Skip to main content

Polymeric biomaterial-inspired cell surface modulation for the development of novel anticancer therapeutics

Abstract

Immune cell-based therapies are a rapidly emerging class of new medicines that directly treat and prevent targeted cancer. However multiple biological barriers impede the activity of live immune cells, and therefore necessitate the use of surface-modified immune cells for cancer prevention. Synthetic and/or natural biomaterials represent the leading approach for immune cell surface modulation. Different types of biomaterials can be applied to cell surface membranes through hydrophobic insertion, layer-by-layer attachment, and covalent conjugations to acquire surface modification in mammalian cells. These biomaterials generate reciprocity to enable cell–cell interactions. In this review, we highlight the different biomaterials (lipidic and polymeric)-based advanced applications for cell–surface modulation, a few cell recognition moieties, and how their interplay in cell–cell interaction. We discuss the cancer-killing efficacy of NK cells, followed by their surface engineering for cancer treatment. Ultimately, this review connects biomaterials and biologically active NK cells that play key roles in cancer immunotherapy applications.

Graphical Abstract

Introduction

Mammalian cell membranes contain a sophisticated heterogeneous environment with various functional moieties of protein, carbohydrates, lipids, and other composites. These functional moieties participate in the identification of adjacent cells through physical contact-mediated direct cell–cell interactions and further dynamic signal transductions [1, 2]. Therefore, the presence of additional functional groups on cell surfaces could effectively manipulate cellular functions. Modification of cell membranes using suitable natural or synthetic biomaterials containing such functional moieties opens a new windows for selective cell–cell interactions [1].

In recent years, many cell surface engineering strategies have been investigated for the development of cutting-edge cell-based therapeutic agents (Fig. 1). These approaches have successfully modulated cell recognition, cell tracking, imaging, and cell-based immunotherapy [2, 3]. For cancer treatments in particular biomaterial-mediated surface modification techniques are undoubtedly clinically effective and robust. Among various tools for cellular surface modification, the anchoring of biomaterials into cell membranes inspires cell–cell and cell–extracellular interactions. Surface presentation of such specific functional moieties on cell membranes has been utilized to augment cellular intrinsic properties and improve surface-mediated cellular signaling.

Fig. 1
figure 1

Scheme representing different approaches for cell surface engineering

Recently, the use of polymer-based materials for effective cell–coating has attracted considerable attention towards developing cell-based therapeutic products. However, the conventional cell–coating processes still present many technical limitations, including interferences in cell division, proliferation, and subsequent disruption in cellular structure [4, 5]. More importantly, undesired coverage of cell receptors/ligands and inhibited secretion of translated protein products by the addition of coating materials on the cell surfaces also downregulate intrinsic cellular functionalities.

Therefore, it is necessary to design and develop such biomaterials with significantly improved cellular surface properties that include: (1) the protection of cell architectures, (2) the preservation of signal releases such as cytokines and growth factors, (3) the stabilization of cell structure after coating, (4) enhanced cell–cell interactions, and (5) suitable degradation or dissociation of coating materials after coated-cell delivery, especially upon in vivo administration.

In this regard, anticancer treatments using surface-modified immune cells have been intensively investigated [6]. For example, chimeric antigen receptors (CAR)-modified T (CAR–T) cell is the most representative immune cell-based immunotherapy formulation for acute lymphoblastic leukemia treatment [7, 8]. Further efforts have focused on improving the anticancer efficacy of CAR–based strategies via increased target specificity and subsequently reduced associated side effects. Facilitated tumor recognition and antigen-specific targeting could extend the therapeutic benefits of CAR-based cellular therapeutics [9]. This is often accomplished by expressing bi-specific/tri-specific recognition moieties that target different antigens in tandem, and bind with single signaling endodomain. Subsequently, a similar surface engineering technique using precise genetic modification has been integrated into NK cell-mediated anticancer immunotherapies as well [10, 11]. NK cells exhibit potential toxicity against various types of solid tumors as self-therapy agents, due to the absence of MHC class I molecules on their surfaces [12, 13]. NK cell can potentially kill the cancer cells through two different methods: (1) direct cancer cell killing through the formation of immunogenic synapse between host and guest cells, and (2) indirect cancer cell cytotoxicity through activating secretory lysozyme containing perforin and granzyme. In addition, many NK cell receptors are also involved in inducing cytotoxicity in cancer cells through selective cancer cell recognition [14]. Cell cytotoxicity predominantly involved through Fas ligand (FasL) and tumor necrosis related apoptosis induced ligand (TRAIL), which presents on both host and guest cell surface, can trigger intracellular caspase activation [15, 16]. The activated NK cells could potentially kill the cancer cells through perforin and granzyme secretion [17,18,19].

Recently efforts have been made to enhance NK cell fitness and antitumor functions, to express an extracellular receptor and functional CD3/TCR signaling. Such studies have demonstrated that, these cells can target different tumors with potential efficiency, specificity and safety profile (Figs. 2a-c). However, these strategies still have some limitations, like complex manufacturing, HLA-restricted killing, and in vivo persistence [20,21,22].

Fig. 2
figure 2

Different strategies enhancing NK cell activity to redirect tumor cell killing efficacy

Motivated by the potential efficacies, encouraging results and their limitations, there is growing interest in the design of surface-engineered NK cells using functional biomaterials for potential cancer immunotherapy (Fig. 2d). Therefore, surface-coated NK cells via multifunctional biomaterials with proximal recognition moieties have become one of the most fascinating research areas. As per the previous reports, the use of non-genetic modulation using these biomaterials could be a safer and easier way to manipulate the composition of cell membranes [23]. Therefore, the incorporation of multifunctional biomaterials onto NK cell surfaces augments tumor–targeting ability and regulates subsequent immune responses through cell-contact biophysical phenomena [24].

Recently, several substantial studies have reported the efficacy of biomaterials on various therapeutic treatments; however, these molecules have hardly been recognized as strong target agents or medications [25]. In a few studies, researchers also found that the multifunctional biomaterials can bind with protein/lipids/small molecules, and make use of these molecules in cell–cell recognition, immune response, and cell signaling. Therefore, a careful selection of biomaterials for cell surface modification, which might be used as a possible therapeutic system, is necessary for efficient NK cell surface engineering for cancer therapy. In this review, we aim to provide an overview of (1) the utilization of prominent cancer recognition moieties, (2) practical synthesis strategies of cell membrane anchors, and (3) the latest advances in the design of multifunctional biomaterials for ex vivo NK cell surface modification.

Cancer recognition

In-depth understanding of the composition and related interaction of cell surface components is necessary to improve target-specific binding between immune cells and tumor cells. By augmenting a variety of synapse interactions, accurate regulation in site-specific binding could be achieved without intervention [26]. To this end, applying additional recognition moieties into immune cells could eradicate rapid transformation of the malignant cell population without affecting normal cells [27]. Specific use of cancer recognition moieties monitored real time therapeutic efficacies of modulated immune cells with personalized preventions. In past years, various recognition moieties, including small molecules, proteins, peptides, aptamer, glycans, and vitamins, have been effectively used to identify and target cancer cells [28].

Recognition moieties could be conjugated with a series of biomaterials for preferential integration into solid tumors. Such conjugated recognition moieties afford potential advantages over conventional vectors, including selective attachment, the avoidance of biological barriers, and enhanced permeation and retention (EPR) effect [29]. Several cancer targeting moieties have been used to facilitate cancer recognition [30,31,32]. In this section, the most potent acid molecule-based cancer recognition moieties, i.e., folic acid, lactobionic acid and boronic acid, are discussed.

Folic acid recognition by folate receptors

Folic acid (FA) is a vitamin B9-based biomolecule and a multicomponent moiety that contains glutamic acid, teropterin, and para-amino benzoic acid. Folic acid (FA), a vitamin B9-based biomolecule, has been most widely used for different cancer targeting, based on its selective binding affinity to folate receptors (FRs) overexpressed in many cancer cells (Fig. 3a). For example, the overexpression of FR–α is generally observed in over 40% of cancer cells, while FR–β is overexpressed in macrophages [33]. Hence, these overexpressed cysteine-rich glycoprotein-based receptors showed high affinity with surrounding the FA and the binding of FA to FRs promotes the clustering of ligand–receptor complex in cellular surfaces, resulting in internalization via endocytosis [34,35,36,37]. At endosomal pH, FA release from receptors and subsequent transport by proton-coupled folate transporters in the cytoplasm occur. Folates also regulate cellular migration, and are associated with tumor progression. It has been reported that the introduction of antifolates in ovarian cancer cells has significantly reduced cell division, independent growth, and the adhesive properties of cancer tissues [35, 38, 39]. Furthermore, FA increases cell proliferation and apoptosis against cancer cells overexpressing FR through JAK–STAT signaling pathway [40].

Fig. 3
figure 3

Scheme representing the role of (a) folic acid, (b) glycan lactobionic acid, and (c) phenyl boronic acid as a cancer recognition moiety for selective anticancer targets

Thus, FA would be considered an interesting cancer cell recognition and targeting ligand, despite its selectivity based on less expression level in normal cells and more efficient internalization to target cancer cells.

Alternatively, designing a FA-based platform to FRs targeting could be a better approach to treat different cancer cells. FA has many advantages over the other targeting moieties: it can easily bind with small molecules present at the cancer cell membrane; it can easily internalize; and it is less overexpressed in normal tissues. The advantages of FA make it a good targeting ligand that selectively binds with cancer cells [41, 42].

However, the FA moiety has been widely used for cancer cell recognition where it also plays a vital role in cell–cell adhesion. Previously, FA was used for cell surface modification to induce cell–cell interaction through FRs at the cancer cell targeted site [35, 43]. In terms of advantages, the FA moiety would be applied for cell surface modification that could enhance the cell–cell interaction. Application of FA conjugated to immunoglobulin (IgG) applied activates NK cells, and selectively binds to FR overexpressed cancer cells [44]. The FA–hepten conjugate is applied to the immune cell to selectively kill cancer cells. It is also reported that the therapeutic efficacy is bimodally dependent on the amount of FA–hepten and showed potential toxicity through the bridge between the cell surface FA and receptors overexpressed in cancer cells [45]. The FA-modified cells would rapidly interact with cancer cells [46]. From another prospective, FA has a few disadvantages, such as a high amount increasing the toxicity toward normal cells and those cells that use them for surface engineering. The high dose of FA also reduced the cytotoxicity of NK cells in mice [47]. Therefore, it is essential to precisely select and design a biomaterial system that can be utilized for NK cell surface modification for cancer immunotherapy.

Glycan lactobionic acid recognition by glycans receptors

Cancer cells have different glycan coatings than normal cells because their glycoproteins and glycolipids are fundamentally altered in cancer cells. Glycan structures are formed by the coordinated action of several enzymes that can catalyze the addition or removal of glycans that are covalently bound to proteins or lipids [48]. The glycans associated with cancer cells, like Lewis’s antigens, sialylated structures, and glycol proteins, are frequently part of the membrane–secreted tumor proteins. For example, asialoglycoprotein receptors (ASGPR) are primarily overexpressed on a variety of tumorous cells, including hepatocellular carcinoma, colon, breast, lung, and prostate cancer cells [49]. These receptors selectively bind with galactosyl residues of the surrounding molecules, and help further the endocytosis process (Fig. 3b). The unique interaction between these galactosyl moieties and cancer cell receptors allows them to be potential cancer targeting ligands [50].

In another way, cancer cells also display the abnormal expression of Galectins (a lectin-based protein) that specifically binds, in addition to aberrant glycan expression. Galectins may be emitted by a wide range of cancer cells, and potentially impaired the function of immune cells like the effector function of T cells, suppressed myeloid cells, and modulated NK cells activity through binding to specific glycans overexpressed on the immune cells [51]. The cancer cell aggression and metastatic correlated with Galectins expression and blockage in tumor microenvironment enhanced the effective function of CD4 + and CD8 + T cells [52]. Additionally, Galectins expression also inhibits the MICA ligand of the NK cell and glycol-dependent interaction [53]. Because NK cell ligands also contain the glycan structure, the binding between NK cell glycan structure and Galectins leads to NK cell evasion [54]. Therefore, the down regulation of Galectins or specific targeting of glycan receptors in solid tumors can lead to tumor suppression and growth inhibition. In another aspect, a strategy that was able to prevent the interaction between inhibitory immune receptors and glycans could serve as better anticancer therapy. The metabolic competition between immune and cancer cell glycans hindered immune cell functions through defective IFN–γ production [55].

Glycol monomer structures such as gluconic acid and lactobionic acid (LBA) act as a Galectin-inhibitors as well as offiring specific binding abilities [56, 57]. Among the monomers, LBA is a sugar molecule also shows the potential inhibitory effects against Galectins, and can also improve tumor immunity in cancer cells [58]. In past years, LBA-based drug delivery systems were specifically utilized for hepatocellular targeting [59,60,61]. It is well known that lectin receptors can easily recognize and endocytose glycol-based drug carriers or sugar-based ligands, especially tumor cells with overexpressed Siglecs receptors through the high binding affinity toward sugar moieties [62].

In the case of NK cells, sugar-mediated modification improved the selective binding ability of NK cells to target cancer cells [63]. Indeed, therapeutic modification of glycol, such as blockage of overexpressed glycans through metabolic memetics, can suppress tumor growth. The tumor suppression can be enhanced presumably due to NK cell activation resulting in decreased Siglecs triggering [48]. In another direction, glycans have also been used for NK cell activation through direct protein glycosylation [64, 65] or biomaterials-mediated glycosylation [66]. However, direct glycosylation onto NK cell surfaces requires the precise identification and selection of well-defined glycans, glycoproteins, and glycopeptides. Therefore, glycol-based biomaterials for NK cell surface engineering has become a rational approach that utilizes glyco-receptor mediated endocytosis [67,68,69,70].

The cell surface with overexpressed sugar moieties could be utilized to modulate cell–cell interaction, transduce signals, and regulate immune response. As compared to normal tissues, aberrant glycosylation overexpression in cancer tissues could be a potential biomarker for cancer cell recognition. Therefore, utilizing LBA as a glycosylation binding agent could be a better therapeutic approach that directly promotes cancer cell suppression.

Glycosylation also affects the immune response of cancer cells. Immune cells are mainly regulated by immunoglobulins and lectins receptors on their surface membranes. Glycosylation of immune cells and cancer cell receptors plays an important role in cancer immunotherapy. For example, in several different forms of cancer, proper ligand–receptor interaction and subsequent activity in the antitumor immunity depend on the glycosylation of PD–1, PD–L1, or B7–H4 [71, 72]. Cancer cells with a high expression of glycoproteins are also susceptible to immune cell cytotoxicity. For example, a high expression of sialyl-based fructosyltransferase 3 (a glycoprotein) onto HepG2 cells resulted in elevated binding affinity with lectin-like receptors of NK cells, and achived downstream NK cell–mediated cytotoxicity [73].

Cancer cells included the expression of O-glycans, N-glycans, sialylated and various fucosylated-based glycans on the cellular surface. Hence, cancer cell overexpressed glycans play a key role in cancer therapy including cancer target, diagnosis and treatment [74]. It is also reported that, the cancer cells expressed glycans facilitate the interaction between cell surface and its surrounding tumor microenvironment through the development of cancer biomarkers like proliferative signaling, resistance to cell death, immune evasion, and angiogenesis. Moreover, cell surface glycoproteins also protect the cell membrane from physical stress and the chemical aspects of cell surface dynamics [75]. Therefore, because of their unique features and their wide range of applications in cancer, glycans have become a promising and fascinating research area in the context of tumor-targeted therapy.

Carbohydrate-based molecules can play important roles in cell–cell interactions. These molecules can self-arrange or be localized near recognized target cells, where they act as a binder with the cancer cell. Saccharides will bind with specific complimentary glycans, or with lectins of target cell. Owing to the specific structure of these molecules, they can modify the immune signals and stability of PD–1.

Glycans such as mannose, galactose and lactobionic acid (LBA) can be used as targeting ligand for hepatocytes, endothelials, and macrophages. Among these carbohydrates, LBA disaccharide molecule, shows good binding affinity with hepatocytes overexpressed ASGPRs receptors through LBA–ASGPR complex formation [59, 76]. The overexpressed ASGPRs promote clathrin-mediated endocytosis during the interaction. ASGPRs mainly consist of two different polypeptide subunits which contain carboxylic groups (–COOH) at the terminal position. Therefore, the hydroxyl (–OH) group of LBA and –COOH of receptors enables the interaction followed by complex formation. Some carbohydrates molecules exhibit cellular and immune response via the enhanced production of proinflammatory modulators [77]. Like other carbohydrates molecules, LBA is also capable of stimulating the immune synapse. Hence, LBA, an effective epitope, could be a promising targeting sugar molecule for effective cancer cell recognition, and treatment via contact with glycoprotein binding.

Boronic acid recognizes sialic acid and carbohydrates on cell surface

Sialic acids (SAs) belong to the sugar family (N–Acetylneuraminic acid) with a nine-carbon backbone structure [78]. The SAs are typically present on several cellular surfaces and secreted glycan molecules on cancer cell surfaces [79]. The aberrant overexpression of SAs on cancer cells can lead to new interplay with immune cells, resulting in the blocking of immune cell activation. The thick coating of SAs on the surface of cancer cells prevents the immune system from eliminating tumors. Siglecs are the sialic acid-binding immunoglobulin-like lectins that can specifically bind to the SAs. Siglecs can be further divided into two sub-categories based on sequence similarity and evolutionary conservation [80]. The first Siglecs group consists of sialoadhesin (Siglec–1), CD22 (Siglec–2), MAG (Siglec–4), and Siglec–15, while the second group consists of CD33 (Siglec–3), Siglec–5, Siglec–6, Siglec–7, Siglec–8, Siglec–9, Siglec–10, Siglec–11, Siglec–14, and Siglec–16 [81]. Most Siglecs serve as activating receptors and are involved in cell–cell interactions, while the majority of Siglecs are responsible for mediating inhibitory signals [82]. Siglecs can effectively bind with target cells via two different interactions: (1) trans-interaction and (2) cis-interaction. The trans-interaction improved binding between effective and target cells, whereas the cis-interaction acted as a masking function on the cell surface, significantly inhibiting SAs binding efficacy. In another direction, abnormal overexpression of SAs on cancer cells could promote cancer growth and proliferation [83], manipulate extracellular matrix interactions, cell–cell interactions, and cell-membrane decoration [84], and effectively overwhelms towards immune cell response and encourage cancer metastasis [85]. Thus, it may also play a role in the molecular mimicry that allows cancer to avoid host cell immune responses.

The hypersialylation (overexpressed SAs) on the cancer cell makes major ligands for the Siglecs binding (present on the immune cell surface). Siglecs and SAs bind to support immunosuppressive signaling, and provide protection to tumor cell. For example, Siglecs–7 and Siglecs–9 present on the NK cell surface bind with SAs on the cancer cell, inhibiting the NK cell toxicity towards cancer cells [86,87,88]. In the tumor environment, sialyl–Tn (sTn) antigens, sialyl Lewis x (sLex) antigen, GD2 disialoganglioside with the sialyl Lewis A epitope (sLea) and sTn epitope, have been used for cancer biomarkers and treatment [89]. Hence, the cancer cell immune response is majorly affected by the Siglecs [90,91,92]. For example, Hudak et al. [86] reported that Siglec–7, an immunoglobulin-like lectin 7 excessively overexpressed in cancer cells, inhibits the cytotoxicity of NK cells by bonding with sialic acid. The immunoreceptor tyrosine-based inhibitory motif (ITIM) found in both Siglecs–7 and Siglecs–9 attracts SHP phosphatases to the region of activation, and prevents the kinase phosphorylation cascade from proceeding. In this study, they also reported that surface–engineered NK cells with increasing ligand density coreceptors would be sufficient to suppress cell killing. Here, a modified synthetic glycopolymer was introduced on the cancer cell surface, and shows that increasing sialylation on cancer cells inhibits the NK cell activation through Siglecs–7 binding. In another sence, Nicoll et al. [93] demonstrated that, Siglec–7 specifically interacts with CD–33 related sialylated probe of cancer cells, and inhibits NK cell cytotoxicity. Here, transfection-based methodology was used to examine the impact of siglec–7–gangliosides GD3 interactions on NK cell killing activity. Siglec-7 is normally concealed on NK cells, but sialidase can make it visible so that it can interact with GD3 on target cells to stop killing. In another way, the expression of GD3 on the cancer cells also increased the siglec–7 mediated killing efficacy.

In recent years, phenylboronic acid (PBA)-based biomaterials have been used to recognize characteristic carbohydrates present on cancer cell surfaces [94,95,96,97,98,99,100], due to the selective binding of PBA with various sialic acids in the tumor microenvironment (Fig. 3c) [101,102,103,104,105,106]. Since this binding ability of boronic acid under different pH conditions was first reported in 1959 [107], a new design strategy has been suggested to develop PBA-decorated biomaterials for cancer treatment [108, 109]. Superior in vitro and in vivo antitumor activities in restricting tumor growth and increasing the survival time of tumor–bearing mice were obtained via PBA–SAs mediated interaction[102, 106, 110]. These studies suggested the potential targetability and pH-triggered drug release profile from PBA–SAs biomaterial. Owing to the ability to form 1,2 and 1,3 diols boronic ester, PBA can also be used in saccharide biosensors applications. Similarly, PBA and SAs can form the reversable borates, which promotes the uptake of cancer cells [111]. Elsewhere, after endocytosis in the tumor cell, the dissociation of the boronic ester complex leads to a decrease in pH level, and an increase in ATP level [105, 112, 113].

Therefore, by introducing PBA on immune cells, it can selectively bind with overexpressed sialyls or siglecs moieties, and can increase the immune cell cytotoxicity toward targeted cells. In other words, PBA would be an effective moiety, and its materials can also be anchor to NK cell membrane; thus, it would also be a promising targeting ligand for selectively targeting cancer cell binding and the treatment of tumor cells.

Synthesis strategies of cell membrane anchors

Today, the ex vivo surface engineering of NK cells by means of modified biomaterials has become a powerful technique for tailoring cell surface propertie, and thereby augmenting the intercellular interactions between NK cells and surrounding target cells. This biomaterial-mediated cell membrane modification could be regulated by the structure of biomaterials, density, and functional groups of biomaterials on NK cell surfaces. Therefore, the design of unique biomaterials without affecting intrinsic cellular properties has become a challenging task. Primary prerequisites for the development of these biomaterials for ex vivo cell surface engineering is the selection of proper anchoring moieties to facilitate membrane binding and presentation onto NK cell membranes. Subsequently, the fabrication of such biomaterials could be initiated by a series of conjugation chemistries including click chemistry, coupling reaction, michael addition, and azide coupling reaction.

Therefore, this section summarizes current investigations using lipid–PEG molecules and polymer conjugates for cell surface modification. In addition, new engineering strategies for designing functional biomaterials specifically for NK cells are discussed.

Lipid–PEG conjugates act as direct anchors for cell membranes vector

In recent years, lipid-PEG derivatives were successfully used for mammalian cell membrane surface modifications [114,115,116]. Lipids can be easily anchored on the cell membrane by hydrophobic interactions with the lipid bilayers of target cells. However, the selection of lipid chain should be preciously determined. As the dynamic and heterogeneous nature of cellular membrane components (i.e., a variety of biological molecules, lipids, and proteins along with intrinsic negative charge and hydrophobic cavity in cell surfaces) can be affected by the lipid insertion [117, 118]. Therefore, charge-based electrostatic interaction and lipid-mediated hydrophobic insertion have been used as major driving forces for the insertion and presentation of biomaterials onto target cellular surfaces. More importantly, the incorporation of such lipid–PEG conjugates on the cell membrane was greatly affected by the length of lipid chains or subsequently modulated hydrophobicity [119, 120]. The incorporation of lipid–PEG on the cell membrane is mainly affected by the length or hydrophobicity of conjugated lipids. Figure 4 describes representative commercially available lipids and PEG derivatives that could be utilized for the design of various cell–coating biomaterials. The anchoring of lipid–based oligonucleotides is a potential tool for cell engineering, and has a wide range of biomedical applications. The specific and unique amphiphilic structure of lipid-based oligonucleotides enables them to self-assemble and potentially monitor cell behavior [121]. Jin et al. synthesized DNA-based phosphorylated lipid (DNA–lipid–P) conjugates for cell membrane anchoring varying lengths of carbon chains (i.e., C6, C9, C12 and C15) in lipid moieties. The anchoring efficiency of synthesized phospholipids onto HepG2 and U–2 OS cells increased when C15 was conjugated. In addition, phosphorylation of this lipid conjugates also influenced the membrane anchoring as compared with non-phosphorylated lipids through alkaline phosphate (ALP)–dependent cell membrane anchoring. With less hydrophobicity, DNA–lipid–P exhibits relatively minimal interactions with the cell membrane in the absence of ALP. Dephosphorylation-induced with increasing hydrophobicity, DNA–lipid–P adheres to cell membranes that were highly ALP-expressing [122].

Fig. 4
figure 4

Commercially available lipids and their PEG-based derivatives can be utilized to develop cell–coating biomaterials

The development of modified lipid–PEG anchors is an easier and more rapid method for various cell membrane anchoring. The terminal moiety of PEG can be further modified to make a biocompatible surface anchor. Kato et al. developed biocompatible oleyl (single and double– tailed) chain-based PEG conjugates for various cells, i.e., NIH3T3, 32D, Ba/F3, hybridoma 9E10 cell–coating [123]. The two–tailed lipid anchor exhibited superior membrane anchoring ability, as compared to the single–tail lipid, implying that increasing the hydrophobicity of the anchoring materials improved the interaction with cell membranes, and resulted in rapid anchoring ability with high retention time. However, a few cells did not express on the cell membrane, so gene transfer methods show some limitations. Therefore, to design protein–enriched surface–based cells, anchoring lipid materials were further conjugated with some protein, i.e., streptavidin, antibody, and EGFP. The proteins-based anchoring lipid conjugates were successfully coated on the cells, and did not show any sign of toxicity.

The soluble part of the polymer on the living cell surface has been used in various biomedical applications. Cationic polymers (poly–lysine and polyethylene amine polymers) are used as gene delivery applications. However, due to the presence of the anionic charge on the cell membrane, the cationic polymer interacts strongly and shows cytotoxicity to living cells [124,125,126]. Therefore, anionic (poly vinyl alcohol) or neutral polymers (polyethylene glycol) can be applied to the cell membrane to induce cell–cell interaction or cell fusion. Owing to the absence of an anchoring moiety in both polymers, after conjugating a lipid/hydrophobic moiety, these modified polymers are expected to anchor the cell membrane. Lipid-conjugated synthetic polymeric derivatives exhibited a similar anchoring capability onto cell membranes. The lipid–conjugated synthetic polymeric derivatives can be applied to cell membrane by three different methods: (1) Lipid–polymer polymers can directly anchor to the cell membrane through hydrophobic insertion, whereas lipid moiety anchors to cell membrane and polymeric moiety to outside the membrane. (2) PEG conjugated with functional moiety (i.e., NHS) can be used for direct covalent modification of the cell membrane. (3) Ionic polymers (cationic or anionic) can be used through electrostatic interaction on the cell membrane, whereas ionic polymer can be attached with cell membrane functional moieties [127,128,129].

The interaction between the cell membrane surface and the polymer has major influence on the uptake and exclusion. The different modes of interaction also affect the dynamics and stability of modified polymers on the cell membrane. To improve the acceptance rate of therapeutic cells, the modified synthetic polymers are used to enclose cell surface antigens and immobilize bioactive molecules on the cellular surfaces [130]. Modified synthetic polymers i.e., hydrophobic containing (PEG–lipid, PVA–alkyl), charged polymers (cationic PEI and anionic PVA), and NHS-modified PEG were taken to modified living cell membrane. Each modified polymer acted differently with changing the surface availability of ligands. The PEG–lipid and PVA–alkyl anchored through hydrophobic interaction, charged polymers anchored through electrostatic interaction and NHS–PEG formed the covalent bonds with available surface membrane proteins. After successfully applying all the polymers to CCRF–CEM or HEK293 cell membrane, the PEG–lipid, PVA–alkyl, and NHS–PEG polymers show rapid homogenous surface coating, while PEI polymer destroyed the cell membrane integrity. Cationic PEI polymer strongly interact with cell membrane and shows toxic effects to cells [131, 132].

The lipid composition and chain length have been considered important for regulating cell surface anchoring [133]. Figure 5 demonstrates a series of lipid–PEG conjugates with different carbon chain length, including PEG–DMPE (C14), PEG–DPPE (C16), and PEG–DSPE (C18). After incubation of hepatocytes cells with different lipid-based conjugates, PEG–DMPE (C14) exhibited uniform presentation with the highest anchoring efficiency on hepatocyte cell membranes. As compared to DMPE, the DSPE and DPPE–based conjugates were not sufficiently present on the cell membrane [134]. However, the effect of carbon chain length in lipids on cell surface anchoring should be precisely controlled, since controversial results were also reported. When two tail lipids conjugated with single–strand DNA oligonucleotides were applied for Jurkat cell surface anchoring, C16 lipid–DNA exhibited faster anchoring, compared with C18 lipid–DNA conjugates [135]. Similarly, Lee et al. have designed an antibody drug conjugate (ADC), i.e., trastuzumab emtansine with amine terminated DMPE–PEG to modified NK cell surface membrane. The designed ADC was successfully embedded on NK cells through hydrophobic interaction between the lipid–bilayer and anchoring lipid moiety [136].

Fig. 5
figure 5

Hydrophobicity inducing surface coated primary hepatocytes cells. The primary hepatocytes cells coated with different FITC–labbled Lipid–PEG conjugates

The limited chemical space and retention of lipid anchors on the cell membranes is the structural parameter of different anchors. The passive exogenous insertion of different lipid motifs or chemically defined structures i.e., single–chain lipid, double–chain lipid, cholesterol, and vitamin E through hydrophobic insertion could be an alternative approach to modify the cell membranes [119]. The different lipidic moieties show that different–different anchoring efficacy depends on the hydrophobic nature of the anchoring molecules and cell membrane [137]. To check the cell–coating efficacy of different lipids, Uvyn et al. synthesized dinitrophenol (antibody recruiting motifs) functionalized with single–tail lipid, two–tail lipid, cholesterol, and PEG–biotin as cell membrane anchoring moieties. After incubating the different lipid–conjugates with CT26 cancer cells for 2 h, the two–tail lipid–polymer showed homogeneous coating on the cell membrane as compared to other conjugated motifs [138].

The incorporation of additional biomolecules (such as peptide or surface proteins) could augment cell–cell adhesion of surface–coated cells. One example is benzylguanine (BG) functionalized DSPE–PEG tagged with SNAP–tag protein for the cell membrane coating. The cell surface bounded proteins are mainly critical for the cell signaling, cell–cell communication and cell recognition properties. Therefore, to simplify this issue, Rudd et al. used benzylguanine (BG) functionalized DSPE–PEG tagged with SNAP–tag protein for the cell membrane coating. This method simplifies and controls the phospholipid protein expression on cell membrane, and would be a promising method for proteins targeting cellular membranes [139].

Similarly, peptide-conjugated lipids could also be applied to facilitate cell–cell attachment. Here, by introducing PEG–lipid on the cell membrane coating, the cell–cell attachment was significantly increased through peptides interactions. Two different type of oligopeptides EIAALEK (fuE3), and KIAALKE (fuK3) sequences, were conjugated with the end group of lipid-PEG moiety and used specifically for heterodimeric interaction between cells [140]. Interestingly, photocleavable materials in lipid–PEG conjugates offer light responsive cell–adhesion ligands [141,142,143,144]. It has also been reported that the cell membrane modified with PEG single–chain lipids could show better stably anchoring profile, as compared to dual–chain lipid–based PEG conjugates [145]. Based on such versatility, light induced PEG–lipids serve as turn-off-type cell-anchoring substances for targeted cell attachment to locations that are not exposed to light [146].

The non-invasive remote regulation of cell anchoring to biomaterial has been made possible by the development of stimuli-responsive cleavable materials that are sensitive to heat, voltage, and light. The photo-sensitive materials are the most promising, and offer the ability to control cell surface modification even at single cell level accuracy. Therefore, Yamahira et al. conjugated a photocleavable molecule (4-[4-(1-hydroxy-ethyl)-2-methoxy-5-nitrophenoxy] butyric acid) with PEG–single chain lipid, which was further conjugated with additional lipid moiety to make a dual–lipid based PEG conjugate. Here, for the anchoring comparison, different chain length lipids were conjugated with PEG–single lipid conjugate. The insertion of second lipid chain to PEG–single lipid prevents cell membrane anchoring. Before exposer of light, dual chain lipid–PEG inhibits the cell membrane anchoring and after the exposure to 365 nm light, cell anchoring switched off on the same surface [147].

Polymer-based conjugates anchor cell membrane

Cell membrane anchoring using a variety of natural or synthetic polymers provides new opportunities for bioengineering applications. Polymer derivatives with multifunctional moieties could also be easily introduced on the cell membrane with the aid of bioactive membrane anchors [148, 149]. In recent years, three major strategies have been suggested for cell membrane anchoring: (1) amphiphilic polymers and PEGs via hydrophobic insertion, (2) covalent conjugation of polymers with active functional components in cell membrane surface, and (3) layer-by-layer surface deposition by ionic polymers.

However, the covalent conjugation and layer-by-layer cell surface methods have limitations such as disturbing the signaling functions of surface proteins and compromising efficacies. The covalent surface modification has been achieved through only chemical, enzymatic, or metabolic treatment which introduce a variety of functional groups including azide, amine, carboxylic, biotin, N-hydroxy succinimide ester, maleimide, and ketones. However, these chemical methods would perturb the cell membrane or cells, and can also disturb physiological signaling processes. Similarly, the layer-by-layer deposition of ionic polymers has exhibited detrimental effects toward target cells due to the cytotoxicity of such ionic materials and random surface coverage inhibiting signal binding or secretion through membrane surfaces.

Synthetic polymers (PEG, PEI, PAA and PVA) [150] and natural polymers (alginate, chitosan, hyaluronic acid and dextran etc.) [151,152,153] are good biopolymers with biodegradable, biocompatible and well-arranged structural properties. These polymers could be utilized for the surface modification of NK cells for biomedical applications (Fig. 6). However, the lack of ideal cell anchoring and cell mimicking functional moieties render the surface features of these polymers inappropriate for cell surface engineering.

Fig. 6
figure 6

Chemical structures of different carboxylate, amine, aldehyde, azide, thiol, maleimide, DBCO, and alkene-based biopolymers that can be utilized for cell membrane modification

The negative charge on the cell membrane helps for electrostatic binding sites for the charged polymers. The negative charged based-polymers cannot attach to cell membrane due to charge repulsion forces while the cation-based polymers can bind with cell membrane due to opposite charge binding. Different studies based on polycations were used to modify mammalian cells, like poly–L–lysine, PEI and PAH [154,155,156], since the interaction between cationic polymers and cell membrane shows potential cytotoxicity [157]. Additionally, incubating cells with polycationic polymers forms a speckled cell covering by electrostatic adherence to the cell surfaces. Therefore, the suitable modification of these polymers could be utilized as cell–coating biomaterials.

Hydrophobic modification of polymers provides an opportunity to change the properties of these biopolymers to achieve a regular cell–coating without altering the cell membrane physiology [5, 158, 159]. The availability of versatile polymer modification methods such as EDC/NHS-based coupling reactions, NHS–ester based coupling, or Michael addition, can be used for biopolymer modification. As discussed in Sect. 3.1, the different types of lipids, including tail basis and C-chain length basis lipids, can be utilized to design coating platforms. However, two tail lipids, i.e., DSPE– and DMPE–based polymers, showed homogenous cell coating efficiency [138]. Therefore, conjugation of DSPE and DMPE lipids with polymers could provide better cell–coating efficacy, and enhance cell–cell interaction (Fig. 7).

Fig. 7
figure 7

Schemes representing the general synthesis strategies of different hydrophobic-PEG biomaterials for cell–coating applications

Systematic surface engineering without interfering with NK cells’ own death receptor ligands is crucial for effectively augmenting cell recognition ability and the therapeutic efficacy of NK cells [160]. However, previously reported cell–coating materials still exhibit various limitations to achieving sufficient therapeutic efficacy of surface-coated NK cells. Due to the rapid dynamic nature of NK cell membrane, a single usage of cholesterol or one-tailed lipid molecules could be internalized into NK cell cytoplasm, and induce insufficient membrane-binding capability.

To this end, new design strategies for multifunctional polymeric anchors for enhanced NK cell surface engineering should be suggested: (1) two-tailed lipid molecules for efficient cell anchoring efficacy, (2) poly(ethylene glycol) (PEG) for increased solubility of the fabricated polymeric anchors and the inhibition of cytoplasmic influx, (3) cationic amino acids (such as arginine) to augmente electrostatic interaction with cellular membranes, and (4) cancer recognition moiety to facilitate membrane–mediated recognition of target cancer cells.

Figures 8 and 9 represent possible design approaches for natural and synthetic polymeric anchors for NK cell–coating materials. The incorporation of such modular moieties could both offer sufficient localization onto NK cell membranes, and facilitate target recognition, and eventually enhance anticancer efficacies of surface–modified NK cells to treat a series of cancers. The addition of PEG also showed higher immune response and the prolonged persistence of targeted moieties [161, 162].

Fig. 8
figure 8

Schemes representing the modification strategies of natural biopolymers for designing two–tail lipid-based cell–coting platforms

Fig. 9
figure 9

Schemes representing the modification strategies of synthetic polymers for designing cell–coting platforms

Natural polysaccharides boost the immune system and help to combat several ailments [77]. Polysaccharides, which are closely related to immune regulation, mainly increase the activity of immune cells, promote the secretion of cytokinin, and enhance immune functions. Polysaccharides also enhanced the activity of NK cells and macrophages to increase phagocytosis toward foreign particles [163]. In addition, the overexpression of MMPs in tumor tissues remodulates the ECM to promote metastasis. The use of bioactive polysaccharides could also inhibit metastasis by the suppression angiogenesis and inhibit the activity of MMPs [164]. Polysaccharides, like hyaluronic acid and chondroitin sulfate, can easily avoid the ATP binding and could directly target CD44 overexpressing cancer cells [165, 166]. Polysaccharides are widely used for anticancer drug delivery applications to bypass the side effects of many drugs. Polysaccharides may also stimulate the intrinsic anticancer immune system themselves. In this regard, many polysaccharides, such as chitosan, hyaluronic acid, dextran, inulin, sodium alginate, and other polysaccharides have been explored as potential drug delivery systems, and as well for immune adjuvants for cancer treatment via the promoting release of interferon–γ and granzymes [167,168,169,170]. The polysaccharides also regulate the immune mechanism via different pathways, including NF–κB, TLR4, and notch pathways. These pathways have major influence on NK cells, dendritic cells, and macrophages to improve the immune microenvironment [171]. Polysaccharides are highly safe for human applications; however their rapid elimination, short half-life, and lack of specific binding hinder their pharmacological activities. Some researchers maintain that the proper modifications of polysaccharides have more immunomodulatory activity than those without functional moieties [167].

Therefore, Fig. 8 demonstrates the developments of next-generation polysaccharides-based biomaterials for cell–coating applications. Polysaccharide backbone could be a framework for multiple chemical conjugation along with inducing interaction with membranes of target cancer cells.

For example, functionalized chitosan-based materials have been developed for plasma cell membrane coating [172]. With the aid of cholesterol as hydrophobic anchoring moiety, the surface integration of the resultant materials maintained integrity up to 6 h without cellular internalization. Another example of chitosan–cholesterol–biotin was also applied for plasma cell membrane coating, and showed stable surface anchoring up to 8 h [173]. Therefore, the incorporation of natural polymer backbone with hydrophobic lipid anchor could be an innovative strategy for cell surface modification and labeling to prolong cell membrane behavior. Similarly, synthetic block–copolymers have also been used for cell–coating materials (Fig. 9), for the purpose of enhancing cellular activities including migration of stem to injured tissues, hiding transplant cells from immune response, and promoting natural differentiation in stem cells [174,175,176]. While block–copolymer incorporation can occur without modification of the cell membrane, such modification can disturb the functions, compared to the therapeutic efficacy of cells. Therefore, the insertion of two–tailed hydrophobic lipid segments shows rapid cell membrane anchoring.

Block–copolymers modified with functional moieties can easily impact cell membrane and these materials can be utilized to improve cell-based therapies. In recent years, modified block-co–polymers have been applied for cell membrane modification. For example, poly(oxanorbornene)-based block copolymer conjugated with Brij moiety was anchored on the living Jurkat cell membrane. The synthesized Brij-based copolymer was introduced on the periphery of cell membrane, without showing any toxic effects up to 24 h of incubation. The polymer concentration below the critical micelle concentration (5 µM) was used to anchor at the 3T3 fibroblast cell membrane to facilitate homogenous surface coating without internalization. The flow cytometry results suggested the surface coating of the cell membrane without altering the cell membrane and cellular internalization. After successfully coating the cell membrane, the polymers did not show cytotoxicity after 24 h of incubation. Additionally, photosensitizer molecule conjugated into this copolymer was used to generate localized 1O2 near the cell surface for trigger cell death without harming healthy cells [177]. Living cell microencapsulation can provide alternative methods to develop new therapeutic effects. Unfortunately, due to the large size of microencapsulated cells, they have mass transportation issues, and clinically are not suitable for veins [178]. The transplantation of microencapsulated islet cells shows limited vascular supply, which generates a hypoxic cell condition [179]. Therefore, to avoid this issue and minimize the capsule size of encapsulated islet cells, Willson et al., utilized a layer–by–layer approach for intraportal islet transplantation. The poly (L-Lysine)-g-poly(ethylene glycol) conjugated with biotin (PBB) was synthesized, and used for the islet cells surface modification through a layer–by–layer cell–coating method using streptavidin, after the successful layer-by-layer surface coating of islets using a PBB copolymer that facilitates cell surface growth without toxicity and loss of cell functions. Therefore, the nanothin coating of islets cells provides a novel approach, and reveals an informative relationship between cationic charge density, cell membrane attachment and biocompatibility for targeted cell delivery applications [180].

Anchored NK cell mediated biomedical applications

Surface-functionalized NK cells have highly advantageous properties towards targeting malignant tumors by generating facilitated immune synapse interaction and cell–cell binding [181]. Subsequently, modified NK cells showed the enhanced therapeutic potentials of upregulated target recognition and cancer killing efficacy against various tumors. Therefore, the effective surface presentation of functional moieties (such as glycan, nanoparticles, aptamers, and antibody) onto NK cell membranes has been developed, as described in this section.

NK cells surface engineering using glycans

The glycoengineering of NK cells can provide precise targetability and desirable cell recognition. Glyco analogs show high specific affinity with Siglecs target cell membrane. For example, sialic acid analogs, like C9 of Neu5Ac, -N-biphenylcarboxamide-Neu5Ac and 9-N-m-phenoxybenzamide-Neu5Ac showed specificity and high affinity with CD22 [182]. Cancer cell overexpressed inhibitory ligands could avoid immune cell detection and following attack. For example, sialylation on cancer cells disturbs cell–cell interaction between NK cells and tumor cells.

The glycol-based molecules have been added by both direct and indirect conjugation at the cell membrane surface [183,184,185]. These processes can also be achieved by both covalent and non-covalent processing. Linker molecules are often required to conjugate different functional groups including maleimide, azide, hydrazide, and alkynes. However, the direct glycan conjugation strategy onto cellular surfaces still has a few drawbacks, such as low binding efficiency and transient nature [186].

To enhance glycan-mediated cell–cell interaction, a few strategies have also been suggested whereby a glycocalyx engineering approach could be applied to elucidate the roles of specific sialosides in facilitating Siglec–based immunoevasion. The tumor cells engineered with sialylated glycopolymers are prevented from NK cell killing due to strong Siglec binding (Fig. 10a) [86].

Fig. 10
figure 10

Physical attributes of different glycans at NK cell surfaces affecting the immunological functions to augmented cancer cell killing. a Glycoengineering approach for cancer cell membrane modification to prevent killing from NK cell, (b) glycomodulation of NK cells via CD22 specific ligands, (c) surface modulation of NK cells by Siglic–7, and (d) metalbolic glycoengineering of NK cells for antibody-mediated targeting cancer cell killing

CD22 highly expressed various type cancer cells and it can be selectively targeted by Sigleic–2 glycan. To achieve tumor-specific CD22 targeting, highly specific Neu5Ac analogue ligands were conjugated with NK–92MI cells through chemoenzymatic generation to enhance the therapeutic efficacy of the cell for better cancer treatment. The two unnatural Neu5Ac analogues, CMP–BPCNeu5Ac and CMP–MPBNeu5Ac conjugated on NK–92MI cells show 20–50 fold higher binding affinity with the CD22 ligand in comparison to CMP-Neu5Ac to install the natural ligands [187]. The modified sialoside analogues containing heterocyclic ring moieties also show high affinity and selectivity towards siglecs. For example, the addition of triazole, benzamide, acetylene, azides, and dimethyl benzamide substituent in sialosides through click chemistry improved the affinity and selectivity toward CD22 and CD33 overexpressed cells [188].

Glycoengineering with 9-O modified sialic acid of NK cells, can significantly improve the selective binding with Siglec–2 B-cell-restricted antigen CD22+ cells. The addition of MPB-sia group through the sialic acid biosynthetic pathway greatly enhanced the binding affinity with CD22+ cells. The glycoengineered NK cells represent CD22–dependent cytotoxicity against CD22+ primary lymphoma cells (Fig. 10b) [189].

Chemoenzymatic glycans (ST6Gal1) surface editing of NK cells shows high affinity and specificity toward the modulation of Siglec–7 signaling. The ST6Gal1-mediated surface glycan was added to the NK cell surface to generate cis-type interaction toward target cell. The higher glycan ligands on the NK cell surface enhanced Siglic–7 phosphorylation and SHP–1 release, which suppresses the NK–induced target tumor killing while lower-level help to release Siglec–7 which helps to restore NK cell killing efficacy. Hence, low level of high–affinity ligands in cis manner destroys Siglec–7 clusters which release Siglec–7 through secretion pathways and triggered immune killing efficacy (Fig. 10c) [190].

The ex vivo glycoengineering with IL–21on the NK cell surface induced effective antitumor targeting and infiltration, and enhanced autoimmune response and potential therapeutic response. IL–21 is an imperative regulatory marker that can efficiently activate and start transcription factors to modify autoimmune reactions. However, the attachment of IL–21 to NK cells by tradition nal methods has a week effect on NK cells. Therefore, tumor–specific release IL–21 nanoparticles were designed to enhance the activation, expansion, and safety performance of NK cells. Bio-orthogonal chemical groups (N3 and BCN) were separately incorporated on the surface of NK cells. After that, cytokinin IL–21-based redox responsive nanoparticles were conjugated on –N3-containing NK cell surface. Finally, bio-orthogonal live-cell nanoparticles augmented the interaction between NK and target cell and the in situ controlled release of IL–21 from nanoparticles, promoting the more effective and therapeutic efficacy of NK cells [191].

Similarly, metabolically glycoengineered NK cells conjugated with antibodies through biorthogonal reaction were used to efficiently enhance targeted cancer killing efficacy. The NK cells were glycoengineered with 9-azido N-acetyl neuraminic acid methyl ester (N3–SA), and generated the surface glycan with a free azide group. Elsewhere, a monoclonal antibody (mAB) was separately conjugated with DBCO–PEG4–NHS ester to generate mAB–DBCO (Fig. 10d). The synthesized mAB–DBCO was further connected with N3–NK cells through biorthogonal reaction. The glycoengineered antibody-based NK cells demonstrated better binding efficacy with EGFR positive KRAS mutant SW480 human colorectal cancer cells as compared to native antibodies and uncoated NK cells [192].

Nanoparticle-mediated NK cells

The physical specification of nanomaterials can modulate the NK cell immune response for cancer treatment. The nanomaterials design strategies and specifications, like structure, surface charge, valency, and hydrophilic/hydrophilic nature, can be designed to induced potential immune response between NK cell and cancer cells [193]. Nanoparticles play an important role in the activation of NK cells, and also act as a good immunomodulator to increase immunotherapeutic efficacy. Nanoparticles shown different pharmacodynamic and pharmacokinetic properties and lower dose of nanoparticles somehow modulates the immune system with minimum side-effects [194].

The tumor killing vitality of NK cell was influenced by the tumor–secreted regulators. For example, tumor secreted TGF–β is a negative regulator for IFN–γ production by NK cells. TGF–β also downregulates the NK cell activation receptors, like NKp30, NKp46, and NKG2D, which reduced the toxicity of NK cells toward tumor cells. Park et al. reported a strategy to modulate TGF–β signaling using nanoscale liposomal gel, which showed tumor killing efficacy through NK cell activation. Nanoscale liposomal gel potentially inhibits TGF–β signaling, and successfully delivered the IL–12 into the tumor site. The results of this study demonstrated the delivery of hydrophobic and hydrophilic immunomodulators to enhance cancer killing efficacy against melanoma and breast cancer cells [195]. Similarly, Liu et al. designed a nanoemulsion system for the co-delivery of TGF–β inhibitor and selenocysteine (SeC) to enhance NK cell tumor-killing efficacy against triple-negative breast cancer cells. The designed nano emulsion system effectively downregulates TGF–β/TGF–β RI/Smad2/3 signaling and enhance immune response through NKG2DL expression on the cancer cell and NKG2D expression on the NK cell surface (Fig. 11a). The site-specific release of TGF–β inhibitor and SeC significantly increased the antitumor efficacy against breast cancer [196].

Fig. 11
figure 11

The nanomaterials affecting the immunological functions of NK cells. a A nanomicelles system activating the NK cells and augmenting NK cell functions for cancer treatment, (b) physical attributes of different nanoparticles altering the immunological functions of NK cells, and (c) Acid-responsive chemo drug-based micelles immune synapse generated by the release of dox in acidic environment

Recent nanotechnology technologies further enable the development of a variety of nanoparticle systems to achieve better therapeutic efficacy of NK cells [197]. Therefore, the surface modulation of NK cells using nanoparticle system is also useful for NK cell homing and infiltration to achieve efficacy of NK cell cancer immunotherapy. Recently, various nanoparticle-based strategies showed promising NK cell tumor infiltration, where NK cells stimulate the production of pro-inflammatory cytokines, and induced cytotoxicity against tumor cells [198, 199]. In addition, the nanoparticles attached to the NK cell surface can also control the moment of NK cell after local delivery using an external magnetic field. For example, Jang et al. developed Cys5.5–conjugated Fe2O3/SiO2 nanoparticles for modification of the surface of NK cells [200]. Cys5.5 Fe2O3/SiO2 was attached to NK cells, and modified Cys5.5 Fe2O3/SiO2@NK cells were used for cancer recognition and tumor targeting (Fig. 11b). Magnetic field induced increase in the relative ratio of NK cells infiltrated into tumor tissue by 17 times, even at the minimum concentration of nanoparticles (20 μg Fe/mL), which is relatively lower than conventional magnetic resonance monitoring agents. Therefore, the results were explored to apply a cell-based therapy with MRI imaging for tumor treatment without surgery. On other hand, Wu and coauthors synthesized biocompatible polydopamine-coated Fe3O4 nanoparticles (Fe3O4@PDA), which were encapsulated by NK cells [201]. The synthesized Fe3O4@PDA nanoparticles were successfully internalized into NK cells without affecting the intrinsic properties of NK cells. The presence of markers at the NK cell surface indicates the biocompatibility of nanoparticles, and hence induced the maturation of NK cells. The Fe3O4@PDA–loaded NK cells significantly induced apoptosis against A549 cells and inhibited tumor growth as compared to unmodified NK cells. Fe3O4@PDA loaded NK cells also pointedly reduced the Ki67 + tumor cells with an increasing number of apoptotic cells (Fig. 11b). In another study, the NK cell surface was engineered using iron oxide nanoparticles (IONPs) to localize delivery to the desired targeted site of action under a magnetic field [202]. IONPs were immobilized on the NK cell surface using a robust bioconjugation technique by attaching sulfo–NHS biotin to the surface amine, and attached streptavidin–coated IONPs. The activity of NK cells, including phenotype and functions, was maintained over time. In addition, the bio-hybrid therapeutic approach was successfully homed with magnetic response to improve antitumor efficacy in contrast to naked NK cells.

Recently, nanoparticle-mediated photothermal strategies have attracted great attention for cancer immunotherapy. Photothermal therapy is oxygen–independent, hence it can be used efficiently to kill solid tumors by overcoming tumor hypoxic conditions. Nanoparticles-based photothermal strategies generated heat by inducing infrared irradiation, and were able to target tumor tissues, and have desirable tumor–killing efficacy [203]. The nanoparticle photothermal agents promote the development and infiltration of cytotoxic CD8+ T cells through pro-inflammatory cytokines release. Therefore, the nanoparticulated photothermal agents provide an active immune response in addition to reversing the TME's immunologically "cool" state [204]. Combined photo–immunotherapy offers a promising therapeutic approach. For example, Zhang et al. produced 2D manganese-based coordination nanosheets (CONASHs), and then the attachment of polyether imide and DNAzyme; the DNAzyme@Mn–CONASHs were anchored on the surface of the NK cell [205]. Further, the TLS11a aptamer was conjugated with DNAzyme@Mn–CONASHs@NK through a glycan biosynthesis approach. The synthesized platform was homogenously anchored on the surface of NK cells, which was confirmed by CLSM images labeled with Cys3–TLS11a aptamer–DBCO. The stability of the TLS11a aptamer in NK cells was controlled for up to 24 h. In addition, the morphology of HepG2 cells that were treated without TLS11a aptamer NK cells did not change, but when treated with NK cells decorated with TLS11a aptamer, were significantly changed. After in vivo injection into mice, the DNAzyme@Mn–CONASHs based on NK cells showed better tumor inhibition compared to the case for DNAzyme@Mn–CONASHs alone.

Cationic biomaterials are transfection agents, that primarily activate and trigger immune cells like macrophages, dendritic, or other immune cells (Fig. 11b). Cationic biomaterials also demonstrate adaptive immune responses by promoting immunity with stimulation to produce porin-inflammatory cytokine by immune cells. For example, PEI, a cationic polymer shows a tendency to activate macrophages through TLR–4 interaction. Elsewhere, the intratumoral injection of PEI increases the expression of IL–12, and decreases the expression of IL–10 [206, 207]. Kwang-Soo Kim and coauthors synthesized PEI–immobilized PDA–coated Zn/Fe magnetic nanoparticles (41 nm) for NK–92MI cell activation. The synthesized magnetic nanoparticles did not show toxicity against NK cells, and successfully enhanced cytolytic activity against triple-negative breast cancer cell [208]. In other research, Im et al. designed a pH-responsive micelle system for the site–specific delivery of chemo drug (DOX) through immunotherapy (Fig. 11c). When the micelle-NK cells were cultured with cancer cells, the pH between NK and cancer cells decreased, due to releasing lytic granules from NK cells. Therefore, the interaction between NK, and cancer cells generates acidic conditions that triggered the site–specific release of chemo drug [209].

Aptamer-immobilized NK cells

Aptamers are the 3D structure of DNA/RNA oligonucleotides, that can specifically bind to selective proteins. Currently, aptamers are widely well-accepted ideal candidates for the targeted delivery, diagnosis, and immunomodulatory agents of cancer. A great benefit of aptamers is their facile chemical modification and molecular engineering. The chemical modification used aptamers prior to enhancing the stability against serum, and enhanced targeting specific binding ability [210]. In cancer immunotherapy, aptamers allow for simultaneous cancer cell recognition, and induced immune synapse. Aptamers show high affinity with various co-inhibitory immune checkpoints including CTLA–4, PD–L1, PD–1, TIM–3, and LAG–3. Aptamers also activate the immune cells through INF–γ release to prevent tumor growth [211,212,213].

To develop aptamers functionalized NK cells for targeted cancer immunotherapy, Yang et al., synthesized CD-30 specific aptamer-based different lipid anchors including single and double chains. As mentioned in Fig. 12a, different lipid moieties, like an 18–C single chain, 18–C × 2 two-tailed lipid, cholesterol, and vitamin-E were used for the CD–30 specific aptamer conjugation. After 30 min incubation of lipid–aptamer with NK cells, aptamer–18–C × 2 two-tailed lipids showed good and homogeneous surface coating without the intact cell morphology of NK cells. Owing to the absence of intracellular signaling domains, the 18–C × 2–NK cells showed potential cell–specific binding without affecting the activity of NK cells, and enhanced targeted cell killing against lymphoma cells for improved immunotherapy [214].

Fig. 12
figure 12

Aptamer-based NK cell modifications (a) Lipid-based different aptamers NK cell anchoring and CD30 specific targeted cell binding, and (b) direct NK cell functionalization using aptamers for specific cancer cell recognition

As NK cell-based therapeutic medicine circulates in the body with certain shear conditions before reaching the targeted site of action, hence the body shear flow environment can significantly influence cell–cell interactions [215, 216]. Therefore, NK cells coated with multifunctional biomaterials can control the cell–cell interaction in shear stress conditions. Multivalent aptamers also have an advantage in targeted immunotherapy under body shear conditions. The benefit of the multivalent aptamer is the ability to direct focus on immune signals to the cell group of interest. To develop cell–cell interaction, one partition of aptamer can bind to tumor cell surface antigen or in TME, and another partition can bind with the specific receptors of immune cells. Therefore, to prove the concept of multivalent aptamer mimicking cell–cell interaction under shear stress, Shi and coworkers synthesized in situ aptamer–based polyvalent antibody mimic (PAM) and monovalent antibody mimic (MAM) for cancer cell targeting and enhance cell–cell binding. The synthesis of polyvalent aptamers involved three major steps: (1) DNA initiator on cell surface, (2) the formation of supramolecular DNA scaffolds on the cell surface, and (3) the direct connection of DNA scaffolds with aptamers (Fig. 12b). The PAM@NK cells are more capable than MAM@NK cells in recognition and binding to K562 cancer cells under shear stress conditions [216].

Antibody-dependent NK cell-mediated cytotoxicity

The concept of antibody for targeted therapy (magic bullet) was first advanced by Paul Ehrlich [217] who suggested the specific detection and suppression of malignant tissues. Five sequence-based antibodies like IgG, IgM, IgD, IgE, and IgA are found in two different subunits: (1) constant fragment and (2) fragment of antigen binding. The constant fragment antibodies are linked with immune effector functions and show binding to IgG receptors (FcyRs) and neonatal FcR (FcRn) [218]. These antibodies can sufficiently interact with the overexpressed surface receptors of tumors, accumulate into tumor tissues, and significantly inhibit tumor growth. The interaction between antibody and tumor receptors allows immune effector functionalities like Fc receptors overexpressed by immune cells [219].

The unconjugated antibodies have lower immune effector function efficacy, hence modification via chemical or genetic agents may increase the targeted immunotherapy. Mainly, the capacity of antibodies to generate adaptive immune effectors such as NK and T cells can induce anticancer activity. In terms of NK cells, the overexpression of phosphatase motifs (SHP-1/2 or SHIP) also inhibits the immune system [220]. Therefore, systematic antibody modification or conjugation may expand the ability to stimulate immune effector functionalities [221,222,223]. In past years, the antibody–dependent engagement of immune mechanisms has been used for cancer cell killing. However, during the multistep cell–coating process, the addition of antibodies through genetic engineering has some technical complications and safety issues. The major limitation is interference with cell endogenous functions during surface modification (Fig. 13a). Therefore, antibody-based surface engineering at the cell membrane using chemo-bio tools in single-step has emerged as a complementary and useful technique (Fig. 13b). In the Jie Li and coworkers study, NK cell surface was engineered using GDP Fucose-conjugated human IgG (GF–hIgG) antibody with Herceptin to target HER2 + breast cancer [224]. The half-life of Herceptin was maintained up to 20 h, and it selectively binds with BT474‒, a HER2 + breast cancer cells, via cluster formation. These surface–modified NK cells with Herceptin were shown to produce more than 7–times lysis of BT474 cells, as compared to the unmodified NK cells. The killing effects of Herceptin–NK cells were also confirmed in other HER2 + cells that showed potential cell-mediated cytotoxicity with effective targetability in ex vivo and in vivo death. In another study, Dapeng Li and coauthors had 3H4 and CA147 IgM antibodies, and utilized them for NK cell surface engineering [225]. The modified NK cells efficiently bind with HLA–E–VL9 and inhibit NKG2A/CD94 which suggested the 3H4 IgM enhances NKG2A + NK cell death. Finally, this study suggested that mouse 3H4 IgM and human CA147 antibodies potentially work as an NK checkpoint inhibitor, and target HLA–E–VL9 (Fig. 14a).

Fig. 13
figure 13

a Scheme representing the multistep cell–coating process and affecting functions of NK cells and (b) one-step NK cell–coating process augmented the NK cell functionalities and targeted cancer cell binding

Fig. 14
figure 14

a Antbody affecting the binding between NK and cancer cell, and (b) non-genetic NK cell surface modification approach for effective cancer targeting

Moreover, Gong and coworkers synthesized anti–EGFR nanobody 7D12 for cancer cell targeting as an alternative genetic approach method. NK92MI cells were metabolized to generate azo-modified sialic acid, and it was then further conjugated with aza–DEBCO-modified anti-EGFR nanobodies by bio–orthogonal click chemistry method [226]. After that, 7D12NK cells were applied in the different EGFR+ tumor cell lines, including MDA-MB-468, A549, A431, LoVo, and RKO cells. The developed therapeutic approach 7D12NK cells demonstrated active targeting toward EGFR overexpressed cancer cells. This new type of biotherapeutic approach is capable of penetrating solid tumor tissues, with low immunogenicity and improved therapeutic efficacy of the NK cells in both ex vivo and in vivo treatments (Fig. 14b).

Conclusion and future perspectives

The development of novel biomaterials is being pursued to satisfy the urgent requirements of cell–based therapies with their unique performance. In recent years different cell-coating strategies have been used for the development of cell-based therapies including CAR–based therapy. The cell–coating techniques successfully tuned the surface properties of cells like cell tracking, imaging, and cell–cell interactions. Controlling cell surface engineering can be achieved by modifying a cell membrane with an active polymer. In general, three major types of polymers have been used in cell surface engineering, namely: (1) ionic polymers for layer-by-layer deposition, (2) reactive functional polymers for direct surface functionalization, and (3) amphiphilic polymers for hydrophobic insertion. In the following context, the utilization of these materials and their perspective application to augmented cell functionalities are briefly discussed. This review has introduced different types of polymeric biomaterials for controlling NK cell behavior, a detailed discussion of surface functionalized NK cells with their potential anticancer applications, and the state-of-the-art in the design of some new biomaterials to alleviate NK cell surface behavior for cancer immunotherapy. Some most common important cancer recognition moieties including folic acid, lactobionic acid, and phenylboronic acid, and their potential applications for cell surface utilizations are also discussed. Elsewhere, most of the currently developed cell-based methodologies and modified NK cells with biomaterials and/or other active templates are introduced. To resolve the inevitable cell–cell interactions by the conventionally available methods, new strategies have also been proposed for large-scale development of biomaterials for cell–surface modifications.

Advanced biomaterials synthesis techniques and their post-functionalization allow for the rational design and effectiveness of polymers toward new cell-based therapies. For example, biomaterials functionalized with lipid, PEGs, and cancer recognition moieties can represent great potential for immune cell surface engineering and immunotherapy applications. In particular, hydrophobic insertion and cancer recognition moiety can enrich the surface functionalities of NK cells for cancer treatment. Along with advancements in biomaterials functionalization, various architectures can precisely enhance the NK cell surface functionalities. Controlled conjugation and specific topology of polymeric biomaterial can be anchored on the NK cell surface with tunable and unique surface properties towards cancer treatment applications. However, the disadvantages of layer-by-layer and direct conjugation of materials with cell membranes cause many side effects. Elsewhere, the CAR–NK cell-based therapy is also a cost-effective technique. Focusing on this issue, the functionalization of different biomaterials has been proposed to develop surface-engineered immune cells for cancer immunotherapy applications. The development of a new cell-based therapy system has brought good hope for cancer patients. Cell surface engineering methodologies have significantly improved the therapeutic abilities of NK cells through different sources of functional motifs. Cell surface modification improves the protection of cells from harsh tumor microenvironments, inducing cell–cell communication, stability, and effective targeting ability without interfering with the intrinsic properties of cells. Therefore, the surface engineering of NK cells using multifunctional biomaterials is a new predictable concept for the development of novel cell-based therapeutics systems for solid tumor treatment.

Availability of data and materials

Not applicable.

Abbreviations

CAR:

Chimeric antigen receptors

NK:

Natural killer

TRAIL:

Tumor necrosis related apoptosis induced ligand

FasL:

Fas ligand

EPR:

Enhanced permeation and retention

FA:

Folic acid

ASGPR:

Asialoglycoprotein receptors

LBA:

Lactobionic acid

PD-1:

Programmed cell death protein 1

PD-L1:

Programmed death-ligand 1

ITIM:

Immunoreceptor tyrosine based inhibitory motif

PBA:

Phenylboronic acid

PEG:

Polyethylene glycol

PVA:

Polyvinyl alcohol

PEI:

Polyethylene amine

DMPE:

1,2-Dimyristoyl-sn-glycero-3-phosphoethanolamine

DPPE:

1,2-Dipalmitoyl-sn-glycero-3-phosphoethanolamine

DSPE:

1,2-Distearoyl-sn-glycero-3-phosphoethanolamine

ECM:

Extracellular matrix

MMP:

Matrix metalloproteinases

DBCO:

Dibenzocyclooctyne

IFN-γ:

Interferon gamma

PAM:

Polyvalent antibody mimic

MAM:

Monovalent antibody mimic

References

  1. Kellam B, De Bank PA, Shakesheff KM. Chemical modification of mammalian cell surfaces. Chem Soc Rev. 2003;32:327–37.

    Article  CAS  Google Scholar 

  2. Wu Z, Xiao M, Lai W, Sun Y, Li L, Hu Z, et al. Nucleic Acid-Based Cell Surface Engineering Strategies and Their Applications. ACS Appl Bio Mater. 2022;5:1901–15.

    Article  CAS  Google Scholar 

  3. Bi X, Yin J, Chen Guanbang A, Liu C-F. Chemical and Enzymatic Strategies for Bacterial and Mammalian Cell Surface Engineering. Chem – A Eur J. 2018;24(8042):50.

    Google Scholar 

  4. Amaral AJR, Pasparakis G. Cell membrane engineering with synthetic materials: Applications in cell spheroids, cellular glues and microtissue formation. Acta Biomater. 2019;90:21–36.

    Article  CAS  Google Scholar 

  5. Kim S, Kim K. Lipid-mediated ex vivo cell surface engineering for augmented cellular functionalities. Biomater Adv. 2022;140: 213059.

    Article  CAS  Google Scholar 

  6. Bashor CJ, Hilton IB, Bandukwala H, Smith DM, Veiseh O. Engineering the next generation of cell-based therapeutics. Nat Rev Drug Discov. 2022;21:655–75.

    Article  CAS  Google Scholar 

  7. Dai X, Mei Y, Cai D, Han W. Standardizing CAR-T therapy: Getting it scaled up. Biotechnol Adv. 2019;37:239–45.

  8. Dai H, Wang Y, Lu X, Han W. Chimeric Antigen Receptors Modified T-Cells for Cancer Therapy. JNCI J Natl Cancer Inst. 2016;108:439.

    Article  Google Scholar 

  9. Laskowski TJ, Biederstädt A, Rezvani K. Natural killer cells in antitumour adoptive cell immunotherapy. Nat Rev Cancer. 2022;22:557–75.

    Article  CAS  Google Scholar 

  10. Biederstädt A, Rezvani K. Engineering the next generation of CAR-NK immunotherapies. Int J Hematol. 2021;114:554–71.

    Article  Google Scholar 

  11. Daher M, Rezvani K. Outlook for New CAR-Based Therapies with a Focus on CAR NK Cells: What Lies Beyond CAR-Engineered T Cells in the Race against Cancer. Cancer Discov. 2021;11:45–58.

    Article  CAS  Google Scholar 

  12. Shin MH, Kim J, Lim SA, Kim J, Kim S-J, Lee K-M. NK Cell-Based Immunotherapies in Cancer. Immune Netw. 2020;20(2):1–27.

    Article  Google Scholar 

  13. Lee G, Karunanithi S, Posner B, Niederstrasser H, Cheng H, Federov Y, et al. Chemical screening identifies novel small molecule activators of natural killer cell cytotoxicity against cancer cells. Cancer Immunol Immunother. 2022;71:1671–80.

    Article  CAS  Google Scholar 

  14. Peng Y, Zhang W, Chen Y, Zhang L, Shen H, Wang Z, et al. Engineering c-Met-CAR NK-92 cells as a promising therapeutic candidate for lung adenocarcinoma. Pharmacol Res. 2023;188:106656.

    Article  CAS  Google Scholar 

  15. Fang F, Xiao W, Tian Z. NK cell-based immunotherapy for cancer. Semin Immunol. 2017;31:37–54.

    Article  CAS  Google Scholar 

  16. Jeong S, Kim YG, Kim S, Kim K. Enhanced anticancer efficacy of primed natural killer cells via coacervate-mediated exogenous interleukin-15 delivery. Biomater Sci. 2022;10:5968–79.

    Article  CAS  Google Scholar 

  17. Du N, Guo F, Wang Y, Cui J. NK Cell Therapy: A Rising Star in Cancer Treatment. Cancers. 2021;13(16):4129.

    Article  CAS  Google Scholar 

  18. Maddineni S, Silberstein JL, Sunwoo JB. Emerging NK cell therapies for cancer and the promise of next generation engineering of iPSC-derived NK cells. J Immunother Cancer. 2022;10:e004693.

    Article  Google Scholar 

  19. Jones AB, Rocco A, Lamb LS, Friedman GK, Hjelmeland AB. Regulation of NKG2D Stress Ligands and Its Relevance in Cancer Progression. Cancers. 2022;14(9):2339.

    Article  CAS  Google Scholar 

  20. Mensali N, Dillard P, Hebeisen M, Lorenz S, Theodossiou T, Myhre MR, et al. NK cells specifically TCR-dressed to kill cancer cells. EBioMedicine. 2019;40:106–17.

    Article  Google Scholar 

  21. Liu E, Tong Y, Dotti G, Shaim H, Savoldo B, Mukherjee M, et al. Cord blood NK cells engineered to express IL-15 and a CD19-targeted CAR show long-term persistence and potent antitumor activity. Leukemia. 2018;32:520–31.

    Article  CAS  Google Scholar 

  22. Li Y, Hermanson DL, Moriarity BS, Kaufman DS. Human iPSC-Derived Natural Killer Cells Engineered with Chimeric Antigen Receptors Enhance Anti-tumor Activity. Cell Stem Cell. 2018;23:181-192.e5.

    Article  CAS  Google Scholar 

  23. Ma S, Xu Y, Song W. Functional bionanomaterials for cell surface engineering in cancer immunotherapy. APL Bioeng. 2021;5:021506.

  24. Maverakis E, Kim K, Shimoda M, Gershwin ME, Patel F, Wilken R, et al. Glycans in the immune system and The Altered Glycan Theory of Autoimmunity: A critical review. J Autoimmun. 2015;57:1–13.

    Article  CAS  Google Scholar 

  25. Hudak JE, Bertozzi CR. Glycotherapy: New Advances Inspire a Reemergence of Glycans in Medicine. Chem Biol Cell Press. 2014;21:16–37.

    CAS  Google Scholar 

  26. Manhas J, Gupta RK, Roy PP. A Review on Automated Cancer Detection in Medical Images using Machine Learning and Deep Learning based Computational Techniques: Challenges and Opportunities. Arch Comput Methods Eng. 2022;29(2893):933.

    Google Scholar 

  27. Yoo J, Park C, Yi G, Lee D, Koo H. Active Targeting Strategies Using Biological Ligands for Nanoparticle Drug Delivery Systems. Cancers. 2019;11(5):640.

    Article  CAS  Google Scholar 

  28. Zhao Z, Ukidve A, Kim J, Mitragotri S. Targeting Strategies for Tissue-Specific Drug Delivery. Cell. 2020;181:151–67.

    Article  CAS  Google Scholar 

  29. Golombek SK, May J-N, Theek B, Appold L, Drude N, Kiessling F, et al. Tumor targeting via EPR: Strategies to enhance patient responses. Adv Drug Deliv Rev. 2018;130:17–38.

    Article  CAS  Google Scholar 

  30. Rosenblum D, Joshi N, Tao W, Karp JM, Peer D. Progress and challenges towards targeted delivery of cancer therapeutics. Nat Commun. 2018;9:1410.

    Article  Google Scholar 

  31. Wu Q, Yuan C, Liu N, Shu J, Wang J, Qian J, et al. Fast detection, a precise and sensitive diagnostic agent for breast cancer. J Exp Clin Cancer Res. 2022;41:201.

    Article  CAS  Google Scholar 

  32. Kim GJ, Nie S. Targeted cancer nanotherapy. Mater Today. 2005;8:28–33.

    Article  Google Scholar 

  33. Scaranti M, Cojocaru E, Banerjee S, Banerji U. Exploiting the folate receptor α in oncology. Nat Rev Clin Oncol. 2020;17:349–59.

    Article  Google Scholar 

  34. Marverti G, Marraccini C, Martello A, D’Arca D, Pacifico S, Guerrini R, et al. Folic Acid-Peptide Conjugates Combine Selective Cancer Cell Internalization with Thymidylate Synthase Dimer Interface Targeting. J Med Chem. 2021;64:3204–21.

    Article  CAS  Google Scholar 

  35. Nagai H, Hatanaka W, Matsuda M, Kishimura A, Katayama Y, Mori T. Folate receptor-specific cell-cell adhesion by using a folate-modified peptide-based anchor. J Biomater Sci. 2019;30:983–93.

    Article  CAS  Google Scholar 

  36. Kim H, Jo A, Baek S, Lim D, Park S-Y, Cho SK, et al. Synergistically enhanced selective intracellular uptake of anticancer drug carrier comprising folic acid-conjugated hydrogels containing magnetite nanoparticles. Sci Rep. 2017;7:41090.

    Article  CAS  Google Scholar 

  37. Zheng Z, Li Z, Xu C, Guo B, Guo P. Folate-displaying exosome mediated cytosolic delivery of siRNA avoiding endosome trapping. J Control Release Elsevier. 2019;311–312:43–9.

    Article  Google Scholar 

  38. Ledermann JA, Canevari S, Thigpen T. Targeting the folate receptor: diagnostic and therapeutic approaches to personalize cancer treatments. Ann Oncol. 2015;26:2034–43.

    Article  CAS  Google Scholar 

  39. Cal PMSD, Frade RFM, Chudasama V, Cordeiro C, Caddick S, Gois PMP. Targeting cancer cells with folic acid–iminoboronate fluorescent conjugates. Chem Commun. R Soc Chem. 2014;50:5261–63.

  40. Xiong H, Zhang Z-G, Tian X-Q, Sun D-F, Liang Q-C, Zhang Y-J, et al. Inhibition of JAK1, 2/STAT3 Signaling Induces Apoptosis, Cell Cycle Arrest, and Reduces Tumor Cell Invasion in Colorectal Cancer Cells. Neoplasia. 2008;10:287–97.

    Article  CAS  Google Scholar 

  41. Newstead S. Structural basis for recognition and transport of folic acid in mammalian cells. Curr Opin Struct Biol. 2022;74:102353.

    Article  CAS  Google Scholar 

  42. Chen C, Ke J, Zhou XE, Yi W, Brunzelle JS, Li J, et al. Structural basis for molecular recognition of folic acid by folate receptors. Nature. 2013;500:486–9.

    Article  CAS  Google Scholar 

  43. Leamon CP, Low PS. Folate-mediated targeting: from diagnostics to drug and gene delivery. Drug Discov Today. 2001;6:44–51.

    Article  CAS  Google Scholar 

  44. Jaime-Ramirez AC, McMichael E, Kondadasula S, Skinner CC, Mundy-Bosse BL, Luedke E, et al. NK Cell-Mediated Antitumor Effects of a Folate-Conjugated Immunoglobulin Are Enhanced by Cytokines. Cancer Immunol Res. 2016;4:323–36.

    Article  CAS  Google Scholar 

  45. Lu Y, Sega E, Low PS. Folate receptor-targeted immunotherapy: Induction of humoral and cellular immunity against hapten-decorated cancer cells. Int J Cancer. 2005;116:710–9.

    Article  CAS  Google Scholar 

  46. Park HS, Kim J, Cho MY, Cho YJ, Suh YD, Nam SH, et al. Effectual Labeling of Natural Killer Cells with Upconverting Nanoparticles by Electroporation for in Vivo Tracking and Biodistribution Assessment. ACS Appl Mater Interfaces. 2020;12:49362–70.

    Article  CAS  Google Scholar 

  47. Sawaengsri H, Wang J, Reginaldo C, Steluti J, Wu D, Meydani SN, et al. High folic acid intake reduces natural killer cell cytotoxicity in aged mice. J Nutr Biochem. 2016;30:102–7.

    Article  CAS  Google Scholar 

  48. RodrÍguez E, Schetters STT, van Kooyk Y. The tumour glyco-code as a novel immune checkpoint for immunotherapy. Nat Rev Immunol. 2018;18:204–11.

    Article  Google Scholar 

  49. Guo R, Yao Y, Cheng G, Wang SH, Li Y, Shen M, et al. Synthesis of glycoconjugated poly(amindoamine) dendrimers for targeting human liver cancer cells. RSC Adv. 2012;2:99–102.

    Article  CAS  Google Scholar 

  50. Chen J, Liu T, Gao J, Gao L, Zhou L, Cai M, et al. Variation in Carbohydrates between Cancer and Normal Cell Membranes Revealed by Super-Resolution Fluorescence Imaging. Adv Sci. 2016;3:1600270.

    Article  Google Scholar 

  51. Méndez-Huergo SP, Blidner AG, Rabinovich GA. Galectins: emerging regulatory checkpoints linking tumor immunity and angiogenesis. Curr Opin Immunol. 2017;45:8–15.

    Article  Google Scholar 

  52. Ilarregui JM, Croci DO, Bianco GA, Toscano MA, Salatino M, Vermeulen ME, et al. Tolerogenic signals delivered by dendritic cells to T cells through a galectin-1-driven immunoregulatory circuit involving interleukin 27 and interleukin 10. Nat Immunol. 2009;10:981–91.

    Article  CAS  Google Scholar 

  53. Wang W, Guo H, Geng J, Zheng X, Wei H, Sun R, et al. Tumor-released Galectin-3, a Soluble Inhibitory Ligand of Human NKp30, Plays an Important Role in Tumor Escape from NK Cell Attack. J Biol Chem. 2014;289:33311–9.

    Article  CAS  Google Scholar 

  54. Suzuki Y, Sutoh M, Hatakeyama S, Mori K, Yamamoto H, Koie T, et al. MUC1 carrying core 2 O-glycans functions as a molecular shield against NK cell attack, promoting bladder tumor metastasis. Int J Oncol. 2012;40:1831–8.

    CAS  Google Scholar 

  55. Chang C-H, Qiu J, O’Sullivan D, Buck MD, Noguchi T, Curtis JD, et al. Metabolic Competition in the Tumor Microenvironment Is a Driver of Cancer Progression. Cell. 2015;162:1229–41.

    Article  CAS  Google Scholar 

  56. Modenutti CP, Capurro JIB, Di Lella S, Martí MA. The Structural Biology of Galectin-Ligand Recognition: Current Advances in Modeling Tools, Protein Engineering, and Inhibitor Design. Front Chem. 2019;7:1–14.

    Article  Google Scholar 

  57. Arifuzzaman M, Hamza A, Zannat SS, Fahad R, Rahman A, Hosen SMZ, et al. Targeting galectin-3 by natural glycosides: a computational approach. Netw Model Anal Heal Informatics Bioinforma. 2020;9:14.

    Article  Google Scholar 

  58. Stannard KA, Collins PM, Ito K, Sullivan EM, Scott SA, Gabutero E, et al. Galectin inhibitory disaccharides promote tumour immunity in a breast cancer model. Cancer Lett. 2010;299:95–110.

    Article  CAS  Google Scholar 

  59. Iacobazzi RM, Porcelli L, Lopedota AA, Laquintana V, Lopalco A, Cutrignelli A, et al. Targeting human liver cancer cells with lactobionic acid-G(4)-PAMAM-FITC sorafenib loaded dendrimers. Int J Pharm. 2017;528:485–97.

    Article  CAS  Google Scholar 

  60. Fu F, Wu Y, Zhu J, Wen S, Shen M, Shi X. Multifunctional Lactobionic Acid-Modified Dendrimers for Targeted Drug Delivery to Liver Cancer Cells: Investigating the Role Played by PEG Spacer. ACS Appl Mater Interfaces. 2014;6:16416–25.

    Article  CAS  Google Scholar 

  61. Bansal D, Yadav K, Pandey V, Ganeshpurkar A, Agnihotri A, Dubey N. Lactobionic acid coupled liposomes: an innovative strategy for targeting hepatocellular carcinoma. Drug Deliv. 2016;23:140–6.

    Article  CAS  Google Scholar 

  62. Lim J, Sari-Ak D, Bagga T. Siglecs as Therapeutic Targets in Cancer. Biol. 2021;10(11):1178.

    Article  CAS  Google Scholar 

  63. Xiao H, Woods EC, Vukojicic P, Bertozzi CR. Precision glycocalyx editing as a strategy for cancer immunotherapy. Proc Natl Acad Sci U S A. 2016;113:10304–9.

    Article  CAS  Google Scholar 

  64. Hartmann J, Van TT, Kaudeer J, Oberle K, Herrmann J, Quagliano I, et al. The Stalk Domain and the Glycosylation Status of the Activating Natural Killer Cell Receptor NKp30 Are Important for Ligand Binding. J Biol Chem. 2012;287:31527–39.

    Article  CAS  Google Scholar 

  65. Patel KR, Nott JD, Barb AW. Primary human natural killer cells retain proinflammatory IgG1 at the cell surface and express CD16a glycoforms with donor-dependent variability. Mol Cell Proteomics. 2019;18:2178–90.

    Article  Google Scholar 

  66. Figuereido I, Paiotta A, Dal Magro R, Tinelli F, Corti R, Re F, et al. A New Approach for Glyco-Functionalization of Collagen-Based Biomaterials. Int J Mol Sci. 2019;20(7):1747.

    Article  Google Scholar 

  67. Zhu Q, Shen Z, Chiodo F, Nicolardi S, Molinaro A, Silipo A, Yu B. Chemical synthesis of glycans up to a 128-mer relevant to the O-antigen of Bacteroides vulgatus. Nat Commun. 2020;11:4142.

  68. Li C, Wang L-X. Chemoenzymatic Methods for the Synthesis of Glycoproteins. Chem Rev. 2018;118:8359–413.

    Article  CAS  Google Scholar 

  69. Joseph AA, Pardo-Vargas A, Seeberger PH. Total Synthesis of Polysaccharides by Automated Glycan Assembly. J Am Chem Soc. 2020;142:8561–4.

    Article  CAS  Google Scholar 

  70. Wang Y-C, Lin V, Loring JF, Peterson SE. The ‘sweet’ spot of cellular pluripotency: protein glycosylation in human pluripotent stem cells and its applications in regenerative medicine. Expert Opin Biol Ther. 2015;15:679–87.

    Article  CAS  Google Scholar 

  71. Beatson R, Tajadura-Ortega V, Achkova D, Picco G, Tsourouktsoglou T-D, Klausing S, et al. The mucin MUC1 modulates the tumor immunological microenvironment through engagement of the lectin Siglec-9. Nat Immunol. 2016;17:1273–81.

    Article  CAS  Google Scholar 

  72. Kim B, Sun R, Oh W, Kim AMJ, Schwarz JR, Lim S-O. Saccharide analog, 2-deoxy-d-glucose enhances 4–1BB-mediated antitumor immunity via PD-L1 deglycosylation. Mol Carcinog John Wiley & Sons, Ltd. 2020;59:691–700.

    Article  CAS  Google Scholar 

  73. Higai K, Ichikawa A, Matsumoto K. Binding of sialyl Lewis X antigen to lectin-like receptors on NK cells induces cytotoxicity and tyrosine phosphorylation of a 17-kDa protein. Biochim Biophys Acta - Gen Subj. 2006;1760:1355–63.

    Article  CAS  Google Scholar 

  74. Diniz F, Coelho P, Duarte HO, Sarmento B, Reis CA, Gomes J. Glycans as Targets for Drug Delivery in Cancer. Cancers (Basel). 2022;14:1–19.

    Article  Google Scholar 

  75. Nardy AFFR, Freire-de-Lima L, Freire-de-Lima CG, Morrot A. The sweet side of immune evasion: Role of Glycans in the Mechanisms of Cancer Progression. Front Oncol. 2016;6:1–7.

    Article  Google Scholar 

  76. Pan Q, Lv Y, Williams GR, Tao L, Yang H, Li H, et al. Lactobionic acid and carboxymethyl chitosan functionalized graphene oxide nanocomposites as targeted anticancer drug delivery systems. Carbohydr Polym. 2016;151:812–20.

    Article  CAS  Google Scholar 

  77. Sindhu RK, Goyal A, Das J, Neha, Choden S, Kumar P. Immunomodulatory potential of polysaccharides derived from plants and microbes: A narrative review. Carbohydr Polym Technol Appl. 2021;2:100044.

  78. Varki A. Glycan-based interactions involving vertebrate sialic-acid-recognizing proteins. Nature. 2007;446:1023–9.

    Article  CAS  Google Scholar 

  79. Rodriguez E, Boelaars K, Brown K, Eveline Li RJ, Kruijssen L, Bruijns SCM, et al. Sialic acids in pancreatic cancer cells drive tumour-associated macrophage differentiation via the Siglec receptors Siglec-7 and Siglec-9. Nat Commun. 2021;12:1270.

    Article  CAS  Google Scholar 

  80. Crocker PR, Paulson JC, Varki A. Siglecs and their roles in the immune system. Nat Rev Immunol. 2007;7:255–66.

    Article  CAS  Google Scholar 

  81. Zheng Y, Ma X, Su D, Zhang Y, Yu L, Jiang F, et al. The Roles of Siglec7 and Siglec9 on Natural Killer Cells in Virus Infection and Tumour Progression. J Immunol Res. 2020;2020:6243819.

    Article  Google Scholar 

  82. Movsisyan LD, Macauley MS. Structural advances of Siglecs: insight into synthetic glycan ligands for immunomodulation. Org Biomol Chem. 2020;18:5784–97.

    Article  CAS  Google Scholar 

  83. Munkley J. Aberrant Sialylation in Cancer: Therapeutic Opportunities. Cancers. 2022;14(17):4248.

    Article  CAS  Google Scholar 

  84. Munkley J, Scott E. Targeting Aberrant Sialylation to Treat Cancer. Med. 2019;6(4):102.

    CAS  Google Scholar 

  85. Jiang K-Y, Qi L-L, Kang F-B, Wang L. The intriguing roles of Siglec family members in the tumor microenvironment. Biomark Res. 2022;10:22.

    Article  Google Scholar 

  86. Hudak JE, Canham SM, Bertozzi CR. Glycocalyx engineering reveals a Siglec-based mechanism for NK cell immunoevasion. Nat Chem Biol. 2014;10:69–75.

    Article  CAS  Google Scholar 

  87. Daly J, Carlsten M, O’Dwyer M. Sugar Free: Novel Immunotherapeutic Approaches Targeting Siglecs and Sialic Acids to Enhance Natural Killer Cell Cytotoxicity Against Cancer. Front Immunol. 2019;10:1047.

    Article  CAS  Google Scholar 

  88. Büll C, Nason R, Sun L, Van Coillie J, Madriz Sørensen D, Moons SJ, et al. Probing the binding specificities of human Siglecs by cell-based glycan arrays. Proc Natl Acad Sci. 2021;118: e2026102118.

    Article  Google Scholar 

  89. Munkley J. The Role of Sialyl-Tn in Cancer. Int J Mol Sci. 2016;17(3):275.

    Article  Google Scholar 

  90. Dharmadhikari G, Stolz K, Hauke M, Morgan NG, Varki A, de Koning E, et al. Siglec-7 restores β-cell function and survival and reduces inflammation in pancreatic islets from patients with diabetes. Sci Rep. 2017;7:45319.

    Article  CAS  Google Scholar 

  91. Gicheva N, Macauley MS, Arlian BM, Paulson JC, Kawasaki N. Siglec-F is a novel intestinal M cell marker. Biochem Biophys Res Commun. 2016;479:1–4.

    Article  CAS  Google Scholar 

  92. Dobie C, Skropeta D. Insights into the role of sialylation in cancer progression and metastasis. Br J Cancer. 2021;124:76–90.

    Article  CAS  Google Scholar 

  93. Nicoll G, Avril T, Lock K, Furukawa K, Bovin N, Crocker PR. Ganglioside GD3 expression on target cells can modulate NK cell cytotoxicity via siglec-7-dependent and -independent mechanisms. Eur J Immunol. 2003;33:1642–8.

    Article  CAS  Google Scholar 

  94. Zhang X, Xu G, Gadora K, Cheng H, Peng J, Ma Y, et al. Dual-sensitive chitosan derivative micelles for site-specific drug release in the treatment of chicken coccidiosis. RSC Adv. 2018;8:14515–26.

    Article  CAS  Google Scholar 

  95. Guo H, Li H, Gao J, Zhao G, Ling L, Wang B, et al. Phenylboronic acid-based amphiphilic glycopolymeric nanocarriers for in vivo insulin delivery. Polym Chem. 2016;7:3189–99.

    Article  CAS  Google Scholar 

  96. Wang J, Liu LG, Jiao WQ, Yang H, Liu J, Liu D. Phenylboronic acid-conjugated chitosan nanoparticles for high loading and efficient delivery of curcumin. Carbohydr Polym. 2021;256: 117497.

    Article  CAS  Google Scholar 

  97. Ma R, Yang H, Li Z, Liu G, Sun X, Liu X, et al. Phenylboronic Acid-Based Complex Micelles with Enhanced Glucose-Responsiveness at Physiological pH by Complexation with Glycopolymer. Biomacromol. 2012;13:3409–17.

    Article  CAS  Google Scholar 

  98. Huang L, Zhang Q, Dai L, Shen X, Chen W, Cai K. Phenylboronic acid-modified hollow silica nanoparticles for dual-responsive delivery of doxorubicin for targeted tumor therapy. Regen Biomater. 2017;4:111–24.

    CAS  Google Scholar 

  99. Khan T, Igarashi K, Tanabe A, Miyazawa T, Fukushima S, Miura Y, et al. Structural Control of Boronic Acid Ligands Enhances Intratumoral Targeting of Sialic Acid to Eradicate Cancer Stem-like Cells. ACS Appl Bio Mater. 2020;3:5030–9.

    Article  CAS  Google Scholar 

  100. Chen H, Lee M, Lee J, Kim JH, Gal YS, Hwang YH, et al. Formation and characterization of self-assembled phenylboronic acid derivative monolayers toward developing monosaccharide sensing-interface. Sensors. 2007;7:1480–95.

    Article  CAS  Google Scholar 

  101. Xu Y, Huang Y, Lu W, Liu S, Xiao Y, Yu J. 4-Carboxyphenylboronic acid-decorated, redox-sensitive rod-shaped nano-micelles fabricated through co-assembling strategy for active targeting and synergistic co-delivery of camptothecin and gemcitabine. Eur J Pharm Biopharm. 2019;144:193–206.

    Article  CAS  Google Scholar 

  102. Lee JY, Chung SJ, Cho HJ, Kim DD. Phenylboronic acid-decorated chondroitin sulfate A-based theranostic nanoparticles for enhanced tumor targeting and penetration. Adv Funct Mater. 2015;25:3705–17.

    Article  CAS  Google Scholar 

  103. Kang BY, Hur W, Kim SM, Kim S, Lee J, Tak E, et al. Phenylboronic acid conjugated to doxorubicin nanocomplexes as an anti-cancer drug delivery system in hepatocellular carcinoma. Nanomedicine Nanotechnology Biol Med. 2021;34:102389.

    Article  CAS  Google Scholar 

  104. Jeong JY, Hong E-H, Lee SY, Lee J-Y, Song J-H, Ko S-H, et al. Boronic acid-tethered amphiphilic hyaluronic acid derivative-based nanoassemblies for tumor targeting and penetration. Acta Biomater. 2017;53:414–26.

    Article  CAS  Google Scholar 

  105. Kim J, Lee YM, Kim H, Park D, Kim J, Kim WJ. Phenylboronic acid-sugar grafted polymer architecture as a dual stimuli-responsive gene carrier for targeted anti-angiogenic tumor therapy. Biomaterials. 2016;75:102–11.

    Article  CAS  Google Scholar 

  106. Deshayes S, Cabral H, Ishii T, Miura Y, Kobayashi S, Yamashita T, et al. Phenylboronic acid-installed polymeric micelles for targeting sialylated epitopes in solid tumors. J Am Chem Soc. 2013;135:15501–7.

    Article  CAS  Google Scholar 

  107. Jp L, Jo E. Polyol Complexes and Structure of the Benzeneboronate Ion. J Org Chem. 1959;24:769–74.

    Article  Google Scholar 

  108. Zhang Q, Liu Y, Fei Y, Xie J, Zhao X, Zhong Z, et al. Phenylboronic Acid-Functionalized Copolypeptides: Facile Synthesis and Responsive Dual Anticancer Drug Release. Biomacromol. 2022;23:2989–98.

    Article  CAS  Google Scholar 

  109. Whited J, Rama CK, Sun X-L. Synthesis and Evaluation of Protein-Phenylboronic Acid Conjugates as Lectin Mimetics. ACS Omega. 2018;3:13467–73.

    Article  CAS  Google Scholar 

  110. Shao M, Qi Y, Sui D, Xu F-J. Phenylboronic acid-functionalized polyaminoglycoside as an effective CRISPR/Cas9 delivery system. Biomater Sci. 2021;9:7104–14.

    Article  CAS  Google Scholar 

  111. Debiais M, Vasseur J-J, Smietana M. Applications of the Reversible Boronic Acids/Boronate Switch to Nucleic Acids. Chem Rec. 2022;22:e202200085.

    Article  CAS  Google Scholar 

  112. Tang X, Sheng Q, Xu C, Li M, Rao J, Wang X, et al. pH/ATP cascade-responsive nano-courier with efficient tumor targeting and siRNA unloading for photothermal-immunotherapy. Nano Today. 2021;37:101083.

    Article  CAS  Google Scholar 

  113. Yoshinaga N, Uchida S, Dirisala A, Naito M, Osada K, Cabral H, et al. mRNA loading into ATP-responsive polyplex micelles with optimal density of phenylboronate ester crosslinking to balance robustness in the biological milieu and intracellular translational efficiency. J Control Release. 2021;330:317–28.

    Article  CAS  Google Scholar 

  114. van Meer G, Voelker DR, Feigenson GW. Membrane lipids: where they are and how they behave. Nat Rev Mol Cell Biol. 2008;9:112–24.

    Article  Google Scholar 

  115. Luchini A, Vitiello G. Mimicking the Mammalian Plasma Membrane: An Overview of Lipid Membrane Models for Biophysical Studies. Biomimetics. 2020;6:3.

  116. Vance JE. Phospholipid Synthesis and Transport in Mammalian Cells. Traffic. 2015;16:1–18.

    Article  CAS  Google Scholar 

  117. Triffo SB, Huang HH, Smith AW, Chou ET, Groves JT. Monitoring Lipid Anchor Organization in Cell Membranes by PIE-FCCS. J Am Chem Soc. 2012;134:10833–42.

    Article  CAS  Google Scholar 

  118. Noiri M, Asawa K, Okada N, Kodama T, Murayama Y, Inoue Y, et al. Modification of human MSC surface with oligopeptide-PEG-lipids for selective binding to activated endothelium. J Biomed Mater Res Part A. 2019;107:1779–92.

    Article  CAS  Google Scholar 

  119. Vabbilisetty P, Boron M, Nie H, Ozhegov E, Sun X-L. Chemical Reactive Anchoring Lipids with Different Performance for Cell Surface Re-engineering Application. ACS Omega. 2018;3:1589–99.

    Article  CAS  Google Scholar 

  120. Weber RJ, Liang SI, Selden NS, Desai TA, Gartner ZJ. Efficient Targeting of Fatty-Acid Modified Oligonucleotides to Live Cell Membranes through Stepwise Assembly. Biomacromol. 2014;15:4621–6.

    Article  CAS  Google Scholar 

  121. Li X, Feng K, Li L, Yang L, Pan X, Yazd HS, et al. Lipid–oligonucleotide conjugates for bioapplications. Natl Sci Rev. 2020;7:1933–53.

    Article  CAS  Google Scholar 

  122. Jin C, He J, Zou J, Xuan W, Fu T, Wang R, et al. Phosphorylated lipid-conjugated oligonucleotide selectively anchors on cell membranes with high alkaline phosphatase expression. Nat Commun. 2019;10:2704.

    Article  Google Scholar 

  123. Kato K, Itoh C, Yasukouchi T, Nagamune T. Rapid Protein Anchoring into the Membranes of Mammalian Cells Using Oleyl Chain and Poly(ethylene glycol) Derivatives. Biotechnol Prog. 2004;20:897–904.

    Article  CAS  Google Scholar 

  124. Chanana M, Gliozzi A, Diaspro A, Chodnevskaja I, Huewel S, Moskalenko V, et al. Interaction of Polyelectrolytes and Their Composites with Living Cells. Nano Lett. 2005;5:2605–12.

    Article  CAS  Google Scholar 

  125. 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;27:5828–35.

    Article  CAS  Google Scholar 

  126. Choksakulnimitr S, Masuda S, Tokuda H, Takakura Y, Hashida M. In vitro cytotoxicity of macromolecules in different cell culture systems. J Control Release. 1995;34:233–41.

    Article  CAS  Google Scholar 

  127. Hong D, Yang SH. Cationic Polymers for Coating Living Cells. Macromol Res. 2018;26:1185–92.

    Article  CAS  Google Scholar 

  128. Peng H, Chelvarajan L, Donahue R, Gottipati A, Cahall CF, Davis KA, et al. Polymer Cell Surface Coating Enhances Mesenchymal Stem Cell Retention and Cardiac Protection. ACS Appl Bio Mater. 2021;4:1655–67.

    Article  CAS  Google Scholar 

  129. Kampf N. The use of polymers for coating of cells. Polym Adv Technol. 2002;13:895–904.

    Article  Google Scholar 

  130. Contreras JL, Xie D, Mays J, Smyth CA, Eckstein C, Rahemtulla FG, et al. A novel approach to xenotransplantation combining surface engineering and genetic modification of isolated adult porcine islets. Surgery. 2004;136:537–47.

    Article  Google Scholar 

  131. Teramura Y, Kaneda Y, Totani T, Iwata H. Behavior of synthetic polymers immobilized on a cell membrane. Biomaterials. 2008;29:1345–55.

    Article  CAS  Google Scholar 

  132. Li X, Sun H, Li H, Hu C, Luo Y, Shi X, et al. Multi-Responsive Biodegradable Cationic Nanogels for Highly Efficient Treatment of Tumors. Adv Funct Mater. 2021;31:2100227.

    Article  CAS  Google Scholar 

  133. Aleknavičienė I, Talaikis M, Budvytyte R, Valincius G. The Impact of an Anchoring Layer on the Formation of Tethered Bilayer Lipid Membranes on Silver Substrates. Molecules. 2021;26(22):6878.

    Article  Google Scholar 

  134. Tatsumi K, Ohashi K, Teramura Y, Utoh R, Kanegae K, Watanabe N, et al. The non-invasive cell surface modification of hepatocytes with PEG-lipid derivatives. Biomaterials. 2012;33:821–8.

    Article  CAS  Google Scholar 

  135. Selden NS, Todhunter ME, Jee NY, Liu JS, Broaders KE, Gartner ZJ. Chemically Programmed Cell Adhesion with Membrane-Anchored Oligonucleotides. J Am Chem Soc. 2012;134:765–8.

    Article  CAS  Google Scholar 

  136. Lee DY, Lim KS, Valencia GM, Jung M, Bull DA, Won Y-W. One-Step Method for Instant Generation of Advanced Allogeneic NK Cells. Adv Sci. 2018;5:1800447.

    Article  Google Scholar 

  137. Bagheri Y, Chedid S, Shafiei F, Zhao B, You M. A quantitative assessment of the dynamic modification of lipid–DNA probes on live cell membranes. Chem Sci. 2019;10:11030–40.

    Article  CAS  Google Scholar 

  138. Uvyn A, De Coen R, Gruijs M, Tuk CW, De Vrieze J, van Egmond M, et al. Efficient Innate Immune Killing of Cancer Cells Triggered by Cell-Surface Anchoring of Multivalent Antibody-Recruiting Polymers. Angew Chemie Int Ed. 2019;58:12988–93.

    Article  CAS  Google Scholar 

  139. Rudd AK, Valls Cuevas JM, Devaraj NK. SNAP-Tag-Reactive Lipid Anchors Enable Targeted and Spatiotemporally Controlled Localization of Proteins to Phospholipid Membranes. J Am Chem Soc. 2015;137:4884–7.

    Article  CAS  Google Scholar 

  140. Yoshihara A, Watanabe S, Goel I, Ishihara K, Ekdahl KN, Nilsson B, et al. Promotion of cell membrane fusion by cell-cell attachment through cell surface modification with functional peptide-PEG-lipids. Biomaterials. 2020;253:120113.

    Article  CAS  Google Scholar 

  141. Nakanishi J, Kikuchi Y, Takarada T, Nakayama H, Yamaguchi K, Maeda M. Photoactivation of a Substrate for Cell Adhesion under Standard Fluorescence Microscopes. J Am Chem Soc. 2004;126:16314–5.

    Article  CAS  Google Scholar 

  142. Ohmuro-Matsuyama Y, Tatsu Y. Photocontrolled Cell Adhesion on a Surface Functionalized with a Caged Arginine-Glycine-Aspartate Peptide. Angew Chemie Int Ed. 2008;47:7527–9.

    Article  CAS  Google Scholar 

  143. Lee TT, García JR, Paez JI, Singh A, Phelps EA, Weis S, et al. Light-triggered in vivo activation of adhesive peptides regulates cell adhesion, inflammation and vascularization of biomaterials. Nat Mater. 2015;14:352–60.

    Article  CAS  Google Scholar 

  144. Kato K, Umezawa K, Funeriu DP, Miyake M, Miyake J, Nagamune T. Immobilized culture of nonadherent cells on an oleyl poly(ethylene glycol) ether-modified surface. Biotechniques. 2003;35:1014–21.

    Article  CAS  Google Scholar 

  145. Kida A, Iijima M, Niimi T, Maturana AD, Yoshimoto N, Kuroda S. Cell Surface-Fluorescence Immunosorbent Assay for Real-Time Detection of Hybridomas with Efficient Antibody Secretion at the Single-Cell Level. Anal Chem. 2013;85:1753–9.

    Article  CAS  Google Scholar 

  146. Jarzębska NT, Yamaguchi S, Izuta S, Kosaka T, Yamahira S, Nagamune T, et al. Photo-responsive materials with strong cell trapping ability for light-guided manipulation of nonadherent cells. Biomater Sci. 2019;7:4514–8.

    Article  Google Scholar 

  147. Yamahira S, Misawa R, Kosaka T, Tan M, Izuta S, Yamashita H, et al. Photoactivatable Materials for Versatile Single-Cell Patterning Based on the Photocaging of Cell-Anchoring Moieties through Lipid Self-Assembly. J Am Chem Soc. 2022;144:13154–62.

    Article  CAS  Google Scholar 

  148. Armstrong JPK, Perriman AW. Strategies for cell membrane functionalization. Exp Biol Med. 2016;241:1098–106.

    Article  CAS  Google Scholar 

  149. von Nickisch-Rosenegk M, Teschke T, Bier FF. Construction of an artificial cell membrane anchor using DARC as a fitting for artificial extracellular functionalities of eukaryotic cells. J Nanobiotechnology. 2012;10:1.

    Article  Google Scholar 

  150. Teramura Y, Iwata H. Cell surface modification with polymers for biomedical studies. Soft Matter. 2010;6:1081–91.

    Article  CAS  Google Scholar 

  151. Gasperini L, Mano JF, Reis RL. Natural polymers for the microencapsulation of cells. J R Soc Interface. 2014;11:20140817.

    Article  Google Scholar 

  152. Wilson JT, Cui W, Kozlovskaya V, Kharlampieva E, Pan D, Qu Z, et al. Cell Surface Engineering with Polyelectrolyte Multilayer Thin Films. J Am Chem Soc. 2011;133:7054–64.

    Article  CAS  Google Scholar 

  153. Li X, Hetjens L, Wolter N, Li H, Shi X, Pich A. Charge-reversible and biodegradable chitosan-based microgels for lysozyme-triggered release of vancomycin. J Adv Res. 2023;43:87–96.

    Article  CAS  Google Scholar 

  154. Ribeiro RDC, Pal D, Jamieson D, Rankin KS, Benning M, Dalgarno KW, et al. Temporary Single-Cell Coating for Bioprocessing with a Cationic Polymer. ACS Appl Mater Interfaces. 2017;9:12967–74.

    Article  CAS  Google Scholar 

  155. Dzamukova MR, Zamaleeva AI, Ishmuchametova DG, Osin YN, Kiyasov AP, Nurgaliev DK, et al. A Direct Technique for Magnetic Functionalization of Living Human Cells. Langmuir. 2011;27:14386–93.

    Article  CAS  Google Scholar 

  156. Zhang P, Liu Y, Xia J, Wang Z, Kirkland B, Guan J. Top-Down Fabrication of Polyelectrolyte-Thermoplastic Hybrid Microparticles for Unidirectional Drug Delivery to Single Cells. Adv Healthc Mater. 2013;2:540–5.

    Article  CAS  Google Scholar 

  157. Wytrwal-Sarna M, Knobloch P, Lasota S, Michalik M, Nowakowska M, Kepczynski M. Effect of polycation nanostructures on cell membrane permeability and toxicity. Environ Sci Nano. 2022;9:702–13.

    Article  CAS  Google Scholar 

  158. Farnoud AM, Nazemidashtarjandi S. Emerging investigator series: interactions of engineered nanomaterials with the cell plasma membrane; what have we learned from membrane models? Environ Sci Nano. 2019;6:13–40.

    Article  CAS  Google Scholar 

  159. Kepczynski M, Jamróz D, Wytrwal M, Bednar J, Rzad E, Nowakowska M. Interactions of a Hydrophobically Modified Polycation with Zwitterionic Lipid Membranes. Langmuir. 2012;28:676–88.

    Article  CAS  Google Scholar 

  160. Ames E, Murphy WJ. Advantages and clinical applications of natural killer cells in cancer immunotherapy. Cancer Immunol Immunother. 2014;63:21–8.

    Article  Google Scholar 

  161. Xue Y, O’Mara ML, Surawski PPT, Trau M, Mark AE. Effect of Poly(ethylene glycol) (PEG) Spacers on the Conformational Properties of Small Peptides: A Molecular Dynamics Study. Langmuir. 2011;27:296–303.

    Article  CAS  Google Scholar 

  162. Huang Q, Li D, Kang A, An W, Fan B, Ma X, et al. PEG as a spacer arm markedly increases the immunogenicity of meningococcal group Y polysaccharide conjugate vaccine. J Control Release. 2013;172:382–9.

    Article  CAS  Google Scholar 

  163. Li N, Wang C, Georgiev MI, Bajpai VK, Tundis R, Simal-Gandara J, et al. Advances in dietary polysaccharides as anticancer agents: Structure-activity relationship. Trends Food Sci Technol. 2021;111:360–77.

    Article  CAS  Google Scholar 

  164. Yu Z, Sun Q, Liu J, Zhang X, Song G, Wang G, et al. Polysaccharide from Rhizopus nigricans inhibits the invasion and metastasis of colorectal cancer. Biomed Pharmacother. 2018;103:738–45.

    Article  CAS  Google Scholar 

  165. Wu Y, Li F, Zhang X, Li Z, Zhang Q, Wang W, et al. Tumor microenvironment-responsive PEGylated heparin-pyropheophorbide-a nanoconjugates for photodynamic therapy. Carbohydr Polym. 2021;255:117490.

    Article  CAS  Google Scholar 

  166. Lin WJ, Lee WC. Polysaccharide-modified nanoparticles with intelligent CD44 receptor targeting ability for gene delivery. Int J Nanomedicine. 2018;13:3989–4002.

    Article  CAS  Google Scholar 

  167. Ferreira SS, Passos CP, Madureira P, Vilanova M, Coimbra MA. Structure–function relationships of immunostimulatory polysaccharides: A review. Carbohydr Polym. 2015;132:378–96.

    Article  CAS  Google Scholar 

  168. Chen F, Huang G. Preparation and immunological activity of polysaccharides and their derivatives. Int J Biol Macromol. 2018;112:211–6.

    Article  CAS  Google Scholar 

  169. Seidi F, Jenjob R, Phakkeeree T, Crespy D. Saccharides, oligosaccharides, and polysaccharides nanoparticles for biomedical applications. J Control Release. 2018;284:188–212.

    Article  CAS  Google Scholar 

  170. Jin Y, Mu Y, Zhang S, Li P, Wang F. Preparation and evaluation of the adjuvant effect of curdlan sulfate in improving the efficacy of dendritic cell-based vaccine for antitumor immunotherapy. Int J Biol Macromol. 2020;146:273–84.

    Article  CAS  Google Scholar 

  171. Li Y, Wang X, Ma X, Liu C, Wu J, Sun C. Natural Polysaccharides and Their Derivates: A Promising Natural Adjuvant for Tumor Immunotherapy. Front Pharmacol. 2021;12:1–18.

    Google Scholar 

  172. Wang H-Y, Jia H-R, Lu X, Chen B, Zhou G, He N, et al. Imaging plasma membranes without cellular internalization: multisite membrane anchoring reagents based on glycol chitosan derivatives. J Mater Chem B The Royal Society of Chemistry. 2015 3:6165–73.

    Article  CAS  Google Scholar 

  173. Jia H-R, Wang H-Y, Yu Z-W, Chen Z, Wu F-G. Long-Time Plasma Membrane Imaging Based on a Two-Step Synergistic Cell Surface Modification Strategy. Bioconjug Chem. 2016;27:782–9.

    Article  CAS  Google Scholar 

  174. Lee DY, Cha BH, Jung M, Kim AS, Bull DA, Won YW. Cell surface engineering and application in cell delivery to heart diseases. J Biol Eng. 2018;12:1–11.

    Article  Google Scholar 

  175. Huang ML, Smith RAA, Trieger GW, Godula K. Glycocalyx Remodeling with Proteoglycan Mimetics Promotes Neural Specification in Embryonic Stem Cells. J Am Chem Soc. 2014;136:10565–8.

    Article  CAS  Google Scholar 

  176. Liu Q, Lyu Z, Yu Y, Zhao Z-A, Hu S, Yuan L, et al. Synthetic Glycopolymers for Highly Efficient Differentiation of Embryonic Stem Cells into Neurons: Lipo- or Not? ACS Appl Mater Interfaces. 2017;9:11518–27.

    Article  CAS  Google Scholar 

  177. Church DC, Pokorski JK. Cell Engineering with Functional Poly(oxanorbornene) Block Copolymers. Angew Chemie Int Ed. 2020;59:11379–83.

    Article  CAS  Google Scholar 

  178. Esfahani RR, Jun H, Rahmani S, Miller A, Lahann J. Microencapsulation of Live Cells in Synthetic Polymer Capsules. ACS Omega. 2017;2:2839–47.

    Article  CAS  Google Scholar 

  179. De Vos P, Van Straaten JF, Nieuwenhuizen AG, de Groot M, Ploeg RJ, De Haan BJ, et al. Why do microencapsulated islet grafts fail in the absence of fibrotic overgrowth? Diabetes. 1999;48:1381–8.

    Article  Google Scholar 

  180. Wilson JT, Cui W, Chaikof EL. Layer-by-Layer Assembly of a Conformal Nanothin PEG Coating for Intraportal Islet Transplantation. Nano Lett. 2008;8:1940–8.

    Article  CAS  Google Scholar 

  181. Nayyar G, Chu Y, Cairo MS. Overcoming Resistance to Natural Killer Cell Based Immunotherapies for Solid Tumors. Front Oncol. 2019;9:51.

    Article  Google Scholar 

  182. Peng W, Paulson JC. CD22 Ligands on a Natural N-Glycan Scaffold Efficiently Deliver Toxins to B-Lymphoma Cells. J Am Chem Soc. 2017;139:12450–8.

    Article  CAS  Google Scholar 

  183. Ma B, Guan X, Li Y, Shang S, Li J, Tan Z. Protein Glycoengineering: An Approach for Improving Protein Properties. Front Chem. 2020;8:1–14.

    Article  CAS  Google Scholar 

  184. Shi W, Li W, Zhang J, Li T, Song Y, Zeng Y, et al. One-step synthesis of site-specific antibody–drug conjugates by reprograming IgG glycoengineering with LacNAc-based substrates. Acta Pharm Sin B. 2022;12:2417–28.

    Article  CAS  Google Scholar 

  185. Kightlinger W, Warfel KF, DeLisa MP, Jewett MC. Synthetic Glycobiology: Parts, Systems, and Applications. ACS Synth Biol. 2020;9:1534–62.

    Article  CAS  Google Scholar 

  186. Nischan N, Kohler JJ. Advances in cell surface glycoengineering reveal biological function. Glycobiology. 2016;26:1–8.

    Article  CAS  Google Scholar 

  187. Hong S, Yu C, Wang P, Shi Y, Cao W, Cheng B, et al. Glycoengineering of NK Cells with Glycan Ligands of CD22 and Selectins for B-Cell Lymphoma Therapy. Angew Chemie- Int Ed. 2021;60:3603–10.

    Article  CAS  Google Scholar 

  188. Rillahan CD, Macauley MS, Schwartz E, He Y, McBride R, Arlian BM, et al. Disubstituted sialic acid ligands targeting siglecs CD33 and CD22 associated with myeloid leukaemias and B cell lymphomas. Chem Sci. 2014;5:2398–406.

    Article  CAS  Google Scholar 

  189. Wang X, Lang S, Tian Y, Zhang J, Yan X, Fang Z, Weng J, Lu N, Wu X, Li T, Cao H, Li Z, Huang X. Glycoengineering of Natural Killer Cells with CD22 Ligands for Enhanced Anticancer Immunotherapy. ACS Cent. Sci. 2020;6:382–9.

  190. Hong S, Yu C, Rodrigues E, Shi Y, Chen H, Wang P, et al. Modulation of Siglec-7 Signaling Via in Situ-Created High-Affinity cis-Ligands. ACS Cent Sci. 2021;7:1338–46.

    Article  CAS  Google Scholar 

  191. Meng D, Pan H, He W, Jiang X, Liang Z, Zhang X, et al. In Situ Activated NK Cell as Bio-Orthogonal Targeted Live-Cell Nanocarrier Augmented Solid Tumor Immunotherapy. Adv Funct Mater. 2022;2202603:1–15.

    Google Scholar 

  192. Wang X, Luo X, Tian Y, Wu T, Weng J, Li Z, et al. Equipping Natural Killer Cells with Cetuximab through Metabolic Glycoengineering and Bioorthogonal Reaction for Targeted Treatment of KRAS Mutant Colorectal Cancer. ACS Chem Biol. 2021;16:724–30.

    Article  CAS  Google Scholar 

  193. Lee D, Huntoon K, Lux J, Kim BYS, Jiang W. Engineering nanomaterial physical characteristics for cancer immunotherapy. Nat Rev Bioeng. 2023. https://doi.org/10.1038/s44222-023-00047-3.

  194. Yao X, Jovevski JJ, Todd MF, Xu R, Li Y, Wang J, et al. Nanoparticle-Mediated Intracellular Protection of Natural Killer Cells Avoids Cryoinjury and Retains Potent Antitumor Functions. Adv Sci. 2020;7:1902938.

    Article  CAS  Google Scholar 

  195. Park J, Wrzesinski SH, Stern E, Look M, Criscione J, Ragheb R, et al. Combination delivery of TGF-β inhibitor and IL-2 by nanoscale liposomal polymeric gels enhances tumour immunotherapy. Nat Mater. 2012;11:895–905.

    Article  CAS  Google Scholar 

  196. Liu C, Lai H, Chen T. Boosting Natural Killer Cell-Based Cancer Immunotherapy with Selenocystine/Transforming Growth Factor-Beta Inhibitor-Encapsulated Nanoemulsion. ACS Nano. 2020;14:11067–82.

    Article  CAS  Google Scholar 

  197. Lu Y, Luo Q, Jia X, Tam JP, Yang H, Shen Y, et al. Multidisciplinary strategies to enhance therapeutic effects of flavonoids from Epimedii Folium: Integration of herbal medicine, enzyme engineering, and nanotechnology. J Pharm Anal. 2023;13:239–54.

    Article  Google Scholar 

  198. Elsabahy M, Wooley KL. Cytokines as biomarkers of nanoparticle immunotoxicity. Chem Soc Rev. 2013;42:5552–76.

    Article  CAS  Google Scholar 

  199. Kim KS, Kim DH, Kim DH. Recent advances to augment NK cell cancer immunotherapy using nanoparticles. Pharmaceutics. 2021;13:1–14.

    Google Scholar 

  200. Jang E-S, Shin J-H, Ren G, Park M-J, Cheng K, Chen X, et al. The manipulation of natural killer cells to target tumor sites using magnetic nanoparticles. Biomaterials. 2012;33:5584–92.

    Article  CAS  Google Scholar 

  201. Wu L, Zhang F, Wei Z, Li X, Zhao H, Lv H, et al. Magnetic delivery of Fe3O4@polydopamine nanoparticle-loaded natural killer cells suggest a promising anticancer treatment. Biomater Sci. 2018;6:2714–25.

    Article  CAS  Google Scholar 

  202. Burga RA, Khan DH, Agrawal N, Bollard CM, Fernandes R. Designing Magnetically Responsive Biohybrids Composed of Cord Blood-Derived Natural Killer Cells and Iron Oxide Nanoparticles. Bioconjug Chem. 2019;30:552–60.

    Article  CAS  Google Scholar 

  203. Liu Y, Bhattarai P, Dai Z, Chen X. Photothermal therapy and photoacoustic imaging via nanotheranostics in fighting cancer. Chem Soc Rev. 2019;48:2053–108.

    Article  CAS  Google Scholar 

  204. Dai X, Li X, Liu Y, Yan F. Recent advances in nanoparticles-based photothermal therapy synergizing with immune checkpoint blockade therapy. Mater Des. 2022;217:110656.

    Article  CAS  Google Scholar 

  205. Zhang D, Zheng Y, Lin Z, Lan S, Zhang X, Zheng A, et al. Artificial Engineered Natural Killer Cells Combined with Antiheat Endurance as a Powerful Strategy for Enhancing Photothermal-Immunotherapy Efficiency of Solid Tumors. Small. 2019;15:1902636.

    Article  CAS  Google Scholar 

  206. Yim H, Park W, Kim D, Fahmy TM, Na K. A self-assembled polymeric micellar immunomodulator for cancer treatment based on cationic amphiphilic polymers. Biomaterials. 2014;35:9912–9.

    Article  CAS  Google Scholar 

  207. Mulens-Arias V, Rojas JM, Pérez-Yagüe S, Morales MP, Barber DF. Polyethylenimine-coated SPIONs trigger macrophage activation through TLR-4 signaling and ROS production and modulate podosome dynamics. Biomaterials. 2015;52:494–506.

    Article  CAS  Google Scholar 

  208. Kim KS, Han JH, Choi SH, Jung HY, Park JD, An HJ, et al. Cationic Nanoparticle-Mediated Activation of Natural Killer Cells for Effective Cancer Immunotherapy. ACS Appl Mater Interfaces. 2020;12:56731–40.

    Article  CAS  Google Scholar 

  209. Im S, Jang D, Saravanakumar G, Lee J, Kang Y, Lee YM, et al. Harnessing the Formation of Natural Killer-Tumor Cell Immunological Synapses for Enhanced Therapeutic Effect in Solid Tumors. Adv Mater. 2020;32:2000020.

    Article  CAS  Google Scholar 

  210. Elskens JP, Elskens JM, Madder A. Chemical Modification of Aptamers for Increased Binding Affinity in Diagnostic Applications: Current Status and Future Prospects. Int J Mol Sci. 2020;21(12):4522.

    Article  CAS  Google Scholar 

  211. Gauthier L, Morel A, Anceriz N, Rossi B, Blanchard-Alvarez A, Grondin G, et al. Multifunctional Natural Killer Cell Engagers Targeting NKp46 Trigger Protective Tumor Immunity. Cell. 2019;177:1701-1713.e16.

    Article  CAS  Google Scholar 

  212. Sun H, Zhu X, Lu PY, Rosato RR, Tan W, Zu Y. Oligonucleotide Aptamers: New Tools for Targeted Cancer Therapy. Mol Ther - Nucleic Acids. 2014;3:e182.

    Article  CAS  Google Scholar 

  213. Thomas BJ, Porciani D, Burke DH. Cancer immunomodulation using bispecific aptamers. Mol Ther-Nucleic Acids. 2022;27:894–915.

    Article  CAS  Google Scholar 

  214. Yang S, Wen J, Li H, Xu L, Liu Y, Zhao N, et al. Aptamer-Engineered Natural Killer Cells for Cell-Specific Adaptive Immunotherapy. Small. 2019;15:1900903.

    Article  Google Scholar 

  215. Liang S, Slattery MJ, Wagner D, Simon SI, Dong C. Hydrodynamic Shear Rate Regulates Melanoma-Leukocyte Aggregation, Melanoma Adhesion to the Endothelium, and Subsequent Extravasation. Ann Biomed Eng. 2008;36:661–71.

    Article  Google Scholar 

  216. Shi P, Wang X, Davis B, Coyne J, Dong C, Reynolds J, et al. In Situ Synthesis of an Aptamer-Based Polyvalent Antibody Mimic on the Cell Surface for Enhanced Interactions between Immune and Cancer Cells. Angew Chemie Int Ed. 2020;59:11892–7.

    Article  CAS  Google Scholar 

  217. Lambert J.M. Antibody–Drug Conjugates (ADCs): Magic Bullets at Last!. Mol Pharm. 2015;12:1701–2.

  218. Weiner LM, Surana R, Wang S. Monoclonal antibodies: versatile platforms for cancer immunotherapy. Nat Rev Immunol. 2010;10:317–27.

    Article  CAS  Google Scholar 

  219. Weiner LM, Murray JC, Shuptrine CW. Antibody-Based Immunotherapy of Cancer. Cell. 2012;148:1081–4.

    Article  CAS  Google Scholar 

  220. André P, Denis C, Soulas C, Bourbon-Caillet C, Lopez J, Arnoux T, et al. Anti-NKG2A mAb Is a Checkpoint Inhibitor that Promotes Anti-tumor Immunity by Unleashing Both T and NK Cells. Cell Cell Press. 2018;175:1731-1743.e13.

    Google Scholar 

  221. Li HK, Hsiao CW, Yang SH, Yang HP, Wu TS, Lee CY, et al. A novel off-the-shelf trastuzumab-armed NK cell therapy (ACE1702) using antibody-cell-conjugation technology. Cancers (Basel). 2021;13:1–15.

    Google Scholar 

  222. Wang W, Erbe AK, Hank JA, Morris ZS, Sondel PM. NK Cell-Mediated Antibody-Dependent Cellular Cytotoxicity in Cancer Immunotherapy. Front Immunol. 2015;6:368.

    Article  Google Scholar 

  223. Frank MJ, Olsson N, Huang A, Tang SW, Negrin RS, Elias JE, et al. A novel antibody-cell conjugation method to enhance and characterize cytokine-induced killer cells. Cytotherapy. 2020;22:135–43.

    Article  CAS  Google Scholar 

  224. Li J, Chen M, Liu Z, Zhang L, Felding BH, Moremen KW, et al. A Single-Step Chemoenzymatic Reaction for the Construction of Antibody-Cell Conjugates. ACS Cent Sci. 2018;4:1633–41.

    Article  CAS  Google Scholar 

  225. Li D, Brackenridge S, Walters LC, Swanson O, Harlos K, Rozbesky D, et al. Mouse and human antibodies bind HLA-E-leader peptide complexes and enhance NK cell cytotoxicity. Commun Biol. 2022;5:271.

    Article  CAS  Google Scholar 

  226. Gong L, Li Y, Cui K, Chen Y, Hong H, Li J, et al. Nanobody-Engineered Natural Killer Cell Conjugates for Solid Tumor Adoptive Immunotherapy. Small. 2021;17:1–12.

    CAS  Google Scholar 

Download references

Acknowledgements

This research was supported by a National Research Foundation of Korea (NRF) grant funded by the Korea government (MSIT) (2021R1A4A3024237) and a Korean Fund for Regenerative Medicine (KFRM) grant funded by the Korea government (the Ministry of Science and ICT, the Ministry of Health & Welfare) (RS-2022-00070304).

Funding

National Research Foundation of Korea (2021R1A4A3024237).

The Korean Fund for Regenerative Medicine (KFRM) grant (RS-2022–00070304).

Author information

Authors and Affiliations

Authors

Contributions

AKJ: Conceptualization, Design of work, Writing – review & editing draft. SK: Review and editing, Technical inputs and Finalization of draft. KK: Conceptualization, Supervision, Writing – review & editing, Finalization of draft & Funding. All authors read and approved the final manuscript. KK is co-corresponding author.

Corresponding author

Correspondence to Kyobum Kim.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Competing interests

The authors declare that they have no competing financial or personal interests.

Additional information

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Jangid, A.K., Kim, S. & Kim, K. Polymeric biomaterial-inspired cell surface modulation for the development of novel anticancer therapeutics. Biomater Res 27, 59 (2023). https://doi.org/10.1186/s40824-023-00404-8

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s40824-023-00404-8

Keywords