplantation of islets of Langerhans is a promising method to treat insulin-dependent type I diabetes, innate immunological response activated by blood coagulation plays a key role in the loss of islets at the beginning. Thus, to overcome this limitation and to inhibit blood coagulation on the islet surface immobilization, a plasminogen activator, urokinase, was immobilized on the islet surface via a PVA derivative with alkyl chains and thiol groups (Figure 5.14). When the PVA derivative was added to an islet suspension, the alkyl side chains spontaneously anchored into the lipid bilayer membranes of islet cells. Subsequently, urokinase modified with maleimide groups was immobilized onto the islet surface by thiol/maleimide covalent bonding with the PVA layer. Urokinase-immobilized islets exhibited fibrinolytic properties, indicating that blood coagulation can be controlled on the islet surface. Furthermore, fibrinolysis activity of surface-immobilized urokinase was determined by the fibrin-plate-based assay (Figure 5.15). Urokinase activates plasminogen in the fibrin gel to form plasmin. Plasmin triggers a proteolytic cascade that results in fibrin gel degradation, which forms a transparent window in the gel plate. A large transparent window (2.5 cm2) was observed around the spotted urokinase-islets, which reflected the fibrinolytic activity of urokinase immobilized on the islets. This is an elegant example of versatility and efficiency of noncovalent immobilization of biomolecules on the cell surface.

image
Figure 5.15 Fibrin plate assay of urokinase-immobilized islets. Urokinase-immobilized islets and nontreated islets (100 islets, respectively) were placed on a fibrin gel plate and incubated at 37°C for 13 h. Plasmin that is produced from plasminogen in the presence of urokinase-immobilized islets, dissolves larger area of the fibrin gel (larger black circle). Source: Reprinted with permission from Ref. [56] © Elsevier 2008.

5.3 Electrostatic Interactions Mediated Cell Surface Modification

5.3.1 Engineering of Cell Surface by Harnessing Electrostatic Interactions

The cell membrane is intrinsically negatively charged due to a high composition of negatively charged lipids such as phosphatidylserine and phosphatidylinositol lipids. The possibility of constructing a thin polymer membrane on the surface of cells through electrostatic interactions between negatively charged cell surfaces and cationic polymers and then further modifying them using a layer-by-layer technique of anionic and cationic polymers [5761] has been explored by several research groups. The ionic polymers that have been employed are poly(allylamine hydrochloride), poly(styrene sulfate), poly-L-lysine, and poly-(ethyleneimine). A layer-by-layer method is a simple and attractive technique to modify cell surfaces (Figure 5.16) [62]. The outermost layer of the polymer controls the surface properties. The thickness of the layer is also controllable by the number of polymer layers. However, most polycations, such as poly-L-lysine and poly-ethylene-imine, can be extremely cytotoxic and severely damage the treated cells. Although cationic polymers are efficient to interact with the cell membrane, often they rupture the cell membrane through electrostatic interactions. This leads to severe cell toxicity. To avoid cationic-polymer-associated cytotoxicity, attempts have been made to minimize electrostatic interactions between cationic polymer and cell surface through modulation of charge on the polymer. However, such weak electrostatic interactions often do not yield desirable effects due to reduction of residence time of polymer on the cell surface. Thus, it is extremely important to carefully design the polymers to achieve a balance between toxicity and efficiency.

image
Figure 5.16 (A) Schematic illustrating MSCs encapsulation process using layer-by-layer assembly. Deposition of cationic PLL layer followed by deposition of anionic HA layer, and process was repeated multiple times. Confocal micrographs of MSCs encapsulated within PLL/HA at (B) 3 days postencapsulation and (C) 7 days postencapsulation. Scale bars represent 8.57 and 117 μm for panels B and C, respectively. Source: Reprinted with permission from Ref. [62] © Wiley-VCH Verlag GMBH & Co. 2007.

5.3.2 Noncovalent Cell Surface Modification for Bioimaging Applications

Cell surface modification can facilitate the characterization of a number of cellular processes and signal transduction pathways in terms of identifying key proteins and signaling molecules [63]. One major application of engineering cell surfaces can be in modifying the cell membrane with molecular and detection probes for subsequent tracking during postinfusion in cell therapy [64]. Chemical reporters have been incorporated to the cell membrane using the metabolic pathways [14]. Orynbayeva et al. [65] engineered chromatic polydiacetylene(PDA) polymer patches onto the cell surfaces for imaging structural perturbations of the membrane bilayer (Figure 5.17). Lipid vesicles comprising of 1,2-dimyristoyl-sn-glycero-phosphoethanolamine and 1,2-dimyristoyl-sn-glycero-3-[phospho-rac-(1-glycerol)] were fused with the cell membrane, and PDA was formed in situ upon ultraviolet irradiation. The tethered nanopatches of PDA on the cell surface had an intense blue color. The structural perturbations of the membrane bilayer due to the conformational changes in the conjugated (ene-yne) polymer backbone led to a blue-to-red transition of the conjugated PDA. During this blue-to-red color transition, the initial nonfluorescent PDA also emitted fluorescence at 560 and 640 nm. The PDA nanopatches can therefore be used to visualize the cell bilayer perturbations induced by, for example, lidocaine, polymyxin-B, cholesterol, and oleic acid and to screen toxic pesticides and other environmentally toxic small molecules.

image
Figure 5.17 (A) Cell surface engineering with PLL-g-PEG copolymers. (B) Photographs of tubes containing PDA-labeled cells after sedimentation. Left tube: Blue cells before treatment; right tube: cells after incubation with oleic acid (180 μM). (C) and (D) Confocal microscopy fluorescence images (left) and transmission differential interference contrast microscopy images (right) of PDA-labeled U937 cells. (C) and (D) Cells before and after polymyxin-B (50 μM) treatment, respectively. Source: Reprinted with permission from Refs. [65,66] © AAAS 2005 and © Wiley-VCH 2005.

In another study, Wilson et al. [66] have used cationic graft copolymers to noncovalently engineer pancreatic islet surfaces using electrostatic interactions for subsequent labeling and visualization (Figure 5.17). Specifically, poly(L-lysine)-graft-poly(ethylene glycol) (PLL-g-PEG) copolymers where the polymer backbone was terminally functionalized with biotin, hydrazide, and azide moieties were used to reengineer the cell surfaces with streptavidin-, aldehyde-, and cyclooctyne-labeled probes.

5.3.3 Noncovalent Cell Surface Engineering to Manipulate Cell Biological Fate

Externally applied chemical moieties and materials can be engineered to cell surfaces to control cell biology. The idea is to stimulate and reversibly control different cell signaling pathways by physical tools that are normally biochemically regulated. This can be achieved by linking different molecular probes to the cell surface. Often native cell surface is not amenable to external stimuli manipulation. In such instances, transiently engineering the cell surface by immobilizing an additional agent on the cell surface would enable to manipulate the cell surface and cell fate through an external stimulus. Ingber and coworkers [67] have demonstrated the use of magnetic nanoparticles on the cell membrane to regulate cellular signaling pathways (Figure 5.18). Specifically, mast cells express membrane high-affinity IgE receptors (FcεRI) that can bind to the Fc region of IgE molecules. Multivalent antigens like dinitrophenyl (DNP) that bind to IgE can then induce FcεRI receptor oligomerization on the cell membrane whereas monovalent antigens will fail to do so. The FcεRI receptor clustering quickly triggers an intracellular signaling response characterized by a rapid rise of cytosolic Ca2+, which in turn leads to the local inflammatory response of mast cells. Instead of using an antigen ligand to induce such signaling pathway, the authors immobilized DNP-attached magnetic nanoparticles on IgE-tethered mast cells [68]. The authors ensured that only one DNP molecule was attached on each particle so that receptor clustering did not occur. When an external magnetic field is applied, magnetic nanoparticles clustered on the cell membrane, which caused the oligomerization of FcεRI receptors, a process that could be visualized by the local Ca2+ concentration change. Such magnetic nanoparticles mediated cellular signaling is reversible, switching on and off the magnetic field can directly correspond to a local Ca2+ concentration oscillation. This is an elegant example for modulation of intracellular properties through controlling an altering cell surface by an external stimulus.

image
Figure 5.18 (A) The biochemical mechanism that stimulates downstream signaling (left) involves the binding of multivalent ligands (green hexagons) that induce oligomerization of individual IgE/FcεRI receptor complexes. In the magnetic switch, monovalent ligand-coated magnetic nanobeads (dark grey circles) bind individual IgE/FcεRI receptor complexes without inducing receptor clustering (centre). (B) The pseudocolored microfluorimetric image shows the local induction of calcium signaling (yellow) in cells near the tip of the electromagnet within 20 s of the field being applied. Scale bar represents 50 μm. (C) Quantification of peak changes in intracellular calcium relative to time 0 measured within individual cells during a 1-min pulse of applied magnetic force as a function of the distance of the tip from the cell surface. (D) Effect of a rapid cyclical magnetic stimulation regimen (40 s on, 20 s off) on intracellular calcium signaling. Source: Reprinted with permission from Ref. [67] © NPG 2007.

5.4 Advantages and Limitations of the Noncovalent Modification of the Cell Surface

5.4.1 Advantages of Noncovalent Modification of the Cell Surface

Cell surface engineering has primarily been a subject of molecular biology and genetic engineering. Although genetic manipulation is an efficient method to modulate the cell surface, potentially it might disrupt the inherent cell properties. Often it might lead to unforeseen and unwanted effects. Thus, selective engineering of cell surface without the need of genetic manipulation is a welcoming approach to modulate the cell surface features and properties. Typically, NHS-activated esters and cyanuric chloride derivatives are frequently used to form covalent bonds to the cell surface functional groups such as amines. In addition, maliemide derivatives can be used to react with thiol groups of membrane molecules. However, although covalent modification provides a robust functionalization approach, degree of functionalization needs to be highly optimized to avoid deactivation of membrane molecules such as proteins, lipids, and receptors. Thus, potential toxic effects may be exerted on membrane biomolecules [68]. Enzymatic treatment and metabolic introduction via sialic acid biosynthetic pathway (Bertozzi group) have been employed to add multiple functional groups such as biotin, azido, and ketone groups to living cell surfaces [14,15]. However, these technologies are limited to the introduction of specified small molecules to cells and may perturb cell physiology. Thus, noncovalent methods to transiently engineer the cell surface could overcome these limitations. Liposome fusion can serve as an important tool among many other techniques that have been previously described to modify and tailor cell surfaces with biologically relevant molecules, significantly contributing to tissue engineering fields in generating 3D tissue constructs and in the field of cell-based targeted therapy.

5.4.2 Limitations of Noncovalent Modification of the Cell Surface

Although noncovalent modification of cell surfaces does not perturb the cellular machinery of the cell, the noncovalent forces like hydrophobic and electrostatic interactions holding the molecular probes to the cell membrane are intrinsically weaker in nature. This can result in a comparatively unstable system leading to temporal sensitivity. While this approach is optimal for transient surface medication, engineering surface through noncovalent modifications might not be retained in the longer term. Thus, an extensive optimization needs to be done to increase the longevity of the engineered cell surfaces.

5.5 Conclusions and Future Perspectives

Noncovalent methods for cell surface engineering provide efficient and facile approaches toward modulating cellular functions and finding solutions to fundamental biological problems. It has potential uses in a wide range of fields of biomedicine, tissue engineering, regenerative medicine, wound healing, cell-based therapies, imaging, and diagnostics. In future, coupling multiple noncovalent forces in a single system that would enhance the longevity of the system could enable the development of more elegant and robust methods of engineering the cell surfaces with different molecular probes. Efforts should also be made in designing dynamic systems, where molecules of choice could be conjugated and released from the cell surface using external stimuli. The eventual design and manufacture of these systems would need further collaborations between biologists and engineers for such products to make it to the marketplace.

Acknowledgments

PKV thanks Department of Biotechnology, Government of India for Ramalingaswami Re-Entry Fellowship, and Institute for Stem Cell Biology and Regenerative Medicine (inStem), Bangalore for financial support.

References

1. Gullberg D, Ekblom P. Extracellular matrix and its receptors during development. Int J Dev Biol. 1995;39:845–854.

2. Scott MD, Murad KL, Koumpouras F, Talbot M, Eaton JW. Chemical camouflage of antigenic determinants: stealth erythrocytes. Proc Natl Acad Sci USA. 1997;94:7566–7571.

3. Huebsch N, Mooney DJ. Inspiration and application in the evolution of biomaterials. Nature. 2009;462:426–432.

4. Zhao W, Karp JM. Controlling cell fate in vivo. Chembiochem. 2009;10:2308–2310.

5. Anker JN, Hall WP, Lyandres O, Shah NC, Zhao J, Duyne PV. Biosensing with plasmonic nanosensors. Nat Mater. 2008;7:442–453.

6. Zhao W, Karp JM. Tumour targeting: nanoantennas heat up. Nat Mater. 2009;8:453–454.

7. Giepmans BNG, Adams SR, Ellisman MH, Tsien RY. The fluorescent toolbox for assessing protein location and function. Science. 2006;312:217–224.

8. Schmidt U, Günther C, Rudolph R, Böhm G. Protein and peptide delivery via engineered polyomavirus-like particles. FASEB J. 2001;15:1646–1648.

9. Akin D, Ragheb K, Sturgis J, Sherman D, Burkholder K, et al. Bacteria-mediated delivery of nanoparticles and cargo into cells. Nat Nanotechnol. 2007;2:441–449.

10. Chen S, Smith LM. Photopatterned thiol surfaces for biomolecule immobilization. Langmuir. 2009;25:12275–12282.

11. Falconnet D, Csucs G, Grandin HM, Textor M. Surface engineering approaches to micropattern surfaces for cell-based assays. Biomaterials. 2006;27:3044–3063.

12. Balyasnikova IV, Franco-Gou R, Mathis JM, Lesniak MS. Genetic modification of mesenchymal stem cells to express a single-chain antibody against EGFRvIII on the cell surface. J Tissue Eng Regen Med. 2010;4:247–258.

13. Saxon E. Cell surface engineering by a modified Staudinger reaction. Science. 2000;287:2007–2010.

14. Prescher JA, Bertozzi CR. Chemistry in living systems. Nat Chem Biol. 2005;1:13–21.

15. Rabuka D, Forstner MB, Groves JT, Bertozzi CR. Noncovalent cell surface engineering: incorporation of bioactive synthetic glycopolymers into cellular membranes. J Am Chem Soc. 2008;130:5947–5953.

16. Paulick MG, Forstner MB, Groves JT, Bertozzi CR. A chemical approach to unraveling the biological function of the glycosylphosphatidylinositol anchor. Proc Natl Acad Sci USA. 2007;104:20332–20337.

17. Yun Lee D, Hee Nam J, Byun Y. Functional and histological evaluation of transplanted pancreatic islets immunoprotected by PEGylation and cyclosporine for 1 year. Biomaterials. 2007;28:1957–1966.

18. Lee DY, Lee S, Nam JH, Byun Y. Minimization of immunosuppressive therapy after islet transplantation: combined action of heme oxygenase-1 and PEGylation to islet. Am J Transplant. 2006;6:1820–1828.

19. Cabric S, Sanchez J, Lundgren T, et al. Islet surface heparinization prevents the instant blood-mediated inflammatory reaction in islet transplantation. Diabetes. 2007;56:2008–2015.

20. Stabler CL, Sun X-L, Cui W, et al. Surface re-engineering of pancreatic islets with recombinant azido-thrombomodulin. Bioconjug Chem. 2007;18:1713–1715.

21. Kadowaki K, Matsusaki M, Akashi M. Control of cell surface and functions by layer-by-layer nanofilms. Langmuir. 2010;26:5670–5678.

22. Belardi B, O’Donoghue GP, Smith AW, Groves JT, Bertozzi CR. Investigating cell surface galectin-mediated cross-linking on glycoengineered cells. J Am Chem Soc. 2012;134:9549–9552.

23. Hangauer MJ, Bertozzi CR. A FRET-based fluorogenic phosphine for live-cell imaging with the Staudinger ligation. Angew Chem Int Ed. 2008;47:2394–2397.

24. Laughlin ST, Bertozzi CR. Metabolic labeling of glycans with azido sugars and subsequent glycan-profiling and visualization via Staudinger ligation. Nat Protoc. 2007;2:2930–2944.

25. Karp JM, Leng Teo GS. Mesenchymal stem cell homing: the devil is in the details. Cell Stem Cell. 2009;4:206–216.

26. Rombouts WJC, Ploemacher RE. Primary murine MSC show highly efficient homing to the bone marrow but lose homing ability following culture. Leukemia. 2003;17:160–170.

27. Kumar S, Nagy TR, Ponnazhagan S. Therapeutic potential of genetically modified adult stem cells for osteopenia. Gene Ther. 2010;17:105–116.

28. Brenner S, Whiting-Theobald N, Kawai T, et al. CXCR4-transgene expression significantly improves marrow engraftment of cultured hematopoietic stem cells. Stem Cells. 2004;22:1128–1133.

29. Zhang D, Fan G-C, Zhou X, et al. Over-expression of CXCR4 on mesenchymal stem cells augments myoangiogenesis in the infarcted myocardium. J Mol Cell Cardiol. 2008;44:281–292.

30. Sackstein R, Merzaban JS, Cain DW, et al. Ex vivo glycan engineering of CD44 programs human multipotent mesenchymal stromal cell trafficking to bone. Nat Med. 2008;14:181–187.

31. Chavakis E, Urbich C, Dimmeler S. Homing and engraftment of progenitor cells: a prerequisite for cell therapy. J Mol Cell Cardiol. 2008;45:514–522.

32. Sarkar D, Vemula PK, Teo GSL, et al. Chemical engineering of mesenchymal stem cells to induce a cell rolling response. Bioconjug Chem. 2008;19:2105–2109.

33. Sarkar D, Vemula PK, Zhao W, Gupta A, Karnik R, Karp JM. Engineered mesenchymal stem cells with self-assembled vesicles for systemic cell targeting. Biomaterials. 2010;31:5266–5274.

34. Dutta D, Pulsipher A, Luo W, Mak H, Yousaf MN. Engineering cell surfaces via liposome fusion. Bioconjug Chem. 2011;22:2423–2433.

35. Nelson CM, Bissell MJ. Of extracellular matrix, scaffolds, and signaling: tissue architecture regulates development, homeostasis, and cancer. Annu Rev Cell Dev Biol. 2006;22:287–309.

36. Gillette BM, Jensen JA, Tang B, et al. In situ collagen assembly for integrating microfabricated three-dimensional cell-seeded matrices. Nat Mater. 2008;7:636–640.

37. Tanaka H, Murphy CL, Murphy C, Kimura M, Kawai S, Polak JM. Chondrogenic differentiation of murine embryonic stem cells: effects of culture conditions and dexamethasone. J Cell Biochem. 2004;93:454–462.

38. Gartner ZJ, Bertozzi CR. Programmed assembly of 3-dimensional microtissues with defined cellular connectivity. Proc Natl Acad Sci USA. 2009;106:4606–4610.

39. Albrecht DR, Underhill GH, Wassermann TB, Sah RL, Bhatia SN. Probing the role of multicellular organization in three-dimensional microenvironments. Nat Methods. 2006;3:369–375.

40. Gray DS, Tan JL, Voldman J, Chen CS. Dielectrophoretic registration of living cells to a microelectrode array. Biosens Bioelectron. 2004;19:771–780.

41. Odde DJ, Renn MJ. Laser-guided direct writing of living cells. Biotechnol Bioeng. 2000;67:312–318.

42. Nahmias Y, Odde DJ. Micropatterning of living cells by laser-guided direct writing: application to fabrication of hepatic-endothelial sinusoid-like structures. Nat Protoc. 2006;1:2288–2296.

43. Barron JA, Krizman DB, Ringeisen BR. Laser printing of single cells: statistical analysis, cell viability, and stress. Ann Biomed Eng. 2005;33:121–130.

44. Inaba R, Khademhosseini A, Suzuki H, Fukuda J. Electrochemical desorption of self-assembled monolayers for engineering cellular tissues. Biomaterials. 2009;30:3573–3579.

45. Ringeisen BR, Othon CM, Barron JA, Young D, Spargo BJ. Jet-based methods to print living cells. Biotechnol J. 2006;1:930–948.

46. Chiou PY, Ohta AT, Wu MC. Massively parallel manipulation of single cells and microparticles using optical images. Nature. 2005;436:370–372.

47. Khademhosseini A, Langer R, Borenstein J, Vacanti JP. Microscale technologies for tissue engineering and biology. Proc Natl Acad Sci USA. 2006;103:2480–2487.

48. Koyfman AY, Braun GB, Reich NO, Yousaf MN. Cell-targeted self-assembled DNA nanostructures. J Am Chem Soc. 2009;131:14237–14239.

49. Dutta D, Pulsipher A, Luo W, Yousaf MN. Synthetic chemoselective rewiring of cell surfaces: generation of three-dimensional tissue structures. J Am Chem Soc. 2011;133:8704–8713.

50. Young L, Jernigan RL, Covell DG. A role for surface hydrophobicity in protein–protein recognition. Protein Sci. 1994;3:717–729.

51. Ko IK, Kean TJ, Dennis JE. Targeting mesenchymal stem cells to activated endothelial cells. Biomaterials. 2009;30:3702–3710.

52. Teramura Y, Kaneda Y, Iwata H. Islet-encapsulation in ultra-thin layer-by-layer membranes of poly(vinyl alcohol) anchored to poly(ethylene glycol)-lipids in the cell membrane. Biomaterials. 2007;28:4818–4825.

53. Teramura Y, Kaneda Y, Totani T, Iwata H. Behavior of synthetic polymers immobilized on a cell membrane. Biomaterials. 2008;29:1345–1355.

54. Teramura Y, Iwata H. Islet encapsulation with living cells for improvement of biocompatibility. Biomaterials. 2009;30:2270–2275.

55. Miura S, Teramura Y, Iwata H. Encapsulation of islets with ultra-thin polyion complex membrane through poly(ethylene glycol)-phospholipids anchored to cell membrane. Biomaterials. 2006;27:5828–5835.

56. Totani T, Teramura Y, Iwata H. Immobilization of urokinase on the islet surface by amphiphilic poly(vinyl alcohol) that carries alkyl side chains. Biomaterials. 2008;29:2878–2883.

57. Krol S, del Guerra S, Grupillo M, Diaspro A, Gliozzi A, Marchetti P. Multilayer nanoencapsulation New approach for immune protection of human pancreatic islets. Nano Lett. 2006;6:1933–1939.

58. Elbert DL, Herbert CB, Hubbell JA. Thin polymer layers formed by polyelectrolyte multilayer techniques on biological surfaces. Langmuir. 1999;15:5355–5362.

59. Chanana M, Gliozzi A, Diaspro A, et al. Interaction of polyelectrolytes and their composites with living cells. Nano Lett. 2005;5:2605–2612.

60. Germain M, Balaguer P, Nicolas JC, et al. Protection of mammalian cell used in biosensors by coating with a polyelectrolyte shell. Biosens Bioelectron. 2006;21:1566–1573.

61. Veerabadran NG, Goli PL, Stewart-Clark SS, Lvov YM, Mills DK. Nanoencapsulation of stem cells within polyelectrolyte multilayer shells. Macromol Biosci. 2007;7:877–882.

62. Decher G. Fuzzy nanoassemblies: toward layered polymeric multicomposites. Science. 1997;277:1232–1237.

63. Stevens MM, George JH. Exploring and engineering the cell surface interface. Science. 2005;310:1135–1138.

64. Daley GQ, Scadden DT. Prospects for stem cell-based therapy. Cell. 2008;132:544–548.

65. Orynbayeva Z, Kolusheva S, Groysman N, Gavrielov N, Lobel L, Jelinek R. Vaccinia virus interactions with the cell membrane studied by new chromatic vesicle and cell sensor assays. J Virol. 2007;81:1140–1147.

66. Wilson JT, Krishnamurthy VR, Cui W, Qu Z, Chaikof EL. Noncovalent cell surface engineering with cationic graft copolymers. J Am Chem Soc. 2009;131:18228–18229.

67. Mannix RJ, Kumar S, Cassiola F, et al. Nanomagnetic actuation of receptor-mediated signal transduction. Nat Nanotechnol. 2008;3:36–40.

68. Yung CW, Fiering J, Mueller AJ, Ingber DE. Micromagnetic-microfluidic blood cleansing device. Lab Chip. 2009;9:1171–1177.

..................Content has been hidden....................

You can't read the all page of ebook, please click here login for view all page.
Reset
18.226.185.87