Chapter 9

Cell Surface Enzymatic Engineering-Based Approaches to Improve Cellular Therapies

Ayman F. Abuelela, Kosuke Sakashita and Jasmeen S. Merzaban,    Biological and Environmental Sciences and Engineering, King Abdullah University of Science and Technology, Thuwal, Kingdom of Saudi Arabia

The cell surface represents the interface between the cell and its environment. As such, the cell surface controls cell–cell interactions and functions such as adhesion and migration, and will transfer external cues to regulate processes such as survival, death, and differentiation. Redefining the cell surface by temporarily (or permanently) modifying the molecular landscape of the plasma membrane affects the way in which the cell interacts with its environment and influences the information that is relayed into the cell along downstream signaling pathways. This chapter outlines the role of key enzymes, the glycosyltransferases, in posttranslationally modifying proteins and lipids to fine-tune cells, ability to migrate. These enzymes are critical in controlling the formation of a platform structure, sialyl Lewis x (sLex), on circulating cells that plays a central role in the recognition and recruitment by selectin counter receptors on endothelial cells that line blood vessels of tissues throughout the body. By developing methods to manipulate the activity of these enzymes and hence the cell surface structures that result, treatments can be envisioned that direct the migration of therapeutic cells to specific locations throughout the body and also to inhibit metastasis of detrimental cells such as circulating tumor cells.

Keywords

Glycosyltransferase; fucosyltransferase; sialyltransferase; core 2 N-acetylglucosaminyltransferase; migration; metastasis; stem cells; circulating tumor cell; leukocyte; adhesion

9.1 Introduction

Cell migration is central to many physiological and pathological functions in vivo from immunity to metastasis of cancer cells, to the homing of therapeutic stem cells to a target tissue, and tissue repair. In some cases, it would be desirable to promote migration of particular cells whereas in others inhibition of migration would be preferable. An understanding of the molecules involved in this migration paradigm will permit better development of technologies for therapies.

This chapter will start by introducing the principle of migration of cells to tissues/organs and how this process is controlled by a specific sugar decoration, mainly sialyl Lewis x/a (sLex/a), on cell surface proteins/lipids. This key epitope is created through the activity of specific glycosyltransferases (GTs). When these enzymes are stimulated to become expressed and/or active in a particular cell or are introduced ex vivo, this will result in migration of that cell to specific locations within the body where counter receptors to the epitope, the selectins, are expressed.

The chapter will focus on how controlling the expression and/or activity of these GTs makes it possible to (i) improve the migration of therapeutic cells (i.e., stem cells) to desired locations within the body in order to optimize therapeutic outcomes of disease, and also (ii) inhibit the migration of harmful cells (i.e., circulating tumor cells, CTCs) in order to control metastasis or inflammation. In this chapter, the role of other enzymes in the modification of cell surface architecture related to refining cell function(s) to benefit therapy will also be considered.

9.1.1 Multistep Paradigm of Cellular Migration

Nature has developed an extremely efficient mechanism for the delivery of circulating cells to specific sites in the body in immune processes such as inflammation and injury. An interesting analogy exemplifying cell movement within the blood stream is offered by Ehrhardt et al. [1], which states “Comparable to a human being plunged into a roaring river, a leukocyte (cell in flow) is exposed to high shears within the mainstream of the blood.” In whichever scenario, arrest and exit from the flow represent a particularly complex task for human or cell. The emigration of cells out of flow (such as in the blood circulation) is a sophisticated process controlled by various adhesion molecules. These comprise, in particular, selectins (E, P, and L), chemokines, and integrins, which function in a coordinated multistep cascade (Figure 9.1) [24]. The first essential event in cellular recruitment involves shear-resistant adhesion of flowing cells onto the endothelial surface, this being most efficiently mediated by the selectins and their respective ligands [5,6]. These interactions initially tether the cell in flow to the vessel wall and, in the context of vascular shear flow, cause the tethered cell to roll along the endothelium lining the inside of the vessel. Selectin-dependent tethering and rolling (Step 1) thus decreases cellular velocity below that of the prevailing hemodynamic stream and brings the cell into close physical proximity with the vessel wall. This process facilitates engagement of specific cell-borne chemokine receptors to pertinent chemokines present in the perivascular areas [79], thereby triggering activation through inside-out signal transduction events [1013] leading to increased adhesiveness of integrin family members (Step 2). Adhesive engagements between the activated cell integrins and their cognate endothelial cell counter receptors then lead to arrest and firm adhesion of the cell to the vessel wall (Step 3) and, ultimately, the process of transmigration (Step 4) [14] to reach the extravascular space. This process was first described in the context of leukocyte recruitment to inflammatory sites and has subsequently been applied to describe other biological migration activities such as the movement of circulating hematopoietic stem progenitor cells (HSPCs) to the bone marrow during bone marrow transplants, and in the metastasis of the CTC. The migration of the cell out of circulation and into the target tissue or organ depends on each of these distinct classes of molecules acting sequentially, each step being a prerequisite for the next. In fact, it has been shown in numerous biological scenarios that the selectins, as the most effective mediators [15] of Step 1 interactions, are essential for recruitment of cells altogether and may be the gatekeepers of the multistep paradigm [16]. In fact, it is suggested that the selectins may even be able to substitute for the absence of chemokine receptor signaling events, in the case that chemokines are not present [17,18].

image
Figure 9.1 The multistep paradigm of cell migration highlighting the main interactions that take place between the cell (leukocyte, stem cell, CTC) in flow and the endothelial cells lining the blood vessels of an organ or tissue (such as the bone marrow). Step 1, “tethering and rolling” is mediated by the selectins and their ligands and acts to slow down the cell. Step 2, “integrin activation” is controlled by the expression of chemokines on the surface of the endothelial cells finding their appropriate chemokine receptors on the surface of the cell in flow triggering G-protein-coupled “inside-out” activation of integrins. This activation results in conformational changes of the integrin(s) expressed on the cell in flow and the subsequent high affinity binding of the integrin(s) to their CAM (cellular adhesion molecules) ligand(s) on the endothelium leading to Step 3 “firm adhesion.” The interaction of all of these molecules with their appropriate receptors culminates at Step 4, the transmigration of the cell in flow into the intended tissue or organ. Each step in this process is a prerequisite for the next; however, the initial step, mediated by the selectins and their ligands, is crucial to the overall migration paradigm. Common selectin glycoprotein ligands identified on leukocytes, stem cells, and cancer cells are depicted in the inset; O-glycans are shown as yellow squiggly lines (image) and N-glycans are shown as green hexagons (image).

9.1.1.1 The Selectins

The selectins are a family of three C-type (Ca2+-dependent) lectins expressed exclusively by bone marrow-derived cells and endothelial cells: L-selectin (expressed on Leukocytes and HSPCs), E-selectin (expressed on Endothelial cells), and P-selectin (expressed on Platelets and endothelial cells). Due to their expression on the endothelium, E-selectin and P-selectin are often referred to as the “vascular selectins” whereas L-selectin is termed the “leukocyte selectin.” These molecules are type I transmembrane glycoproteins sharing a similar topology which comprises an amino-terminal lectin-like domain, an epidermal growth factor (EGF)-like domain, a variable number of consensus repeats (CRs; also known as complement regulatory proteins or “sushi” domains), a single membrane spanning domain, and a carboxy-terminal cytoplasmic domain (Figure 9.2A) [4,6]. The lectin domain primarily mediates ligand binding with minimal contribution provided by the EGF-like domain [1922]. The number (and hence length) of sushi domains facilitates the recognition between selectins and their ligands as it allows a selectin to extend beyond the negatively charged cloud of the glycocalyx. The selectins bind to specialized carbohydrate determinants (Figure 9.2B), comprising sialofucosylations containing an α(2,3)-linked sialic acid substitution on galactose, and an α(1,3)-linked fucose modification on N-acetylglucosamine, prototypically displayed as the terminal tetrasaccharide sialyl Lewis x (sLex; or also to its isomer sLea) [16,23]. These determinants, also known as “CD15s,” may be displayed on either a protein scaffold (i.e., a glycoprotein) or a lipid scaffold (i.e., a glycolipid), and are recognized by mAbs such as CSLEX-1, KM93, and HECA-452 [24,25]. Although additional structural modifications, principally involving sulfation, increase the binding affinity of P-selectin and L-selectin to sLex, no known modifications are needed for optimal binding of E-selectin [26,27]. These characteristics of engaging rapidly and with high tensile strength to their ligands help to make the selectins the most important initiators of adhesion under flow. A brief introduction to each of the selectins is provided below.

image
Figure 9.2 Selectins and their carbohydrate ligands. (A) Selectins are type 1 transmembrane cell surface proteins comprised of five different domains: lectin-binding domain, EGF-like domain, varying numbers of CRs (sushi domains), transmembrane domain, and a short cytoplasmic tail. The lectin-binding domain binds the carbohydrate sLea/x containing ligands while the varying sushi domains help to extend the selectin beyond the glycocalyx of the cell surface. (B) GTs involved in the formation of sLea and sLex structures on selectin ligands. Although type 1 and type 2 lactosamine [Galβ1,4GlcNAc or Galβ1,3GlcNAc] structures tend to be common to many glycans, the expression of the GTs responsible for capping the galactose (Gal, image) of the lactosamine with sialic acid (NeuAc, image)—ST—and the terminal N-acetylglucosamine (GlcNAc, image) of the lactosamine with fucose (Fuc, image)—FT—are often correlated with cells that migrate or home to tissues where selectins are expressed.

L-selectin (CD62L) is expressed constitutively by all myeloid cells, virtually all B cells and naïve T cells, a subpopulation of memory T cells, NK cells, and both early and mature hematopoietic cells in the bone marrow. Unlike the other selectins, however, L-selectin is not expressed on endothelial cells [28]. Functionally, L-selectin is critical for the homing of naïve lymphocytes to high endothelial venules (HEVs) of secondary lymphoid organs [3,27]. L-selectin-dependent lymphocyte rolling on HEV requires localization of L-selectin to the tips of the microvilli characteristic of the surfaces of lymphocytes and other leukocytes, allowing for optimal leukocyte–endothelial interactions. Mapping of L-selectin domains by mAbs has revealed that the NH2 terminal nine amino acids are critical for ligand binding. The EGF-like and the two short sushi domains maintain the spatial conformation and are also important for ligand binding [27]. HEV-borne L-selectin ligands are collectively referred to as peripheral node addressins (PNAds) and include the glycoproteins GlyCAM-1 (glycosylation-dependent cell adhesion molecule-1), CD34, MAdCAM-1 (mucosal vascular addressin cell adhesion molecule-1), podocalyxin, Sgp200, endomucin, and endoglycan. All of these PNAds have mucin-like domains that carry O-glycans [27]. Although the primary role of L-selectin is in promoting lymphocyte homing, its expression by circulating neutrophils may facilitate the secondary tethering of these cells to an already rolling neutrophil at an inflammatory site (secondary tethering) [29,30]; this, however, may be a rare event of which the physiological significance has been questioned [31,32]. If L-selectin is to recognize them, these ligands must be modified by various GTs to give specific sialylated, fucosylated, and sulfated oligosaccharides.

P-selectin (CD62P) is a glycoprotein that is constitutively expressed within secretory granules of endothelial cells (Weibel–Palade bodies) and in α-granules of platelets. Upon activation by inflammatory mediators (histamine, thrombin, LPS), secretory vesicles fuse rapidly with the plasma membrane allowing expression of P-selectin on the surface of the platelet or endothelial cells within minutes [6]. Thus, P-selectin is often involved in early leukocyte recruitment during an inflammatory response. P-selectin is also constitutively expressed at low levels on the endothelium of thymus [33], lung and choroid plexus microvessels [34], bone marrow microvasculature [35,36], in postcapillary venules in skin [37], and on peritoneal macrophages [38]. The transcription of P-selectin can be induced by cytokines such as interleukin (IL)-4 and IL-13 [39,40]. Once expressed on the surface of endothelial cells, P-selectin is rapidly internalized by endocytosis [41]. In addition to mediating the homing of immune cells to inflammatory sites, P-selectin-mediated interactions are also involved in controlling T-cell progenitor (circulating thymic progenitor, CTP) migration from the bone marrow to the site of T-cell development, the thymus [33].

It is important to note that differences appear to exist between the responses of human compared with mouse endothelium to inflammatory cytokines. Whereas IL-1 and tumor necrosis factor-alpha (TNF-α) are potent inducers of P-selectin expression in mouse endothelium, these cytokines only induce the expression of E-selectin (and not P-selectin) in human endothelium [42]. Additionally, experiments in transgenic mice comparing human and mouse P-selectin genes have demonstrated that following injection of TNF-α, the expression of human P-selectin mRNA decreases whereas mouse P-selectin mRNA increases [43]. This suggests that E-selectin may be the more dominant selectin in human inflammation and disease and brings into question the extrapolation to human disease of the emphasis on P-selectin ligand interactions related to mouse models.

E-selectin (CD62E) is a glycosylated glycoprotein that recognizes diverse groups of glycoconjugates on hematopoietic and cancer cells. Unlike P-selectin, E-selectin is not presynthesized in endothelial cells but, rather, is transcriptionally regulated by mediators including TNF-α, IL-1β, IFN-γ, and LPS. However, exceptions to this rule exist including the observations that vascular beds within the skin [37,44] and bone marrow [36,45] constitutively express E-selectin. Expression can occur as early as 2 h after stimulation and decline within 24 h [46]. Despite this delayed expression (as compared to P-selectin), E-selectin overlaps with P-selectin temporarily enhancing leukocyte recruitment during an inflammatory response and functionally leading to a dramatic decrease in the rolling velocity (slow rolling), and enhancing the probability of adhesion [47] and the subsequent steps of the cascade (including integrin activation) [48]. Even following the downregulation of P-selectin, E-selectin continues to enhance leukocyte–endothelial interactions leading to slower rolling velocities and improved adhesion [49]. When the stimulus wanes, E-selectin becomes internalized and degraded in lysosomes [50]. Physiologically, E-selectin-mediated interactions are involved in controlling leukocyte recruitment to injury and inflammatory sites [51], steady-state, tissue-specific homing in cutaneous tropism of skin-homing T cells [52,53], HSPC entry into bone marrow [45,52,53], as well as metastasis of CTCs.

9.1.2 E-Selectin Ligands in Therapy and Disease

Emphasis on E-selectin-mediated interactions and the regulation of migration of human cells in flow is crucial to understanding many physiological phenomena in human disease and therapy since: (i) E-selectin responds differently from P-selectin to stimuli such as inflammatory cytokines, (ii) E-selectin contributes dramatically to decreasing the velocity of the cells in flow for an extended time, and (iii) E-selectin is constitutively expressed in some vascular beds such as the bone marrow, this being important both for an understanding of how to enhance migration of blood stem cells to bone marrow and also for how to inhibit the metastasis of CTCs to this site.

9.1.2.1 E-Selectin Binds Specific Selectin Ligands on HSPCs to Guide them to the Bone Marrow

The vascular endothelial cells within the bone marrow constitutively express E-selectin and vascular cell adhesion molecule-1 (VCAM-1)[36]. Numerous studies in mice, including intravital microscopy studies, have demonstrated the importance of E-selectin, VCAM-1, and stromal cell-derived factor 1 (SDF-1) in mediating the entry of cells in flow (such as HSPCs and immune cells found in the blood) to the bone marrow [5456]. These findings have established that homing to bone marrow involves a multistep cascade initiated by E-selectin binding to its ligands on the HSPCs to mediate tethering and rolling, followed by SDF-1 binding to CXCR4 (on HSPC) leading to activation of VLA-4 (on HSPC), resulting in firm adhesion mediated by VLA-4 binding to VCAM-1 and subsequent transmigration. The Step 1 effectors of this cascade, E-selectin and its ligands, play a fundamental role in the recruitment of circulating cells to the bone marrow as demonstrated by many studies and reviewed by Sackstein [16].

Studies in both humans and mice indicate that E-selectin is constitutively expressed on marrow endothelial cells [36,54,57], and intravital studies have revealed that HSPC migration to marrow occurs at specialized microvascular beds expressing E-selectin [54]. These and several other independent lines of evidence [45,5862] have highlighted a central role for E-selectin-dependent interactions in HSPC recruitment to marrow. E-selectin-dependent binding of HSPC to marrow microvessels, either concurrent with transmigration or subsequently within “vascular niches,” could affect hematopoietic activity and with differential effects depending on the target glycoprotein ligand(s). Importantly, it is known that ligation of P-selectin glycoprotein ligand 1 (PSGL-1) can suppress hematopoiesis in mouse [63] and human [64] HSPCs. This fact, together with evidence that non-PSGL-1 and E-selectin ligands can also trigger apoptosis and growth inhibition of HSPCs from human and mouse [64], highlights a broader role for E-selectin receptor/ligand interactions in hematopoiesis. Indeed, recent work by Winkler et al. demonstrated that E-selectin is a crucial component of the vascular HSC niche that actively induces hematopoietic stem cell (HSC) proliferation and if E-selectin is removed from the equation, HSC quiescence is maintained [65]. This suggests that in the absence of a specific E-selectin-mediated proliferative signal, greater numbers of HSCs remain dormant.

Control of this process may be through the expression upon “metabolic activation” [66] of a specific E-selectin ligand with an ability to bind E-selectin and induce the observed proliferation and commitment.

9.1.2.2 E-Selectin/E-Selectin Ligand Interactions Involved in CTC Metastasis, Regulation, and Maintenance

Metastasis is the culmination of an elaborate cascade of events in which a cancer cell breaks away from the primary tumor, enters blood vessels, and travels through the body until it finally extravasates through the vessels of a distant organ to establish a secondary colony. In order to successfully establish a metastatic clone at a distant site, any CTCs must survive the targeted abuse present in the blood stream that may lead to induction of apoptosis or necrosis as well as avoid elimination by immune cells. CTCs that possess these enhanced survival capabilities (and have overcome this “microevolutionary” process) will generate metastatic colonies in distant organs and even “self-seed” [67] the primary site with more hostile tumor cells. Research focused on elucidating the molecular mediators used by CTCs to adhere to the endothelial cells lining vessels at a target site, i.e., bone marrow, may lead to therapies that can be used to dissuade metastasis.

Endothelial cells expressing E-selectin, such as the bone marrow or activated inflamed vessels, are able to capture CTCs displaying E-selectin ligands on their surface, and initiate the adhesion/metastasis cascade that leads to invasion of the tumor cell (Figure 9.1). These ligands have also been implicated in cancer stem cell regulation and maintenance relating to epithelial-to-mesenchymal transition (EMT) and other events necessary for metastatic growth. E-selectin-dependent interactions have been established in promoting metastasis of many cancers including breast, pancreatic, colon, and prostate [6872]. sLex/a epitopes are significantly upregulated on cancer cells compared with normal cells and this may play a role in promoting CTC adhesion during metastasis [68,72,73]. This increased expression of selectin ligands is due, in part, to altered regulation of the GTs responsible for creating the sLex/a epitope [74,75]. Thus, it is necessary to identify not only the glycoproteins and glycolipids presenting sialo-fucosylated glycans but also the GTs responsible in the upregulated expression of selectin ligands in order to target their downregulation. The role of E-selectin ligands and relevant GTs will be discussed in the following section.

9.1.3 Role of GTs in Controlling Cell Migration

As illustrated in Figure 9.1, the migration of cells to a specified site is dependent on the interaction of a number of ligands and receptor pairs expressed both on the surface of the migrating cell and on the surface of the endothelium, these working together in a coordinated multistep fashion to orchestrate the delivery of cells to a specific site. Although many molecules are involved in organizing the steps involved in this cascade of events, the selectin binding to its ligand(s) literally “gets the ball (cell) rolling” and on its way to enter a tissue/organ. This initial interaction is the prerequisite for the remaining steps in the cascade and is dependent on the careful glycosylation of selectin ligands. The glycan structure most common for the recognition of ligands by selectins is the sLex (or its stereoisomer sLea) addition to the N-/O-glycan on the protein (Figure 9.2). GTs expressed in the ER and Golgi act in an assembly line fashion by adding one glycan at a time to a growing carbohydrate chain whereby the product of one GT reaction becomes the substrate for the following GT. All GTs are type II transmembrane-Golgi/ER localized enzymes with a short NH2 terminal tail in the cytosol followed by a transmembrane domain, which is linked to the catalytic domain in the Golgi/ER lumen by a short stem region. Generally, each GT uses only one of several sugar nucleotide substrates (including UDP-Galactose (Gal), UDP-Glucose (Glc), UDP-N-acetylgalactosamine (GalNAc), UDP-N-acetylglucosamine (GlcNAc), GDP-fucose (Fuc), GDP-mannose (Man), UDP-xylose (Xyl), or CMP-sialic acid (SA)). Each GT also uses one and sometimes more glycan acceptor substrates and, as they are stereospecific, these enzymes are only capable of forming one specific glycosidic bond (i.e., either an α or β anomer) [76]. Glycan synthesis is the consequence of a series of ordered, GT-dependent events that are characterized by glycan polymer elongation or branching. In general, it is the suite of GTs expressed by a cell that will dictate the glycan structures made. However, the expression of specific GTs does not permit prediction of the glycan structures that will be synthesized. Many factors can affect the activity and ability of the GTs to create a given structure such as the availability of sugar nucleotides, the relative location of the GT in the Golgi stacks in relation to the substrates that it can recognize, and even the possibility of modifications of the glycan product by naturally expressed glycosidases [5,77]. In the context of the glycan structures that confer selectin counter receptor activity on a cell, the GT suite will therefore participate in the control of selectin-dependent leukocyte recruitment. Although all three selectins recognize the sLex epitope, each selectin requires particular fine structural details of this epitope or of the peptide backbone to confer optimal binding affinities.

In order for selectin ligands to be formed, GTs responsible for creating the sLea/x structures must be endogenously expressed and active. The native display of these glycan structures on selectin ligands requires the expression and activity of various GTs and sulfotransferases, including the action of α1,3- or α1,4-fucosyltransferases, α2,3-sialyltransferases, β1,4-galactosyltransferases, and β1,6-N-acetyl-glucosaminyltransferases on either a protein scaffold (i.e., N/O-glycans on glycoproteins) or a lipid scaffold (i.e., glycolipids). Specifically, this specialized carbohydrate motif comprises sialofucosylations containing an α(2,3)-linked sialic acid substitution on galactose, and an α(1,4)-linked (for sLea) or α(1,3)-linked (for sLex) fucose modification on N-acetylglucosamine, of either a type 1 (for sLea) or type 2 (for sLex) lactosamine unit, creating a terminal tetrasaccharide sLea or sLex structure, respectively (Figure 9.2). Although E-selectin binds both sLea and sLex, it binds sLex with a higher avidity [78] and, with the exception of human metastatic cancer cells, most human hematopoietic cells and “normal” stem cells prototypically express only type 2 lactosamine units [79,80]. Understanding the expression and regulation of GTs in the creation of E-selectin ligands will guide efforts to manipulate the migration of human cells in circulation. Indeed, knowledge of which GTs are required to create E-selectin ligands has allowed researchers to manipulate cells in order to create ligands on cells ex vivo with the intention of helping direct the migration of therapeutic cells (such as HSPCs) to specific tissues (such as the bone marrow) that express the corresponding selectins [59,79]. Such work will be discussed in the following section.

9.1.3.1 GTs Involved in Creation of E-Selectin Ligands

For the most part, the enzymes responsible for forming E-selectin ligands are similar to those responsible for both L-selectin and P-selectin ligand formation in that they require sLex structures. However, unlike L- and P-selectin ligands, the sLex cap can occur on a number of different glycans from N-glycans to glycolipids. In addition, analysis of FXnull mice (mice with a null mutation of the FX locus that encodes an enzyme within the de novo pathway for GDP-fucose synthesis) by Lowe’s group suggests that E-selectin ligands are much more sensitive than P-selectin ligands to loss of fucose, this further highlighting the importance of fucosylation to the formation of functional E-selectin ligands [81].

9.1.3.2 FTs and E-Selectin Ligand Formation

α1,3-fucosylation is imperative for E-selectin ligand activity and a number of human fucosyltransferases (FTs) have been identified that are able to modify terminal lactosamines. These include FT-III, FT-IV, FT-V, FT-VI, FT-VII, and FT-IX [82,83]. FT-III to FT-VI are all able to modify either sialylated or unsialylated type 2 lactosamines resulting in sLex or Lex (Figure 9.2) structures, respectively, while FT-III (and FT-V) is also able to modify type 1 lactosamines resulting in sLea or Lea.

Analyses of mice with targeted mutations in FT genes reveal that only two of these enzymes are required to form functional selectin counter receptor activity in leukocytes. These studies all support the observation that FT-VII and FT-IV cooperate in the synthesis of fucosylated glycans required for all leukocyte selectin ligand activity [8486]. Specifically FT-VIInull mice have leukocytosis and their blood leukocytes have reduced E- and P-selectin ligand activity that effectively compromises the recruitment of leukocytes to inflammatory sites in in vivo mouse models of inflammation, such as the thioglycollate-induced peritonitis model [85] and the contact hypersensitivity model (CHS) [84,86]. However, residual recruitment is still evident (particularly with myeloid cells) and is due to FT-IV activity. CHS response studies performed in mice deficient in FT-IV, FT-VII, or both enzymes revealed that, although lymphocyte (Th1 and Tc1) recruitment was absent in FT-VIInull animals, both enzymes needed to be knocked out to inhibit the recruitment of myeloid cells. A subtle decrease in the myeloid cell CHS response was seen in FT-IVnull mice whereas lymphocyte recruitment was unaffected. Overall, these observations support the hypothesis that FT-VII dependent fucosylation of selectin ligands accounts for most of the selectin-dependent adhesion in leukocytes especially within the lymphocyte fraction. The dependence of selectin ligand formation on FT-VII is also paralleled in studies with human T lymphocytes [8789]. Indeed, studies suggest that FT-IV acts preferentially on glycolipids whereas FT-VII acts on glycoproteins [90] and, interestingly, gangliosides are able to mediate tethering and rolling [9193] on endothelium. However, glycoproteins are better at facilitating firm adhesion under physiological flow conditions [93] which could also be related to signaling through the glycoprotein selectin ligands to activate integrins directly [17]. Recent studies focused on assessing knockdowns of FT-IV and -VII in human cells suggest that an additional α1,3-fucosyltransferase, FT-IX, is important in decorating E-selectin ligands on human leukocytes but not mouse [94,95]. This is significant since the selectin ligands, particularly those that bind E-selectin, vary between different leukocyte cell populations and among species [96].

9.1.3.3 ST3GalTs and E-Selectin Ligands

α2,3-sialic acid linkages to terminal galactose residues are crucial in the formation of sLex epitopes. The contribution of sialylation to the formation of selectin ligands is well documented as sialidase treatment of leukocytes under static conditions in vitro completely abolishes binding of P-selectin and E-selectin to selectin ligands [97]. This linkage is catalyzed by the activity of the sialyltransferases (STs) ST3GalT-I, -II, -III, -IV, -V, and -VI. Only ST3GalT-III, -IV, and -VI have been shown to sialylate type 2 lactosamines and ST3GalT-IV and ST3GalT-VI are implicated in E-selectin (and P-selectin) ligand formation [97100]. In vivo inflammatory studies of ST3GalT-IVnull mice reveal a mild but significant defect in selectin ligand function, including a mild reduction in E-selectin-dependent rolling, an increase in E-selectin-dependent rolling velocity, a decrease in L-selectin-dependent rolling during inflammation (but no effect on L-selectin-dependent rolling on HEVs [101]), and no contribution to P-selectin ligands [97,102]. Since sialylation is so crucial to selectin binding but yet the phenotype of the ST3GalT-IVnull is so mild, it is suggested that other STs, particularly ST3GalT-VI, contribute to selectin ligand formation. With the generation of the ST3GalT-VInull and ST3GalT-IV/ST3GalT-VI double knockout mouse, the role of these STs appears to be critical for the generation of functional selectin ligands, especially in the case of the formation of P-selectin ligands (ST3GalT-VI is involved in E-selectin-mediated capturing but not in E-selectin-dependent rolling) [100]. Implications for the involvement of additional STs involved in selectin ligand synthesis and function are still warranted, especially for E-selectin and L-selectin.

9.1.3.4 C2GnTs and Functional E-Selectin Ligands

C2GnT (core 2 β(1,6)-N-acetylglucosaminyltransferase) isoenzymes (C2GnT-I, -II, -III) create the core 2 O-glycan branch [103] by adding GlcNAc to the Galβ1,3GalNAc core 1 structure expressed on serine or threonine residues. Capping of the core 2 branch with a sLex moiety creates the P-selectin counter receptor. Analyses of C2GnT-Inull mice, however, show that it is the isoenzyme essential for PSGL-1 modifications on leukocytes [104106]. Despite the essential role of C2GnT-I in P-selectin ligand formation, C2GnT-Inull mice possess a relatively mild phenotype, showing only a partial reduction in selectin ligands and essentially no effect on lymphocyte homing [104], reflecting the possible redundancy of the C2GnT isoenzymes. It has been suggested that C2GnT-II and/or C2GnT-III contribute to selectin ligand formation and cell trafficking [107,108]. C2GnT-I and C2GnT-III are expressed primarily by lymphocytes while C2GnT-II is associated with goblet cell mucin production in the intestinal epithelium. A single knockout of any of these enzymes and double knockout of C2GnT-I and C2GnT-III had no effect on the trafficking of lymphocytes in an inflammatory mouse model of the large intestine. However, mice deficient in C2GnT-I and C2GnT-II or all 3 C2GnT enzymes were not able to overcome a helminth infection, suggesting that C2GnT-dependent modifications on lymphocytes may play a role in migration of lymphocytes (to the large intestine) [108].

In the context of E-selectin ligands, a number of studies indicate that cells deficient in C2GnT-I are able to bind to E-selectin, although at a slightly reduced level compared with wild-type cells [104,107,109]. These studies reveal that 75–80% of E-selectin binding is maintained in the absence of C2GnT-I, suggesting that either the isoenzymes are compensating for the loss of C2GnT-I or that other E-selectin ligands (besides PSGL-1) are dominant mediators of this interaction.

9.2 Use of Enzymes to Modify Cell Surface Carbohydrates on Proteins and Lipids to Enhance Migration to Tissues

Direct injection of therapeutic cells into target tissue sites is only practical for a limited number of applications including pancreatic beta islet transplants through the portal vein into the liver or intracoronary infusions of myoblasts into damaged heart muscles. For the vast majority of therapeutic cellular-based therapies that are targeting systemic multifocal sites of injured or inflamed tissues or organs, intravenous transfer is the routine method of administrating cells. Therefore, it is crucial that the efficient homing of donor cells to the site of intended action be ensured. Since only a subset of intravenously injected cells may engraft in the intended location due to the absence of the key homing molecules on infused cells, the development of methods for improving therapeutic cell trafficking is a high priority. One relatively simple cell engineering method uses the GT enzymatic machinery responsible for creating elaborate glycan structures on cell surface proteins to program the in vivo trafficking of systemically delivered cells. Through the introduction of GTs important for creating the key carbohydrate structure, sialyl Lewis x/a (sLex and sLea), on the cell surface, infused cells can then be attracted by the cognate selectin counter receptor expressed on the endothelium and direct its subsequent migration. These GTs can be introduced genetically using expression vectors containing the GT gene of interest, induced through cytokine signaling mechanisms, or using recombinant active GTs to add on glycans ex vivo.

9.2.1 Genetic Approaches for Modification of Cell Surface Glycoconjugates

It is possible to introduce into cells genes that are required to achieve the formation of the glycoconjugates on the surface of cells that render selectin ligand binding.

Among the 13 family members recognized in the human genome, eight FTs (FT-III, FT-IV, FT-V, FT-VI, FT-VII, FT-IX, FT-X, and FT-XI) are reported to possess the alpha-1,3-fucosyltransferase activity which potentially could confer formation of the sLex structure with distinctive substrate specificities [110,111]. In particular, it has been shown that FT-III, FT-IV, FT-V, FT-VI, and FT-VII are involved in the formation of the selectin ligand and by inhibiting the expression of these FTs in various cell types including leukocytes [95], prostate cancer cells [71], and gastric cancer cells [112], selectin binding is inhibited. Consequently, FTs can be considered as good candidate genes to introduce to therapeutic cells and enhance sLex formation. The mode of introduction of FTs into cells could be either transient or stable. Since the formation of the sLex moiety is a prerequisite only in the initial step of the transition of cells to a tissue, transient expression of FT could adequately satisfy this requirement. The benefit of the transient expression strategy compared with stable expression through viral vector-mediated gene integration is that the former approach avoids genetic disturbance of the genome and so minimizes the likelihood of spontaneous oncogenesis.

Successful physical approaches for the delivery of plasmid vectors into cells for transient expression have been reported using lipofection [113,114] or electroporation [115,116], in both cases with good transfection efficiency, high viability and recovery of transfected cells, and without any significant effect on capability for differentiation or proliferation.

One potentially intriguing approach to overcome the relatively lower efficiency of transient transfection is the use of a nonintegrating version of foamy virus-based vector [117], which has been shown to transduce stem cells rather effectively [118]. Although this vector system is still under development, it could permit a high efficiency by virtue of being a virus-based transduction while at the same time keeping the genome of the target cells intact.

Stable expression of exogenous genes can be achieved using genomic integration mediated by virus-based vectors. Various viral vector systems have been used for transduction of cells including lentivirus-, retrovirus-, adenovirus-, adeno-associated virus-, or baculovirus-based vectors to achieve long-term and stable expression of exogenous genes [119,120].

Although ex vivo enzymatic treatments are generally viewed as being more desirable in terms of simplicity and efficiency, it would seem reasonable to employ a viral vector approach when the introduction of transgenes or correction of a genomic mutation is required.

9.2.2 Cytokine and Growth Factor Induced Expression of GTs Within Cells

Several cytokines are implicated in the regulation of the key GTs involved in selectin ligand synthesis of leukocytes and other circulating cells, such as hematopoietic stem/progenitor cells, as they undergo differentiation and specialization.

In T cells, expression and regulation of GTs important in forming P- and E-selectin ligands require antigen-stimulated activation [121124]. Following this finding, a number of studies focused on defining how key cytokines involved in the activation of T cells affect the expression of GTs that create sLex epitopes on E- and P-selectin ligands. In vitro activation studies using cytokines that induce the polarization of CD4+ and CD8+ T cells toward activated TH1 (TC1) or TH2 (TC2) subsets demonstrated that the GTs C2GnT-I, FT-VII and ST3GalT-IV and -VI are differentially regulated depending on cytokines used. TH1 cytokines, including IL-12, -2, -15, and TGF-β tend to enhance the expression of C2GnTs [107,125,126], FT-VII [127131], and ST3GalTs [127,132] and induce binding to E-selectin and P-selectin, while TH2 cytokines, such as IL-4 and -5, repress their expression. Based on their potent impact in vitro, these cytokines were expected to modulate homing receptor formation under in vivo conditions. Although these cytokines appeared to regulate C2GnT-I in vitro, there was no effect on the selectin ligand formation and proliferative responses in the absence of these cytokines in vivo [133].

Other signals have been implicated in the regulation of selectin ligand formation, such as the vitamin A and D derivatives. A recent study illustrated that exposure of human T cells to physiological levels of retinoic acid (RA) and 1,25D(3) completely eliminated E-selectin ligand activity and subsequent skin-homing abilities, whereas the inactive/precursor forms of vitamins A and D had little inhibitory action [134]. Analysis of GTs involved in forming sLex structures revealed that FT-VII and ST3GalT-III were significantly reduced upon exposure to these vitamin derivatives. This finding is intriguing as RA is currently being used successfully as a differentiation therapy for the treatment of acute myeloid leukemia (AML) subtype 3. In addition to inducing the differentiation of the diseased leukemic cells, it may also influence the ability of the leukemic blasts/stem cells to migrate to the bone marrow where E-selectin is expressed by inhibiting the GTs involved in creating E-selectin ligands.

In hematopoietic stem/progenitor cells and myeloid cells, G-CSF (granulocyte-colony stimulating factor) significantly enhanced the expression of sLex on HSPCs and their subsequent ability to bind E-selectin and home in vivo. This was accompanied by an associated increase in the expression of the GTs, ST3GalT-IV, FT-IV, and FT-VII [135]. Studies on human CD34+ cord blood stem cells exposed to dimethyl-prostaglandin E2 (dm-PGE2) have demonstrated that this treatment increases HSPC numbers and enhances their homing [136138]. However, to date, this enhanced homing has not been correlated with measurements of GT or sLex expression or selectin binding.

Immune blood cells, such as lymphocytes and macrophages, are known to secrete inflammatory factors that could potentially influence cancer cell metastatic progression. In particular, the proinflammatory cytokines, TNF-α, and IL-6 have been reported to be elevated in the blood serum of patients diagnosed with advanced stage breast cancer and this has been correlated with an increase in the number and size of metastatic sites [139]. These cytokines have also been shown to promote the growth and invasiveness of breast, colon, and prostate cancer cells [140142]. In addition, cancer cell chemokine receptors have also been implicated in playing a role in identifying the destination of metastases [143]. It can be inferred that these cytokines lead to enhanced expression of GTs involved in the creation of sLex epitopes on ligands of these cancer cells, which results in enhanced adhesion to selectins [140]. This process will be discussed later.

9.2.3 Use of Recombinant GTs ex vivo

GTs expressed in the ER and Golgi act in an assembly line fashion by adding one glycan at a time to a growing carbohydrate chain decorating a particular glycoprotein. Depending on which GTs are expressed in a particular cell, a diverse set of structures can be made. In order for selectin ligands to be formed, GTs responsible for creating the sLex structure need to be endogenously expressed and active, as outlined in Figure 9.2. The knowledge of which GTs are required to create particular structures allows researchers to manipulate cells in order to create ligands on cells’ surface ex vivo with the intention of helping to direct the migration of therapeutic cells to specific tissues that express the corresponding selectins [59,79,144].

The microvasculature of human and murine bone marrow constitutively expresses E-selectin and, as described earlier, extensive research has determined that HSPC migration to the bone marrow depends on the expression of ligands that bind E-selectin. Both mouse and human HSPCs express a number of E-selectin ligands. Studies to identify the glycoprotein ligands revealed that on human CD34+ HSPCs, at least three different ligands contributed to E-selectin binding. These include the hematopoietic cell E-/L-selectin ligand (HCELL; sLex-decorated glycoform of CD44 that binds E-selectin and L-selectin), cutaneous lymphocyte antigen (CLA; sLex-decorated glycoform of PSGL-1 that binds E-selectin), and a sLex-decorated glycoform of CD43 [96]. In contrast, on mouse HSPCs only CLA and CD43 were identified as ligands suggesting that the modest E-selectin binding detected in mouse HSPCs compared with human HSPCs is due to the absence of HCELL. Indeed, inducing the expression of HCELL on mouse HSPCs leads to significant enhancement of E-selectin binding in mouse HSPCs [96]. HCELL expression was achieved using an innovative tool to custom engineer desirable cell surface glycoforms (i.e., sLex) on therapeutically beneficial cells (i.e., stem cells) using recombinant GTs to transform the cell surface glycans temporarily in order to enhance their migration to specific sites (i.e., E-selectin on bone marrow microvasculature).

Selectin ligands must be α1,3-fucosylated to form the terminal glycan sLex determinant. Often stem cells are inadequately α1,3-fucosylated, this being the case with MSCs [59], HSPCs isolated from cord blood [61,144], as well as with neural stem cells (unpublished data), and so exhibit homing defects. Indeed ex vivo treatment of stem cells (principally human MSCs and HSPCs derived from human cord blood) with FTs, particularly FT-VI and FT-VII, increases cell surface sLex determinants, boosts binding to E-selectin and P-selectin, and enhances homing and engraftment into bone marrow of immunocompromised mice [59,144146]. By generating a comprehensive set of tools (i.e., GTs) for ex vivo design of sLex structures on the surface of a therapeutic cell, strategies to enhance the migration of these cells for the treatment of diseases can be further anticipated.

For many years, scientists have been interested in understanding and characterizing the complex nature of carbohydrates displayed on cell membranes and the GTs responsible for creating these molecules. Early studies focused on isolating small amounts of GTs from natural sources and using these enzymes in reactions that included radioactive donor sugar nucleotides to rebuild cell surface structures following their removal using glycosidases; the proteins that resulted were then identified by SDS-PAGE [147,148]. These studies also helped to characterize carbohydrate-dependent binding interactions such as those between pathogens and host cells. The use of GTs to manipulate glycans became much more practical with advancements in recombinant DNA technologies that allowed for the expression of proteins and enzymes on a large scale. The potential for creating libraries of recombinant GTs for fine tuning glycans on cells or proteins brings extensive possibilities, such as enhancing the stability of proteins, creating enhanced bioactive/humanized mAbs [149151], and directing the movement of cells within the body.

The use of recombinant GTs to modify cells ex vivo has been elegantly formulated and discussed by Sackstein [59,79,152154]. He outlines the guiding principles (summarized in Table 9.1) and methodological details (Figure 9.3) required for the custom engineering of cell surface glycans of a cell population of interest, particularly stem cells, using his “GPS” (glycosyltransferase-programmed stereosubstitution) technology [152,153]. Any GT enzyme can be used to create the glycan of interest using this technique but special attention must be given to ensure that the enzyme assay conditions (storage buffer, reaction buffer) are optimized to maintain cell viability and avoid phenotypic effects (such as the differentiation) on cell(s). One major concern is that many GTs require divalent cations for their catalytic activity [156]. Accordingly, the recombinant human FT-VI enzyme and reaction conditions specifically developed by Sackstein effectively catalyzed α(1,3)-fucosylation at physiologic pH in the absence of divalent cations [59].

Table 9.1

Guiding Principles Developed by Sackstein [152,153] to Apply GPS Technology to Cells

Strategy for Custom Engineering Cell Surface Glycans

I. Identify target glycoconjugate “acceptor”

II. Use appropriate enzyme(s) that result in stereospecific carbohydrate epitopes

III. Confirm the creation of the target modification biochemically and functionally in vitro and in vivo

image
Figure 9.3 Ex vivo GT treatment of cells. Preparations of recombinant GTs, such as FT-VI [59,144] and -VII [146], have been used to treat stem cells. In this example, the FT-VI enzyme is illustrated, along with the reaction conditions needed to catalyze α(1,3)-exofucosylation efficiently (near-physiologic pH; absence of divalent cations, i.e. Mn2+), by adding fucose from a nucleotide sugar donor, GDP-Fuc (guanosine 5′-diphospho-β-L-fucose), to a sialylated lactosamine structure on glycoproteins or glycolipids in order to create the sLex structure critical to guide the migration of cells to E-selectin-expressing endothelial beds such as the bone marrow [59,152]. Once the structure is formed, a number of functional assays are performed to determine the efficacy of the enzymatic treatment including parallel plate flow chamber studies and blot rolling assays [152,155].

Many mammalian proteins have been expressed and prepared as active recombinant proteins [157159]. A number of different expression systems could be used (bacterial cells, mammalian cell lines, baculovirus-infected insect cells, yeast cells, and even silk worm) to express these GTs but, since several reports have shown that N-glycosylation is necessary, higher order systems are preferred for optimal enzymatic activity; in fact, bacteria-expressed FTs exhibit no detectable activity [160,161]. As bacteria-based systems would not work and production of large amounts of the required enzymes is necessary, Sackstein et al. chose to use the methanotrophic yeast Pichia pastoris expression system. This system offers the advantages of generating a high yield of recombinant proteins and an ability for the proteins to be secreted into the medium which facilitates downstream preparation of the enzyme [162]. Recent efforts have been made to investigate the expression and isolation of GTs from silkworm. This is a clear option for the preparation of recombinant enzymes, which cannot be expressed functionally in bacterial expression systems. The silkworm approach uses pupal or larval individuals of a lepidopteran insect, Bombix mori, and a baculovirus-based vector for exogenous gene expression [163]. This system brings several benefits including optimal folding and posttranslational modifications resulting in proper functionality of the many recombinant proteins, an extremely high yield, great amenability to scaling-up, and a short time period for production of the protein. Although few laboratories currently have the necessary technology, several companies offer a commercial protein expression service, which uses this system.

9.2.3.1 Glycans and Their Importance in Mediating Engraftment

Once HSPCs have used their newly acquired homing molecules to enter the bone marrow, successful transplantation then requires engraftment of the stem cells within their particular niche. Implications for the activity of GTs depend not only on improving the homing of therapeutic cells to their niche but also on the successful engraftment of these cells. Normally HSCs are located within specific regions in the bone marrow that dictate the behavior of these cells. The complex interplay of cells and molecules helps to regulate HSPC quiescence, division, and differentiation. Various cell types have been identified that contribute to the HSPC niche, including various mesenchymal stem cells (CD146+ perivascular MSCs [164], nestin+ MSCs [165], and leptin receptor+ perivascular MSCs [166]), CXCL12 abundant reticular (CAR) cells, and MSC-derived osteoblast lineage cells [167169]. In addition, endothelial cells form an overlapping HSC niche [166,170]. These niches also appear to be further regulated by local sympathetic nerves [171]. A model has been proposed whereby the “osteoblastic” niche (endosteal surface of bone) harbors and maintains HSCs in a quiescent state, whereas the “vascular” niche (marrow vessels) promotes HSC proliferation and differentiation [65,172,173]. However, many recent studies have challenged this model [174] stating that the osteoblastic niche is actually extensively vascularized [55,175], and that since the most potent and quiescent HSCs prefer to reside in hypoxic [176,177], poorly vascularized regions, these may not reside in the “osteoblastic” niche of the bone marrow. A particularly interesting study has demonstrated that E-selectin, expressed on around 20% of bone marrow endothelial cells (BMECs), directly promotes HSC proliferation and, in the absence of this selectin, HSC quiescence and self-renewal potential increases [65]. These findings suggest that, in addition to directing the migration of HSPCs to the bone marrow, E-selectin—through a currently undefined ligand—also plays a key role in regulating HSC engraftment and fate. Currently, quite which E-selectin ligand(s) on the HSC mediate this proliferation remains unknown as studies focused on knocking down the key E-selectin ligands identified to date do not present the same HSC phenotype [65,96]. Further studies are warranted to identify these ligands.

9.3 Modification of Enzyme Activity Leads to Changes on Cell Surface Structures that Deter Migration and Metastasis

Metastatic colonization, outgrowth at a distant site, starts with the interaction between CTCs and endothelial cells lining lymph and postcapillary vessels. This interaction involves many adhesion molecules and molecular pathways, such as selectins, integrins, intercellular adhesion molecules (ICAMs), and chemokines with their receptors. These molecules work collectively in a process that obeys the multistep paradigm of HSPC recruitment to bone or leukocytes to sites of inflammation (Figure 9.1). Although still debated, “hematopoietic mimicry” or the “homing concept of metastasis” hypothesis states that at Step 1 the circulating cell, moving under high shear force of circulation, uses selectin-mediated low affinity reversible rolling interactions with endothelial cells expressing E-selectin and P-selectin [68,178,179]. This interaction facilitates extravasation of CTCs and determines organ-specific metastasis depending on the tumor type and metastatic environment. Selectins are the receptors for many glycoconjugate selectin ligands.

9.3.1 Selectin–Selectin Ligand Axis in Metastasis

The role of the selectin–selectin ligand axis in promoting cancer metastasis is supported by a number of findings described below and is summarized in Figure 9.4.

image
Figure 9.4 Selectins mediate tumor progression and metastasis. At the primary site, a prometastatic transition program confers higher migratory abilities of the tumor cell including increased expression of GTs involved in the creation of sLex/a (A). Migratory tumor cells invade the blood circulation searching for a proper metastatic niche at a distant site to form the secondary tumor mass; in selectin-dependent metastasis, tissues that express selectin(s) capture selectin ligand-expressing CTCs allowing tissue tropism of metastasis (B). In the circulation, selectin-expressing leukocytes and platelets (D) coat the tumor cells to form metastatic emboli (C) that help the tumor cell evade the recognition and attack by immune cells. Moreover, metastatic emboli bridge an interaction between tumor cells and selectin-expressing endothelial cells through selectin ligand-expressing leukocytes (E). Premetastatic niche induces selectin expression on the surrounding endothelial linings by releasing inflammatory cytokines (F). Metastatic colonization commences when the CTCs encounter selectin-mediated capture and rolling allowing downstream adhesion and extravasation events (G). Cells devoid of selectin ligands lack this ability to traffic, which decreases their ability to survive in the circulation leading to a lower metastatic potential (H).

Many types of metastatic epithelial tumors, including breast [74,180,181], prostate [182,183], colon [184187], lung [188,189], and pancreatic carcinomas [190], show a substantial ability to roll on endothelial cell layers natively expressing E-selectin. Although this interaction alone is not sufficient to explain the organ specificity of metastasis, it is the initial interaction and the primary adhesion that directs organ-specific metastasis. Lewis lung cancer cells tend to metastasize preferentially to liver where endothelium shows upregulation of E-selectin expression [191,192]. Additionally, prostate and breast carcinomas tend to metastasize to bone, where E-selectin is constitutively expressed and, interestingly, highly metastatic cells show a stronger interaction with bone marrow endothelium compared with weakly metastatic carcinomas [181183,191,192].

Experiments in mouse models deficient in or overexpressing one of the selectins indicate a direct correlation with selectin expression and metastatic potential. For example, Biancone et al. [193] demonstrated a reversion of metastasis from lung to liver when E-selectin was upregulated in vivo. Another study showed that metastasis is inhibited upon immunoglobulin E-selectin blockade [194]. Moreover, studies in P-selectin and L-selectin deficient mice displayed impaired metastatic potential [195,196]. A growing body of evidence supports the proposal that inflammation potentiates cancer dissemination and progression. Immune cells lying in a tumor microenvironment both at the primary site and at premetastatic sites release proinflammatory cytokines that not only induce the upregulation of selectin expression on endothelial monolayers but also upregulate the enzymes responsible for carbohydrate ligands (GTs) on cancer cells. For instance, E-selectin expression in colorectal tissues is directly proportional to inflammatory reactions at the lesion–environment interface [191,192,197], and this may explain the sensitivity of colon cancer to anti-inflammatory therapy [198].

Selectins support the formation of metastatic emboli. L-selectin and P-selectin expressed on leukocytes and platelets, respectively, bind selectin ligand glycoconjugates and soluble mucins on the surface of CTCs forming metastatic emboli [199]. Metastatic emboli support metastasis in two ways: first, they allow better interaction with E-selectin-expressing endothelium through ligand-expressing leukocytes, and secondly, the coat of platelets is thought to offer a shield against the immune system [195,200].

Carcinogenesis is associated with abnormal GT expression [73] and, hence, typically a shift toward increased sLex/a expression [201203] (as outlined below). In fact, levels of sLex within tumors correlate with malignancy and patient survival [75,204208].

Selectin ligand binding, perturbation in GT expression, and subsequent sLex/a expression are correlated with prometastatic modifications such as EMT [209]. EMT is a rapid transformation process involved in embryogenesis, wound healing, and invasiveness of cancer cells, within which a polarized epithelial cell undergoes cellular and biochemical changes to transform into a mesenchymal phenotype. Mesenchymal phenotype is characterized by loss of polarization, increased invasiveness, high motility, elevated resistance to apoptosis, and expression of extracellular matrix components [210214]. Cancer cells undergoing this transformation, often at the tumor host interface, tend to invade and metastasize to surrounding and distant tissues leading to life-threatening manifestations of cancer progression [215223]. A study by Sakuma et al. [209] showed that EMT induction in colon cancer cell lines using EGF and bFGF caused elevated expression of sLex and sLea as well as E-selectin binding. These changes were linked to upregulation of ST and FT expression. Pinho et al. [224] reported that there is a negative correlation between sLex and E-cadherin expression in two mammary tumor models. The loss of the latter is a well-known marker for EMT.

9.3.1.1 Selectin Ligands and Metastasis

To date, many selectin ligands are known to be expressed on carcinoma cells, including PSGL-1 [182,225], CD44v [181,185,226228], podocalyxin-like protein (PCLP) [229], death receptor-3 (DR-3) [186], gangliosides [93], mucin 1 (MUC1) [180,230,231], CD43 [232,233], and Mac-2BP [234] (Figure 9.1). PSGL-1 is the most characterized selectin ligand at the cellular, molecular, and functional levels in addition to being the only ligand currently known to be able to bind all three selectins. PSGL-1 is a type 1 transmembrane sLex/a-bearing mucin-like glycoprotein of 402 or 412 amino acids [235,236]. Moore et al. [237] first identified PSGL-1 as a P-selectin ligand on myeloid cells including neutrophils and HL-60 cells. PSGL-1 is a disulfide homodimer with apparent molecular weights of 240 kD (dimer) and 120 kD (monomer) on SDS-PAGE. The extracellular domain is rich in serine, threonine, and proline, and consists of 15 or 16 decameric repeats extended linearly due to the presence of multiple serine/threonine O-glycosylated sites that straighten the backbone of mucin proteins [238]. Three N-terminal tyrosine amino acids at positions 5, 7, and 10 are arranged in a consensus sequence that favors tyrosine sulfation. The cytoplasmic domain of PSGL-1 has 69 amino acids with signaling functionality but with no consensus sequence for phosphorylation [235]. PSGL-1 is expressed on hematopoietic cells and cancer cells. CD34+ HSPCs and myeloid cells at different states of maturation (except erythrocytes and megakaryocytes) and all peripheral blood leukocytes express PSGL-1 [239,240]. Expression of PSGL-1 on platelets has been reported by Frenette et al., although remains debated [241]. PSGL-1 is also expressed by leukemic cells [242] and carcinoma cells such as colon [225] and prostate cancer cells [182]. PSGL-1 binds to E-selectin in a different way than to L-selectin and P-selectin. Both sialylation and α1,3-fucosylation, that appear on core 2 O-glycan chains, are required for E-selectin binding but sulfation (of either sLex moiety or tyrosine) is not required [243]. The binding to E-selectin of PSGL-1 expressed by CTCs implies a functional role in bone tropism of metastatic cancer cells. Dimitroff et al. demonstrated that PSGL-1, expressed on the surface of the bone-metastatic cell line MDA-PCa-2b, supports rolling on BMECs. This rolling was abolished by using neutralizing antibodies against E-selectin, suggesting that E-selectin/E-selectin ligand interactions mediate this adhesion. Neuraminidase treatment of the MDA-PCa-2b cells also impaired rolling indicating the importance of sialic acid in mediating this interaction and subsequent rolling. Performing a Western blot analysis on the membrane fraction of MDA-PCa-2b for the expression of sLex (detected by HECA-452 mAb) revealed reactivity with CD44 in addition to PSGL-1 [182,183].

Many reports have characterized the binding of CD44 to E-selectin on HSCs, leukocytes, and CTCs. CD44 is known to direct the recruitment of lymphocytes toward inflammatory sites through its interaction with matrix HA (hyaluronin) and endothelial cell expressed E-selectin, and a similar interaction is thought to direct the metastasis of many cancer types [80]. The E-selectin binding form of CD44 is known as HCELL. HCELL is expressed on the surface of HSPCs, human leukemia cells and has recently been located on colon cancer and breast cancer cells. Unlike mucin-like glycoproteins, CD44 bears sLex/a structures on N-linked glycan branches [181,185,226228,244].

MUC1 is a large heavily glycosylated transmembrane protein with an extracellular region comprising a variable number of tandem repeats that include 20 conserved amino acids (HGVTSAPDTRPAPGSTAPPA) each with five potential sites of O-glycosylation (as underlined). In normal cells, MUC1 is expressed on apical surfaces and glycosylated with core 2 branched O-glycans and lactosamine extensions, whereas in cancer cells MUC1 is expressed on the whole surface and decorated with aberrantly truncated O-glycosylated carbohydrate antigens. MUC1 expressed on breast cancer cells tends to present sialylated and unsialylated core 1 structures instead of core 2 structures. It has long been known that ST expression and activity is higher in breast cancer cells than in normal breast tissue. Thus, the exposure of truncated glycan structures may be due to sialylation of core 1 GalNAc or lactosamine structures leading to a blockade of further extension and expression of T and Tn antigens [245]. The MUC1 interaction with E-selectin was first described in colon cancer cells by Zhang et al. [231]; MUC1 from a colon cancer cell line or colon cancer patient serum was shown to be able to inhibit the binding of leukemic cells to E-selectin-expressing cells. MUC1, expressed by colon cancer cells, was reported to bind E-selectin better than CD43 from the same cells under flow conditions [230]. A recent study by Geng et al. [180] used E-selectin and ICAM-coated microtubes as a model for microvascular endothelium to show that the MUC1-expressing metastatic breast cancer cell line, ZR-75-1, was able to roll and adhere to the walls of the microtube under flow conditions, so implicating MUC1 as a ligand for E-selectin on breast cancer metastatic cells. CD24, another mucin-like glycoprotein, was also identified as an E-selectin and P-selectin ligand in breast cancer cell lines (KS and MCF-7) [246,247].

More recently, a number of novel ligands have been identified as E-selectin ligands on cancer cells including PCLP on colon carcinoma cells [229], Mac-2BP on breast cancer cells [234], and DR-3 on colon cancer [186]. In addition to these glycoproteins, gangliosides have also been shown to play a role as E-selectin ligands in breast cancer cell lines [93]. Mac-2BP is a highly glycosylated protein whose expression is correlated with cancer metastasis. Shirure et al. were able to immunoprecipitate E-selectin ligands from metastatic breast cancer (ZR-75-1) lysate using recombinant E-selectin and identified Mac-2BP as a novel ligand through MS analysis. Gene silencing of Mac-2BP significantly reduced the binding of ZR-75-1 cells to E-selectin-expressing human umbilical vein endothelial cell (HUVEC) cells in flow assays [234]. All these studies highlighted the role of selectin/selectin ligand interaction in supporting metastasis. Since it is the glycan decorations, primarily the expression of sLex/a on these proteins (and lipids), that mediate this interaction of the CTC with the endothelial E-selectin, the contribution of GTs to migration and metastasis-supporting programs such as EMT needs to be better understood.

9.3.1.2 Perturbation in the Expression of GTs in Metastasis

It is widely accepted that abnormalities in glycosylation are closely associated with malignant transformation and progression [201203]. Accordingly, highly metastatic carcinomas (e.g., colon, breast, and prostate) exhibit high levels of sLex/a expression [74,180183,185,186,225]. sLex/a biosynthesis involves the concerted action of four GTs [68] and, hence, any dysregulation in GT expression and function has a direct impact on selectin-mediated metastasis. For instance, C2GnT is overexpressed in colorectal adenocarcinoma [248,249]. Lung carcinoma cells were found to acquire a colonizing phenotype upon FT-VII overexpression [188,189,250] and, in colon carcinoma, FucT-III gene silencing caused downregulation of sLex expression and subsequent inhibition of tumor proliferation and metastasis to the liver [251,252]. Matsuura et al. [74] investigated the expression of FTs and STs in breast cancer samples and cell lines and found that FT-VI and -III are significantly elevated, and that by expressing these FTs in breast cancer cells normally negative for sLex and sLea expression (MCF-7), rolling events are amplified on IL-1β activated HUVECs. One decade later, Barthel et al. reported similar results, i.e., elevated expression of FucT-III and -VI in the metastatic prostate cancer cell line, MDA-PCa-2b, compared with nonmetastatic cell lines. When PC3 cells, a prostate cancer cell line that is normally negative for E-selectin binding, were transfected with FucT-III, -VI or -VII, the cells gained sLex expression and the ability to roll on E-selectin-expressing endothelial cells and to metastasize to bone [71,201].

Moreover, it was demonstrated that the most highly expressed ST in patient samples of breast carcinoma was ST3Gal-III. This ST is responsible for sLea biosynthesis and its overexpression was associated with shorter overall survival [253,254]. Correspondingly, ST expression in metastasizing mammary tumors is higher compared with levels in nonmetastasizing tumors [255]. A recent report has indicated that induction of EMT, a prometastatic transformation program, was accompanied by elevation in levels of ST3Gal-I/-III/-IV and FT-III which are responsible for terminal capping of lactosamine units with sialic acid and fucose residues to form sLex/a moieties on colon cancer cells [209]. In combination, these studies indicate the fundamental role of GTs as modifiers of surface glycome features in supporting cancer metastasis and also reveal the potential significance of GT inhibitors as basic research and therapeutic tools.

9.3.2 Metabolic Inhibition of GTs

In common with all glycans, sLex/a is not directly encoded by the genome and its upregulation is achieved through overexpression of GTs responsible for its biosynthesis. The tetrasaccharide sLex/a requires four GT enzymes to catalyze its biosynthesis: N-acetylglucosaminyltransferase, β-galactosyltransferase, α-fucosyltransferase, and α-sialyltransferase. A detailed biosynthesis pathway of sLex/a and an explanation of the differences in glycan structure recognized by the three different selectins are outlined in Figure 9.2. Glycoside bond synthesis by the Leloir pathway involves the transfer of a sugar from an activated sugar nucleotide donor, nucleoside diphospho sugar (NDP sugar), or a nucleoside monophospho sugar (NMP sugar), to the hydroxyl group of an aglycone acceptor, catalyzed by the action of a GT enzyme. GT action may either have an overall retaining or inverting effect on the configuration at the anomeric center of the donor sugar [156]. sLex/a acts as a recognition marker for selectin and is involved in diverse biological processes including cell–cell adhesion, cell growth, cell differentiation, morphogenesis, leukocyte trafficking, and tumor dissemination [46,256]. In addition, aberration of glycan structure to sLex/a and expression occurs in various diseases including inflammatory disorders and tumor metastasis [68,204,257]. Consequently, GT inhibitors as therapeutic tools are promising in targeting diseases associated with surface glycoproteins such as inflammatory diseases and cancer metastasis. Alternatively, inhibitors of GTs provide valuable information about the role of glycans and glycoconjugates in biological processes through functional experiments in which the glycosylation of surface markers is perturbed and the cell surface glycoproteome is altered. Inhibitors are also used to study the active sites of GT enzymes and the molecular mechanisms of their action. Due to their biological and therapeutic significance, numerous inhibitors have been developed to target GTs.

9.3.2.1 GT Inhibitors and Primers

Natural GT inhibitors include papulacandin B [258], polyoxin D [259], tunicamycin [260], castanospermine [261], swainsonine [262], deoxymannojirimycin [263], and N-butyldeoxynojirimycin [264]. Natural inhibitors may cause extensive modifications in glycoconjugate structures such that it is difficult to study or control a specific interaction or enzyme; this introduces the necessity for the design and synthesis of selective agents. Synthetic GT inhibitors can be divided into two classes (refer to Figure 9.5). The first comprises substrate analogs which cause dead-end inhibition whereby a modified substrate is used to mimic one or more of the enzyme substrate(s) in order to block the enzyme or compete with natural substrates and terminate or delay glycan chain initiation or elongation. This class is further separated into three classes: donor substrate analogs (mimetics) of the donor sugar, nucleoside diphosphate, nucleoside monophosphate, or a diphosphate moiety. Because the mammalian system has only nine sugar nucleotide donors, the selectivity of GTs is thought to be associated with the identification of the acceptor rather than with the donor substrate. Within the group of donor analogs, diphosphate mimetics are thought to confer even less selectivity than sugar-modified donor analogs. In order to fulfill the selectivity requirement of GT inhibitors, acceptor mimetics have been developed. Acceptor substrate analogs compete with natural acceptors or block the active site of the enzyme to prevent the glycosyl transfer to the natural acceptor. Acceptor analogs usually have modifications at the hydroxyl group. The group may be deoxygenated, derivatized, or replaced with a halogen or other functionalities. Bisubstrate analogs have two covalently linked motifs mimicking both the acceptor and donor substrates. These analogs are also referred to as transition state analogs because they mimic the transition state of the reaction, and they are thought to be the most active and selective inhibitors of GTs. The second class includes GT primers and metabolic decoys. In contrast to the dead-end GT inhibitors, GT primers prime the enzyme activity using irrelevant substrates leading to the consumption of free active enzyme and the production of irrelevant products. Products of the priming reaction may be either inactive or metabolic decoys, or further modified with downstream enzymes to produce metabolic decoys capable of inhibiting a target pathway or ligand. GT synthetic inhibitors will be discussed in more detail in the following section. For simplicity, the first class will be designated as “inhibitors” and the second class as “primers.”

image
Figure 9.5 Mode of action of GT modulators. Glycan analogs modulate the action of GTs through either inhibitory or priming activity. GT inhibitors are enzyme substrate analogs that mimic one or more of the substrate analogs to block the enzyme or compete with natural substrate leading to termination of natural glycan biosynthesis pathway. GT primers prime the enzyme activity using irrelevant substrate leading to consumption of the enzyme activity in the production of irrelevant products and diversion from the natural glycan biosynthesis pathway.

9.3.2.2 GT Substrate-Analogous Inhibitors

As this chapter is focused on selectin ligand prototype sLex/a, this section in turn will discuss inhibitors against the key GTs involved in its biosynthesis (Figure 9.2) and with which sLex/a expression and/or selectin ligand binding were reported to be affected: C2GnT, FT, and ST.

C2GnT inhibitors: C2GnT (β(1,3)-galactosyl-O-glycosyl-glycoprotein β(1,6)-N-acetylglucosaminyltransferase; EC 2.4.1.102) is an inverting GT that catalyzes the transfer of a GlcNAc residue from UDP-αGlcNAc to β(1,6) linkage on core 1 structure (Galβ1,3GalNAcα-O-pp) to form core 2 branching, one of the major mucin-type O-glycan core structures [265]. An inhibitor to C2GnT GT was developed based on a deoxygenated analog of the enzyme acceptor (Galβ1,3,6-deoxy-GalNAcα-O(CH2)8COOCH3) which produced moderate inhibition compared with the corresponding acceptor (Galβ1,3GalNAcα-O(CH2)8COOCH3) in vitro [266]. A recent study showed that 5-thio-GlcNAc could act as a donor substrate analog against the C2GnT enzyme. This inhibitor is active in vitro and the first to be reported as active in vivo. Although it caused a decrease in levels of cellular O-GlcNAc, the surface glycan profile did not change. In vivo activity is thought to be due to peracetylation of 5T-GlcNAc, which overcomes the problem of poor diffusion of charged analogs. The peracetylated prodrug is converted to the 5T-GlcNAc by the intracellular estrases and, in turn, is converted to UDP-5T-GlcNAc through the salvage pathway [267].

FUT inhibitors: α1,3-fucosyltransferase (FucT-IV, -V, -VI, -VII; EC 2.4.1.152) is an inverting enzyme that catalyzes the transfer of the fucosyl moiety from GDP-fucose to the 3-OH of group of N-acetyllactosamine to give a Lewis X trisaccharide. α1,3/4-fucosyltransferase (EC 2.4.1.65) transfers a fucose sugar from GDP-fucose to the 4-OH of Galβ1,3GlcNAc moiety to give a Lewis A trisaccharide. FTs can act on sialylated LacNAc to give sLex/a tetrasaccharides [268].

As for 5T-GlcNAc, the same research group developed 5-thio-fucose and evaluated its activity against sLex expression and selectin binding activity. Peracetylated 5T-Fuc was taken up by the relevant cells and converted through a salvage pathway into a sugar nucleotide donor analog, GDP-5T-Fuc that, in turn, inhibited the addition of fucose by the FT enzyme. 5T-Fuc was able to inhibit sLex expression in HepG2 cells and impaired adhesion of the cells to selectin-coated surfaces, and activated endothelial monolayer cells [269]. A cell-permeable acetylated analog of a fucose bearing a fluorine atom near the endocyclic oxygen, 2-fluoro-fucose (2F-Fuc), was reported to exhibit global inhibition of FTs. After being deacetylated intracellularly, this fluorinated analog is converted to the corresponding nucleotide sugar substrate GDP-2-fluoro-Fuc by the salvage pathway and is able to inhibit sLex formation in a human myeloid cell line and, subsequently, to hinder E-selectin- and P-selectin-mediated rolling on coated surfaces [270]. A novel study investigated the oral bioavailability of 2F-Fuc in a murine model. 2F-Fuc inhibited protein fucosylation, neutrophil sLex expression, and selectin-mediated adhesion during oral administration with no apparent toxicity in mice over the course of 3 weeks. Inhibitory activities were reversed upon discontinuing the drug administration. 2F-Fuc was also found to inhibit tumor outgrowth in colon and renal cancer metastasis mouse models after oral administration. This study supports the development and potential use of clinically effective GT inhibitors [179,180] in the treatment of cancer and prevention of metastasis.

ST inhibitors: The α2,3-sialyltransferase enzyme (ST3Gal) is a retaining enzyme that catalyzes the transfer of sialic acid from cytidine-5′-monophospho-N-acetylneuraminic acid (CMP-Neu5Ac) to the terminal Galβ1,4GlcNAc (EC 2.4.99.6) or Galβ1,3GlcNAc. Donor-analogous substrates were developed to mimic the common donor substrate, CMP-Neu5Ac, by replacing the anomeric carbon-linked oxygen atom with an ethyl bridge. This analog caused significant inhibition as it probably prevented the nucleophilic attack by the enzyme basic group [271]. Recently, using a cell-permeable acetylated analog of a sialic acid bearing a fluorine atom near the endocyclic oxygen, a global inhibition of ST activity was reported. After being deacetylated intracellularly, this fluorinated analog is converted to the corresponding nucleotide sugar substrate, CMP-2-fluoro-NeuAc, by the salvage pathway. This analog is then able to inhibit sLex formation in a human myeloid cell line as well as E-selectin- and P-selectin-mediated rolling on coated surfaces. It is thought that the observed inhibition is due to both enzyme and feedback inhibition of de novo synthesis of CMP-Neu5Ac by the salvage pathway and, to date, this is the only α2,3-sialyltransferase analog that has been found to produce inhibitory effects in vivo [270,272]. Acceptor-analogous inhibitors were reported to inhibit rat liver α2,3-sialyltransferase activity in vitro using a trisaccharide analog with a modified 3′-OH. The modifications include deoxygenation, fluorination, amination, or methoxy derivatization [273]. The 3′-deoxy lactosamine derivative was also reported, but with no inhibition activity in vitro [266]. Bisubstrate analogs containing the donor CMP-Neu5Ac mimetic and galactose, lactose or lactosamine acceptor have been developed and found to produce weak inhibitory effects against α2,3-sialyltransferase in the case of galactose- and lactose-containing acceptors, whereas they produced potent effects in the case of lactosamine-containing analogs [271,274].

9.3.2.3 GT Primers and Metabolic Decoys

Traditional GT inhibitors are analogs with alterations in specific functionalities that affect the thermodynamics and kinetics of the reaction. These analogs can still bind to the enzyme, but cannot undergo covalent modifications to give reaction products (dead-end inhibition). Conversely, priming activity is achieved when the modified analog binds the active enzyme and undergoes covalent modification to yield modified soluble glycan. Modified glycans can be inactive and thus inhibitory by only the consumption of the active enzyme in an extraneous pathway diverting glycan synthesis from endogenous glycoconjugates (see Figure 9.5). Otherwise, the modified glycan can be biologically active so that it acts as a metabolic decoy, or further modified by the biosynthesis pathway enzymes to a metabolic decoy that can inhibit the glycoconjugate formation or block an effector target. Many of the well-known inhibitors are inactive in vivo probably due to poor permeation into cells and subcellular compartments such as the Golgi apparatus where most glycan formation takes place. Primers can circumvent this problem by conjugating the substrate to a hydrophobic aglycone that enables passive diffusion through the lipophilic cell membrane without affecting the binding kinetics to the target enzyme. Conjugating specific aglycones may also facilitate the isolation of reaction products for further study or create probes for the study of active sites. Also, heterogeneity between cell types in glycosylation patterns can be studied through analysis of the primed products. The monosaccharide glycoside paranitrophenol α-GalNAc (pNp α-GalNAc) was used to inhibit mucin glycosylation through priming modified lactosamine formation and extension in colon cancer and lymphoid tumor cells [275,276]. PNp α-GalNAc is an acceptor analog for β1,3-galactosyltransferase that transfers a Gal residue from UDP-Gal to GalNAc to form core 1 structure. The core 1 structure (Gal β1,3GalNAc) acts as a molecular scaffold for mucin decorations with glycan side chains (i.e., core 2 branching, lactosamine extension, and sLex/a formation). PNp α-GalNAc could enter viable cells, probably due to the hydrophobic aryl group, and consume glycosylation machinery from mucin decoration toward the priming reaction. The aryl group was also used to isolate and secrete the priming reaction products Galβ1,3GalNAcα-pNp and Galβ1,3(Galβ1,4GlcNAcβ-6)GalNAcα-pNp, and to study the glycosylation patterns in these cells. Two independent groups who used GalNAcα-O-benzyl as a primer for such pathways [277279] also reported similar results. Likewise, several hydrophobic glycosides of GlcNAc, GlcNAcα-O-benzyl, GlcNAcβ-O-benzyl, GlcNAcβ-O-phenyl, and GlcNAcβ-O-pnp could act as primers for polylactosamine chain synthesis [280]. More recently, fluorinated peracetylated GalNAc was reported to be incorporated in selectin ligands expressed by human promyelocytic leukemia cells (HL-60) suggesting a metabolic priming activity and inhibition of glycan chain elongation through the blockade of carbon 4 with a fluorine atom, so preventing the addition of either fucose or galactose sugars. Addition of 4F-GalNAc to the culture medium caused a significant decrease in sLex expression and leukocyte adhesion in vivo [281]. Another study confirmed the activity of 4F-GlcNAc in inhibiting sLex expression and selectin binding. However, this research showed that there was no incorporation of 4F-GlcNAc into the growing glycan chain, suggesting that the main mechanism of 4F-GlcNAc activity lies in interference with UDP-GlcNAc biosynthesis and not in the priming and incorporation of 4F-GlcNAc [282].

Disaccharide glycosides can also be used as primers for GTs. GlcNAcβ1,3Galβ-O-naphthalenemethanol (GlcNAcβ1,3Galβ-O-NM) and GlcNAcβ1,4Galβ-O-NM inhibited sLex expression in leukemia, colon cancer, and lung cancer models, and subsequently reduced E-selectin-dependent cell adhesion to activated endothelial cells [283286]. Furthermore, it was shown that the treatment of colon and lung adenocarcinoma with the peracetylated disaccharide primer (Ac)6GlcNAcβ1,3Galβ-O-NM reduced the potential for lung colonization in immunodeficient mice. Acetylated derivatives are thought to be more readily bioavailable to cells and, hence, more active in vivo as the polar hydroxyl groups are masked. Acetyl groups are then removed by the action of intracellular and membrane bound esterases [287]. 4-deoxy derivative of (Ac)6GlcNAcβ1,3Galβ-O-NM acted as an acceptor-analogous substrate against β1,4-galactosyltransferase and inhibited experimental tumor metastasis by Lewis lung carcinoma in vivo [288].

The development of a diverse group of GT inhibitors as modulators of surface selectin glycoconjugates holds great promise for the treatment of metastatic tumors and inflammatory disorders. Although this field is several decades old, most of the inhibitors known lack activity in vivo due to their polar nature and underprivileged cell permeability. Recently, several reports have indicated that this problem can be avoided by using the prodrug technology whereby peracetylated derivatives show enhanced uptake. Peracetylated prodrugs are easily deacetylated intracellularly to give the active form of the drug. Another approach is to use bulky hydrophobic aglycones to confer hydrophobic properties on the inhibitor candidate. Nevertheless the development of GT inhibitors with acceptable bioavailability remains a goal in the development of clinically effective antimetastasis therapeutic agents. On the other hand, since both physiological and pathological cell players share the sLex decoration, increased efforts should be directed toward to the development of selective agents that target-specific GT isoforms that are differentially expressed in CTCs but not in normal immune cells.

9.4 Use of Enzymes to Improve the Therapeutic Function of Cells Through Modification of the Cell Surface

Other types of enzymes have been used or are potentially available for ex vivo enzymatic treatment of the cell surface with the aim of therapeutic applications. One such example is the transient proteolytic treatment of MSCs to improve the efficiency of delivery to the target sites [289,290]. When a systemic infusion is conducted for the delivery of MSCs, one major obstacle is the natural tendency of the injected cells to accumulate initially in the lung, and this may lower the efficiency of delivery of the MSCs to the target sites. Use of pronase rather than the standard use of trypsin for the preparation of the cell suspension results in increased lung clearance of MSCs and a better delivery to the site of inflammation, without any apparent functional defect, presumably due to the enhanced cleavage of cell surface proteins involved in adhesion and migration.

In another example, O-sialoglycoprotein endopeptidase (OSGE) treatment has been shown to enhance the functionality of aged CD4 T cells [291]. In humans and mice, the function of the antibody-mediated immune system becomes limited with age, there being lowered clonal diversity of antibodies and inadequate formation of long-term memory [292], due in part to age-related defects in CD4 T cells, especially a malfunction in the interaction between T cells and antigen presenting cells (APCs) [293]. The treatment of aged CD4 cells with OSGE ex vivo restored the age-dependent loss of function, possibly through a mechanism that removes glycoproteins resulting in increased access to the CD28 molecules for interaction with APCs [291].

9.5 Conclusion

This chapter has discussed in detail that the adhesive recruitment of both therapeutic stem cells as well as detrimental metastatic cancer cells are initiated by glycoprotein–selectin interactions facilitating cell rolling on endothelium. The formation of the key glycan responsible for mediating these interactions, sLex, is controlled by the precise activity of a family of ER/Golgi localized enzymes, the GTs. Through the identification and manipulation of these enzymes, migration can be effectively controlled either to enhance or inhibit it. In applications dealing with enabling a remedial cell to migrate to a predetermined site in vivo, the sLex structure needs to be created by the sequential activity of GTs under conditions that do not alter the viability, metabolism, or physiology of that cell, such as through the application of “GPS” technology. Multiple therapeutic avenues can then be sought including the guiding of stem cells or cell-based drug delivery systems to selectin expressing endothelial cells intrinsic to autoimmune or inflammatory diseases. In contrast, for applications where the focus is to inhibit migration/metastasis of detrimental cancer cells or immune cells in inflammatory disease, this interaction between selectins and glycans can be perturbed by inhibiting the activity of these enzymes. Since every family of GT has numerous members, identifying unique enzymes that are responsible for contributing to sLex formation in these harmful cells, but not in “normal” cells, and engineering selective inhibitors to these unique GTs, the migration of the harmful cells could potentially be targeted. Most of the inhibitors known to date lack this specificity and in addition lack activity in vivo due to their polar nature and underprivileged cell permeability. Although recent approaches have helped to overcome membrane permeabilization issues, the development of inhibitors with acceptable bioavailability and specificity remains a goal in the development of clinically effective antimetastasis therapeutic agents.

References

1. Ehrhardt C, Kneuer C, Bakowsky U. Selectins—an emerging target for drug delivery. Adv Drug Deliv Rev. 2004;56:527–549.

2. Butcher EC. Leukocyte-endothelial cell recognition: three (or more) steps to specificity and diversity. Cell. 1991;67:1033–1036.

3. Butcher EC, Picker LJ. Lymphocyte homing and homeostasis. Science. 1996;272:60–66.

4. Springer TA. Traffic signals for lymphocyte recirculation and leukocyte emigration: the multistep paradigm. Cell. 1994;76:301–314.

5. Lowe JB. Glycosylation in the control of selectin counter-receptor structure and function. Immunol Rev. 2002;186:19–36.

6. Vestweber D, Blanks JE. Mechanisms that regulate the function of the selectins and their ligands. Physiol Rev. 1999;79:181–213.

7. Kunkel EJ, Butcher EC. Chemokines and the tissue-specific migration of lymphocytes. Immunity. 2002;16:1–4.

8. Rossi D, Zlotnik A. The biology of chemokines and their receptors. Annu Rev Immunol. 2000;18:217–242.

9. Sallusto F, Mackay CR, Lanzavecchia A. The role of chemokine receptors in primary, effector, and memory immune responses. Annu Rev Immunol. 2000;18:593–620.

10. Cinamon G, Grabovsky V, Winter E, et al. Novel chemokine functions in lymphocyte migration through vascular endothelium under shear flow. J Leukoc Biol. 2001;69:860–866.

11. Harris ES, McIntyre TM, Prescott SM, Zimmerman GA. The leukocyte integrins. J Biol Chem. 2000;275:23409–23412.

12. van Kooyk Y, Figdor CG. Avidity regulation of integrins: the driving force in leukocyte adhesion. Curr Opin Cell Biol. 2000;12:542–547.

13. Woodside DG, Liu S, Ginsberg MH. Integrin activation. Thromb Haemost. 2001;86:316–323.

14. Weber C. Novel mechanistic concepts for the control of leukocyte transmigration: specialization of integrins, chemokines, and junctional molecules. J Mol Med. 2003;81:4–19.

15. Lawrence MB, Springer TA. Leukocytes roll on a selectin at physiologic flow rates: distinction from and prerequisite for adhesion through integrins. Cell. 1991;65:859–873.

16. Sackstein R. The lymphocyte homing receptors: gatekeepers of the multistep paradigm. Curr Opin Hematol. 2005;12:444–450.

17. Thankamony SP, Sackstein R. Enforced hematopoietic cell E- and L-selectin ligand (HCELL) expression primes transendothelial migration of human mesenchymal stem cells. Proc Natl Acad Sci USA. 2011;108:2258–2263.

18. Yago T, Shao B, Miner JJ, et al. E-selectin engages PSGL-1 and CD44 through a common signaling pathway to induce integrin alphaLbeta2-mediated slow leukocyte rolling. Blood. 2010;116:485–494.

19. Gibson RM, Kansas GS, Tedder TF, Furie B, Furie BC. Lectin and epidermal growth factor domains of P-selectin at physiologic density are the recognition unit for leukocyte binding. Blood. 1995;85:151–158.

20. Kansas GS, Saunders KB, Ley K, et al. A role for the epidermal growth factor-like domain of P-selectin in ligand recognition and cell adhesion. J Cell Biol. 1994;124(4):609–618.

21. Wenzel K, Felix S, Kleber FX, et al. E-selectin polymorphism and atherosclerosis: an association study. Hum Mol Genet. 1994;3:1935–1937.

22. Wenzel K, Hanke R, Speer A. Polymorphism in the human E-selectin gene detected by PCR-SSCP. Hum Genet. 1994;94:452–453.

23. Polley MJ, Phillips ML, Wayner E, et al. CD62 and endothelial cell-leukocyte adhesion molecule 1 (ELAM-1) recognize the same carbohydrate ligand, sialyl-Lewis x. Proc Natl Acad Sci USA. 1991;88:6224–6228.

24. Alon R, Feizi T, Yuen CT, Fuhlbrigge RC, Springer TA. Glycolipid ligands for selectins support leukocyte tethering and rolling under physiologic flow conditions. J Immunol. 1995;154:5356–5366.

25. Fuhlbrigge RC, Kieffer JD, Armerding D, Kupper TS. Cutaneous lymphocyte antigen is a specialized form of PSGL-1 expressed on skin-homing T cells. Nature. 1997;389:978–981.

26. Leppanen A, White SP, Helin J, McEver RP, Cummings RD. Binding of glycosulfopeptides to P-selectin requires stereospecific contributions of individual tyrosine sulfate and sugar residues. J Biol Chem. 2000;275:39569–39578.

27. Rosen SD. Ligands for L-selectin: homing, inflammation, and beyond. Annu Rev Immunol. 2004;22:129–156.

28. Tedder TF, Steeber DA, Chen A, Engel P. The selectins: vascular adhesion molecules. FASEB J. 1995;9:866–873.

29. Bargatze RF, Kurk S, Butcher EC, Jutila MA. Neutrophils roll on adherent neutrophils bound to cytokine-induced endothelial cells via L-selectin on the rolling cells. J Exp Med. 1994;180:1785–1792.

30. Walcheck B, Moore KL, McEver RP, Kishimoto TK. Neutrophil–neutrophil interactions under hydrodynamic shear stress involve L-selectin and PSGL-1 A mechanism that amplifies initial leukocyte accumulation of P-selectin in vitro. J Clin Invest. 1996;98:1081–1087.

31. Kunkel EJ, Chomas JE, Ley K. Role of primary and secondary capture for leukocyte accumulation in vivo. Circ Res. 1998;82:30–38.

32. Mitchell DJ, Li P, Reinhardt PH, Kubes P. Importance of L-selectin-dependent leukocyte–leukocyte interactions in human whole blood. Blood. 2000;95:2954–2959.

33. Rossi FMV, Corbel SY, Merzaban JS, et al. Recruitment of adult thymic progenitors is regulated by P-selectin and its ligand PSGL-1. Nat Immunol. 2005;6:262–270.

34. Kivisakk P, Mahad DJ, Callahan MK, et al. Human cerebrospinal fluid central memory CD4+ T cells: evidence for trafficking through choroid plexus and meninges via P-selectin. Proc Natl Acad Sci USA. 2003;100:8389–8394.

35. Jacobsen K, Kravitz J, Kincade PW, Osmond DG. Adhesion receptors on bone marrow stromal cells: in vivo expression of vascular cell adhesion molecule-1 by reticular cells and sinusoidal endothelium in normal and gamma-irradiated mice. Blood. 1996;87:73–82.

36. Schweitzer KM, Drager AM, van der Valk P, et al. Constitutive expression of E-selectin and vascular cell adhesion molecule-1 on endothelial cells of hematopoietic tissues. Am J Pathol. 1996;148:165–175.

37. Weninger W, Ulfman LH, Cheng G, et al. Specialized contributions by alpha(1,3)-fucosyltransferase-IV and FucT-VII during leukocyte rolling in dermal microvessels. Immunity. 2000;12:665–676.

38. Tchernychev B, Furie B, Furie BC. Peritoneal macrophages express both P-selectin and PSGL-1. J Cell Biol. 2003;163:1145–1155.

39. Yao L, Pan J, Setiadi H, Patel KD, McEver RP. Interleukin 4 or oncostatin M induces a prolonged increase in P-selectin mRNA and protein in human endothelial cells. J Exp Med. 1996;184:81–92.

40. Woltmann G, McNulty CA, Dewson G, Symon FA, Wardlaw AJ. Interleukin-13 induces PSGL-1/P-selectin-dependent adhesion of eosinophils, but not neutrophils, to human umbilical vein endothelial cells under flow. Blood. 2000;95:3146–3152.

41. Hattori R, Hamilton KK, Fugate RD, McEver RP, Sims PJ. Stimulated secretion of endothelial von Willebrand factor is accompanied by rapid redistribution to the cell surface of the intracellular granule membrane protein GMP-140. J Biol Chem. 1989;264:7768–7771.

42. Pan J, Xia L, McEver RP. Comparison of promoters for the murine and human P-selectin genes suggests species-specific and conserved mechanisms for transcriptional regulation in endothelial cells. J Biol Chem. 1998;273:10058–10067.

43. Liu Z, Miner JJ, Yago T, et al. Differential regulation of human and murine P-selectin expression and function in vivo. J Exp Med. 2010;207:2975–2987.

44. Keelan ET, Licence ST, Peters AM, Binns RM, Haskard DO. Characterization of E-selectin expression in vivo with use of a radiolabeled monoclonal antibody. Am J Physiol. 1994;266:H278–H290.

45. Frenette PS, Subbarao S, Mazo IB, von Andrian UH, Wagner DD. Endothelial selectins and vascular cell adhesion molecule-1 promote hematopoietic progenitor homing to bone marrow. Proc Natl Acad Sci USA. 1998;95:14423–14428.

46. Kansas GS. Selectins and their ligands: current concepts and controversies. Blood. 1996;88:3259–3287.

47. Kunkel EJ, Ley K. Distinct phenotype of E-selectin-deficient mice E-selectin is required for slow leukocyte rolling in vivo. Circ Res. 1996;79:1196–1204.

48. Kunkel EJ, Dunne JL, Ley K. Leukocyte arrest during cytokine-dependent inflammation in vivo. J Immunol. 2000;164:3301–3308.

49. Hwang JM, Yamanouchi J, Santamaria P, Kubes P. A critical temporal window for selectin-dependent CD4+ lymphocyte homing and initiation of late-phase inflammation in contact sensitivity. J Exp Med. 2004;199:1223–1234.

50. Subramaniam M, Koedam JA, Wagner DD. Divergent fates of P- and E-selectins after their expression on the plasma membrane. Mol Biol Cell. 1993;4:791–801.

51. Bevilacqua MP, Stengelin S, Gimbrone Jr MA, Seed B. Endothelial leukocyte adhesion molecule 1: an inducible receptor for neutrophils related to complement regulatory proteins and lectins. Science. 1989;243:1160–1165.

52. Picker LJ, Michie SA, Rott LS, Butcher EC. A unique phenotype of skin-associated lymphocytes in humans Preferential expression of the HECA-452 epitope by benign and malignant T cells at cutaneous sites. Am J Pathol. 1990;136:1053–1068.

53. Berg EL, Yoshino T, Rott LS, et al. The cutaneous lymphocyte antigen is a skin lymphocyte homing receptor for the vascular lectin endothelial cell-leukocyte adhesion molecule 1. J Exp Med. 1991;174:1461–1466.

54. Sipkins DA, Wei X, Wu JW, et al. In vivo imaging of specialized bone marrow endothelial microdomains for tumour engraftment. Nature. 2005;435:969–973.

55. Lo Celso C, Fleming HE, Wu JW, et al. Live-animal tracking of individual haematopoietic stem/progenitor cells in their niche. Nature. 2009;457:92–96.

56. Sackstein R. The bone marrow is akin to skin: HCELL and the biology of hematopoietic stem cell homing. J Invest Dermatol. 2004;122:1061–1069.

57. Frenette PS, Mayadas TN, Rayburn H, Hynes RO, Wagner DD. Susceptibility to infection and altered hematopoiesis in mice deficient in both P- and E-selectins. Cell. 1996;84:563–574.

58. Mazo IB, Gutierrez-Ramos JC, Frenette PS, Hynes RO, Wagner DD, von Andrian UH. Hematopoietic progenitor cell rolling in bone marrow microvessels: parallel contributions by endothelial selectins and vascular cell adhesion molecule 1. J Exp Med. 1998;188:465–474.

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

60. Katayama Y, Hidalgo A, Furie BC, Vestweber D, Furie B, Frenette PS. PSGL-1 participates in E-selectin-mediated progenitor homing to bone marrow: evidence for cooperation between E-selectin ligands and alpha4 integrin. Blood. 2003;102:2060–2067.

61. Hidalgo A, Weiss LA, Frenette PS. Functional selectin ligands mediating human CD34+ cell interactions with bone marrow endothelium are enhanced postnatally. J Clin Invest. 2002;110:559–569.

62. Dimitroff CJ, Lee JY, Rafii S, Fuhlbrigge RC, Sackstein R. CD44 is a major E-selectin ligand on human hematopoietic progenitor cells. J Cell Biol. 2001;153:1277–1286.

63. Levesque JP, Zannettino AC, Pudney M, et al. PSGL-1-mediated adhesion of human hematopoietic progenitors to P-selectin results in suppression of hematopoiesis. Immunity. 1999;11:369–378.

64. Winkler IG, Snapp KR, Simmons PJ, Levesque JP. Adhesion to E-selectin promotes growth inhibition and apoptosis of human and murine hematopoietic progenitor cells independent of PSGL-1. Blood. 2004;103:1685–1692.

65. Winkler IG, Barbier V, Nowlan B, et al. Vascular niche E-selectin regulates hematopoietic stem cell dormancy, self renewal and chemoresistance. Nat Med. 2012;18:1651–1657.

66. Manfredini R, Zini R, Salati S, et al. The kinetic status of hematopoietic stem cell subpopulations underlies a differential expression of genes involved in self-renewal, commitment, and engraftment. Stem Cells. 2005;23:496–506.

67. Kim MY, Oskarsson T, Acharyya S, et al. Tumor self-seeding by circulating cancer cells. Cell. 2009;139:1315–1326.

68. Barthel SR, Gavino JD, Descheny L, Dimitroff CJ. Targeting selectins and selectin ligands in inflammation and cancer. Expert Opin Ther Targets. 2007;11:1473–1491.

69. Geng Y, Marshall JR, King MR. Glycomechanics of the metastatic cascade: tumor cell–endothelial cell interactions in the circulation. Ann Biomed Eng. 2012;40:790–805.

70. Khatib AM, Fallavollita L, Wancewicz EV, Monia BP, Brodt P. Inhibition of hepatic endothelial E-selectin expression by C-raf antisense oligonucleotides blocks colorectal carcinoma liver metastasis. Cancer Res. 2002;62:5393–5398.

71. Barthel SR, Wiese GK, Cho J, et al. Alpha 1,3 fucosyltransferases are master regulators of prostate cancer cell trafficking. Proc Natl Acad Sci USA. 2009;106:19491–19496.

72. Burdick MM, Henson KA, Delgadillo LF, et al. Expression of E-selectin ligands on circulating tumor cells: cross-regulation with cancer stem cell regulatory pathways? Front Oncol. 2012;2:103.

73. Kannagi R, Izawa M, Koike T, Miyazaki K, Kimura N. Carbohydrate-mediated cell adhesion in cancer metastasis and angiogenesis. Cancer Sci. 2004;95:377–384.

74. Matsuura N, Narita T, Hiraiwa N, et al. Gene expression of fucosyl- and sialyl-transferases which synthesize sialyl Lewis x, the carbohydrate ligands for E-selectin, in human breast cancer. Int J Oncol. 1998;12:1157–1164.

75. Renkonen J, Paavonen T, Renkonen R. Endothelial and epithelial expression of sialyl Lewis(x) and sialyl Lewis(a) in lesions of breast carcinoma. J Int Cancer. 1997;74:296–300.

76. Czlapinski JL, Bertozzi CR. Synthetic glycobiology: exploits in the Golgi compartment. Curr Opin Chem Biol. 2006;10:645–651.

77. Gadhoum SZ, Sackstein R. CD15 expression in human myeloid cell differentiation is regulated by sialidase activity. Nat Chem Biol. 2008;4:751–757.

78. Tyrrell D, James P, Rao N, et al. Structural requirements for the carbohydrate ligand of E-selectin. Proc Natl Acad Sci USA. 1991;88:10372–10376.

79. Sackstein R. Glycosyltransferase-programmed stereosubstitution (GPS) to create HCELL: engineering a roadmap for cell migration. Immunol Rev. 2009;230:51–74.

80. Jacobs PP, Sackstein R. CD44 and HCELL: preventing hematogenous metastasis at step 1. FEBS Lett. 2011;585:3148–3158.

81. Smith PL, Myers JT, Rogers CE, et al. Conditional control of selectin ligand expression and global fucosylation events in mice with a targeted mutation at the FX locus. J Cell Biol. 2002;158:801–815.

82. Becker DJ, Lowe JB. Fucose: biosynthesis and biological function in mammals. Glycobiology. 2003;13:41R–53R.

83. Ma B, Simala-Grant JL, Taylor DE. Fucosylation in prokaryotes and eukaryotes. Glycobiology. 2006;16:158R–184R.

84. Smithson G, Rogers CE, Smith PL, et al. Fuc-TVII is required for T helper 1 and T cytotoxic 1 lymphocyte selectin ligand expression and recruitment in inflammation, and together with Fuc-TIV regulates naive T cell trafficking to lymph nodes. J Exp Med. 2001;194:601–614.

85. Maly P, Thall A, Petryniak B, et al. The alpha(1,3)fucosyltransferase Fuc-TVII controls leukocyte trafficking through an essential role in L-, E-, and P-selectin ligand biosynthesis. Cell. 1996;86:643–653.

86. Homeister JW, Thall AD, Petryniak B, et al. The alpha(1,3)fucosyltransferases FucT-IV and FucT-VII exert collaborative control over selectin-dependent leukocyte recruitment and lymphocyte homing. Immunity. 2001;15:115–126.

87. Knibbs RN, Craig RA, Natsuka S, et al. The fucosyltransferase FucT-VII regulates E-selectin ligand synthesis in human T cells. J Cell Biol. 1996;133:911–920.

88. Knibbs RN, Craig RA, Maly P, et al. Alpha(1,3)-fucosyltransferase VII-dependent synthesis of P- and E-selectin ligands on cultured T lymphoblasts. J Immunol. 1998;161:6305–6315.

89. Bengtson P, Lundblad A, Larson G, Pahlsson P. Polymorphonuclear leukocytes from individuals carrying the G329A mutation in the alpha 1,3-fucosyltransferase VII gene (FUT7) roll on E- and P-selectins. J Immunol. 2002;169:3940–3946.

90. Huang MC, Laskowska A, Vestweber D, Wild MK. The alpha (1,3)-fucosyltransferase Fuc-TIV, but not Fuc-TVII, generates sialyl Lewis X-like epitopes preferentially on glycolipids. J Biol Chem. 2002;277:47786–47795.

91. Burdick MM, McCaffery JM, Kim YS, Bochner BS, Konstantopoulos K. Colon carcinoma cell glycolipids, integrins, and other glycoproteins mediate adhesion to HUVECs under flow. Am J Physiol Cell Physiol. 2003;284:C977–C987.

92. Nimrichter L, Burdick MM, Aoki K, et al. E-selectin receptors on human leukocytes. Blood. 2008;112:3744–3752.

93. Shirure VS, Henson KA, Schnaar RL, Nimrichter L, Burdick MM. Gangliosides expressed on breast cancer cells are E-selectin ligands. Biochem Biophys Res Commun. 2011;406:423–429.

94. Mondal N, Buffone Jr A, Neelamegham S. Distinct glycosyltransferases synthesize E-selectin ligands in human vs mouse leukocytes. Cell Adh Migr. 2013;7:288–292.

95. Buffone Jr A, Mondal N, Gupta R, McHugh KP, Lau JT, Neelamegham S. Silencing alpha1,3-fucosyltransferases in human leukocytes reveals a role for FUT9 enzyme during E-selectin-mediated cell adhesion. J Biol Chem. 2013;288:1620–1633.

96. Merzaban JS, Burdick MM, Gadhoum SZ, et al. Analysis of glycoprotein E-selectin ligands on human and mouse marrow cells enriched for hematopoietic stem/progenitor cells. Blood. 2011;118:1774–1783.

97. Ellies LG, Sperandio M, Underhill GH, et al. Sialyltransferase specificity in selectin ligand formation. Blood. 2002;100:3618–3625.

98. Okajima T, Fukumoto S, Miyazaki H, et al. Molecular cloning of a novel alpha2,3-sialyltransferase (ST3Gal VI) that sialylates type II lactosamine structures on glycoproteins and glycolipids. J Biol Chem. 1999;274:11479–11486.

99. Sasaki K, Watanabe E, Kawashima K, et al. Expression cloning of a novel Gal beta (1–3/1–4) GlcNAc alpha 2,3-sialyltransferase using lectin resistance selection. J Biol Chem. 1993;268:22782–22787.

100. Yang WH, Nussbaum C, Grewal PK, Marth JD, Sperandio M. Coordinated roles of ST3Gal-VI and ST3Gal-IV sialyltransferases in the synthesis of selectin ligands. Blood. 2012;120:1015–1026.

101. Sperandio M, Frommhold D, Babushkina I, et al. Alpha 2,3-sialyltransferase-IV is essential for L-selectin ligand function in inflammation. Eur J Immunol. 2006;36:3207–3215.

102. Kono M, Ohyama Y, Lee YC, Hamamoto T, Kojima N, Tsuji S. Mouse beta-galactoside alpha 2,3-sialyltransferases: comparison of in vitro substrate specificities and tissue specific expression. Glycobiology. 1997;7:469–479.

103. Schachter H, Brockhausen I. The biosynthesis of branched O-glycans. Symp Soc Exp Biol. 1989;43:1–26.

104. Ellies LG, Tsuboi S, Petryniak B, Lowe JB, Fukuda M, Marth JD. Core 2 oligosaccharide biosynthesis distinguishes between selectin ligands essential for leukocyte homing and inflammation. Immunity. 1998;9:881–890.

105. Snapp KR, Heitzig CE, Ellies LG, Marth JD, Kansas GS. Differential requirements for the O-linked branching enzyme core 2 beta1-6-N-glucosaminyltransferase in biosynthesis of ligands for E-selectin and P-selectin. Blood. 2001;97:3806–3811.

106. Sperandio M, Thatte A, Foy D, Ellies LG, Marth JD, Ley K. Severe impairment of leukocyte rolling in venules of core 2 glucosaminyltransferase-deficient mice. Blood. 2001;97:3812–3819.

107. Merzaban JS, Zuccolo J, Corbel SY, Williams MJ, Ziltener HJ. An alternate core 2 {beta}1,6-N-acetylglucosaminyltransferase selectively contributes to P-selectin ligand formation in activated CD8 T cells. J Immunol. 2005;174:4051–4059.

108. Mullaly SC, Oudhoff MJ, Min PH, et al. Requirement for core 2 O-glycans for optimal resistance to helminth infection. PLoS ONE. 2013;8:e60124.

109. Huang MC, Zollner O, Moll T, et al. P-selectin glycoprotein ligand-1 and E-selectin ligand-1 are differentially modified by fucosyltransferases Fuc-TIV and Fuc-TVII in mouse neutrophils. J Biol Chem. 2000;275:31353–31360.

110. de Vries T, Knegtel RM, Holmes EH, Macher BA. Fucosyltransferases: structure/function studies. Glycobiology. 2001;11:119R–128R.

111. Mollicone R, Moore SE, Bovin N, et al. Activity, splice variants, conserved peptide motifs, and phylogeny of two new alpha1,3-fucosyltransferase families (FUT10 and FUT11). J Biol Chem. 2009;284:4723–4738.

112. Padro M, Cobler L, Garrido M, de Bolos C. Down-regulation of FUT3 and FUT5 by shRNA alters Lewis antigens expression and reduces the adhesion capacities of gastric cancer cells. Biochim Biophys Acta. 2011;1810:1141–1149.

113. Boura JS, Santos FD, Gimble JM, et al. Direct head-to-head comparison of cationic liposome-mediated gene delivery to mesenchymal stem/stromal cells of different human sources: a comprehensive study. Hum Gene Ther Methods. 2013;24:38–48.

114. Madeira C, Mendes RD, Ribeiro SC, et al. Nonviral gene delivery to mesenchymal stem cells using cationic liposomes for gene and cell therapy. J Biomed Biotechnol. 2010;2010:735349.

115. Lim JY, Park SH, Jeong CH, et al. Microporation is a valuable transfection method for efficient gene delivery into human umbilical cord blood-derived mesenchymal stem cells. BMC Biotechnol. 2010;10:38.

116. Madeira C, Ribeiro SC, Pinheiro IS, et al. Gene delivery to human bone marrow mesenchymal stem cells by microporation. J Biotechnol. 2011;151:130–136.

117. Deyle DR, Li Y, Olson EM, Russell DW. Nonintegrating foamy virus vectors. J Virol. 2010;84:9341–9349.

118. Wiktorowicz T, Peters K, Armbruster N, Steinert AF, Rethwilm A. Generation of an improved foamy virus vector by dissection of cis-acting sequences. J Gen Virol. 2009;90:481–487.

119. Dwyer RM, Khan S, Barry FP, O’Brien T, Kerin MJ. Advances in mesenchymal stem cell-mediated gene therapy for cancer. Stem Cell Res Ther. 2010;1:25.

120. Shakhbazov AV, Kosmacheva SM, Kartel NA, Potapnev MP. Gene therapy based on human mesenchymal stem cells: strategies and methods. Tsitol Genet. 2010;44:76–82.

121. Kaufmann M, Blaser C, Takashima S, Schwartz-Albiez R, Tsuji S, Pircher H. Identification of an alpha2,6-sialyltransferase induced early after lymphocyte activation. Int Immunol. 1999;11:731–738.

122. Damle NK, Klussman K, Dietsch MT, Mohagheghpour N, Aruffo A. GMP-140 (P-selectin/CD62) binds to chronically stimulated but not resting CD4+ T lymphocytes and regulates their production of proinflammatory cytokines. Eur J Immunol. 1992;22:1789–1793.

123. Piller F, Piller V, Fox RI, Fukuda M. Human T-lymphocyte activation is associated with changes in O-glycan biosynthesis. J Biol Chem. 1988;263:15146–15150.

124. Picker LJ, Treer JR, Ferguson-Darnell B, Collins PA, Bergstresser PR, Terstappen LW. Control of lymphocyte recirculation in man II: Differential regulation of the cutaneous lymphocyte-associated antigen, a tissue-selective homing receptor for skin-homing T cells. J Immunol. 1993;150:1122–1136.

125. Carlow DA, Corbel SY, Williams MJ, Ziltener HJ. IL-2, -4, and -15 differentially regulate O-glycan branching and P-selectin ligand formation in activated CD8 T cells. J Immunol. 2001;167:6841–6848.

126. Lim YC, Xie H, Come CE, et al. IL-12, STAT4-dependent up-regulation of CD4+ T cell core 2 beta-1,6-N-acetylglucosaminyltransferase, an enzyme essential for biosynthesis of P-selectin ligands. J Immunol. 2001;167:4476–4484.

127. Blander JM, Visintin I, Janeway Jr CA, Medzhitov R. Alpha(1,3)-fucosyltransferase VII and alpha(2,3)-sialyltransferase IV are up-regulated in activated CD4 T cells and maintained after their differentiation into Th1 and migration into inflammatory sites. J Immunol. 1999;163:3746–3752.

128. Nakayama F, Teraki Y, Kudo T, et al. Expression of cutaneous lymphocyte-associated antigen regulated by a set of glycosyltransferases in human T cells: involvement of alpha1, 3-fucosyltransferase VII and beta1,4-galactosyltransferase I. J Invest Dermatol. 2000;115:299–306.

129. Wagers AJ, Waters CM, Stoolman LM, Kansas GS. Interleukin 12 and interleukin 4 control T cell adhesion to endothelial selectins through opposite effects on alpha1, 3-fucosyltransferase VII gene expression. J Exp Med. 1998;188:2225–2231.

130. Lim YC, Henault L, Wagers AJ, Kansas GS, Luscinskas FW, Lichtman AH. Expression of functional selectin ligands on Th cells is differentially regulated by IL-12 and IL-4. J Immunol. 1999;162:3193–3201.

131. van Wely CA, Blanchard AD, Britten CJ. Differential expression of alpha3 fucosyltransferases in Th1 and Th2 cells correlates with their ability to bind P-selectin. Biochem Biophys Res Commun. 1998;247:307–311.

132. Underhill GH, Zisoulis DG, Kolli KP, Ellies LG, Marth JD, Kansas GS. A crucial role for T-bet in selectin ligand expression in T helper 1 (Th1) cells. Blood. 2005;106:3867–3873.

133. Carlow DA, Williams MJ, Ziltener HJ. P-selectin ligand in activated CD8 T cells proceeds in absence of IL2 and IL12Rp40 in vivo. J Immunol. 2005;174.

134. Yamanaka KI, Dimitroff CJ, Fuhlbrigge RC, et al. Vitamins A and D are potent inhibitors of cutaneous lymphocyte-associated antigen expression. J Allergy Clin Immun. 2008;121:148–157.

135. Dagia NM, Gadhoum SZ, Knoblauch CA, et al. G-CSF induces E-selectin ligand expression on human myeloid cells. Nat Med. 2006;12:1185–1190.

136. Hoggatt J, Singh P, Sampath J, Pelus LM. Prostaglandin E-2 enhances hematopoietic stem cell homing, survival, and proliferation. Blood. 2009;113:5444–5455.

137. North TE, Goessling W, Walkley CR, et al. Prostaglandin E2 regulates vertebrate haematopoietic stem cell homeostasis. Nature. 2007;447:1007–1011.

138. Goessling W, Allen RS, Guan X, et al. Prostaglandin E2 enhances human cord blood stem cell xenotransplants and shows long-term safety in preclinical nonhuman primate transplant models. Cell Stem Cell. 2011;8:445–458.

139. Berberoglu U, Yildirim E, Celen O. Serum levels of tumor necrosis factor alpha correlate with response to neoadjuvant chemotherapy in locally advanced breast cancer. Int J Biol Markers. 2004;19:130–134.

140. Geng Y, Chandrasekaran S, Hsu JW, Gidwani M, Hughes AD, King MR. Phenotypic switch in blood: effects of pro-inflammatory cytokines on breast cancer cell aggregation and adhesion. PLoS ONE. 2013;8:e54959.

141. Bates RC, Mercurio AM. Tumor necrosis factor-alpha stimulates the epithelial-to-mesenchymal transition of human colonic organoids. Mol Biol Cell. 2003;14:1790–1800.

142. Lee SO, Lou W, Hou M, de Miguel F, Gerber L, Gao AC. Interleukin-6 promotes androgen-independent growth in LNCaP human prostate cancer cells. Clin Cancer Res. 2003;9:370–376.

143. Muller A, Homey B, Soto H, et al. Involvement of chemokine receptors in breast cancer metastasis. Nature. 2001;410:50–56.

144. Xia L, McDaniel JM, Yago T, Doeden A, McEver RP. Surface fucosylation of human cord blood cells augments binding to P-selectin and E-selectin and enhances engraftment in bone marrow. Blood. 2004;104:3091–3096.

145. Robinson SN, Simmons PJ, Thomas MW, et al. Ex vivo fucosylation improves human cord blood engraftment in NOD-SCID IL-2Rgamma(null) mice. Exp Hematol. 2012;40:445–456.

146. Wan X, Sato H, Miyaji H, et al. Fucosyltransferase VII improves the function of selectin ligands on cord blood hematopoietic stem cells. Glycobiology. 2013;23:1184–1191.

147. Whiteheart SW, Passaniti A, Reichner JS, Holt GD, Haltiwanger RS, Hart GW. Glycosyltransferase probes. Methods Enzymol. 1989;179:82–95.

148. Palcic MM. Glycosyltransferases in glycobiology. Methods Enzymol. 1994;230:300–316.

149. Warnock D, Bai X, Autote K, et al. In vitro galactosylation of human IgG at 1 kg scale using recombinant galactosyltransferase. Biotechnol Bioeng. 2005;92:831–842.

150. Zheng K, Bantog C, Bayer R. The impact of glycosylation on monoclonal antibody conformation and stability. MAbs. 2011;3:568–576.

151. Anthony RM, Nimmerjahn F, Ashline DJ, Reinhold VN, Paulson JC, Ravetch JV. Recapitulation of IVIG anti-inflammatory activity with a recombinant IgG Fc. Science. 2008;320:373–376.

152. Sackstein R. Directing stem cell trafficking via Gps. Method Enzymol. 2010;479:93–105.

153. Sackstein R. Engineering cellular trafficking via glycosyltransferase-programmed stereosubstitution. Ann N Y Acad Sci. 2012;1253:193–200.

154. Sackstein R. Glycoengineering of HCELL, the human bone marrow homing receptor: sweetly programming cell migration. Ann Biomed Eng. 2012;40:766–776.

155. Sackstein R, Fuhlbrigge R. The blot rolling assay: a method for identifying adhesion molecules mediating binding under shear conditions. Methods Mol Biol (Clifton, NJ). 2006;341:217–226.

156. Lairson LL, Henrissat B, Davies GJ, Withers SG. Glycosyltransferases: structures, functions, and mechanisms. Annu Rev Biochem. 2008;77:521–555.

157. Gallet PF, Vaujour H, Petit JM, et al. Heterologous expression of an engineered truncated form of human Lewis fucosyltransferase (Fuc-TIII) by the methylotrophic yeast Pichia pastoris. Glycobiology. 1998;8:919–925.

158. Stacke C, Ziegelmuller P, Hahn U. Comparison of expression systems for human fucosyltransferase IX. Eur J Cell Biol. 2010;89:35–38.

159. Auge C, Malleron A, Tahrat H, et al. Outstanding stability of immobilized recombinant alpha(1 -> 3/4)-fucosyltransferases exploited in the synthesis of Lewis a and Lewis x trisaccharides. Chem Commun 2000;2017–2018.

160. Baboval T, Koul O, Smith FI. N-glycosylation site occupancy of rat alpha-1,3-fucosyltransferase IV and the effect of glycosylation on enzymatic activity. Biochim Biophys Acta. 2000;1475:383–389.

161. Morais VA, Costa MT, Costa J. N-glycosylation of recombinant human fucosyltransferase III is required for its in vivo folding in mammalian and insect cells. Biochim Biophys Acta. 1619;133–138:2003.

162. Li P, Anumanthan A, Gao XG, et al. Expression of recombinant proteins in Pichia pastoris. Appl Biochem Biotechnol. 2007;142:105–124.

163. Kato T, Kajikawa M, Maenaka K, Park EY. Silkworm expression system as a platform technology in life science. Appl Microbiol Biotechnol. 2010;85:459–470.

164. Sacchetti B, Funari A, Michienzi S, et al. Self-renewing osteoprogenitors in bone marrow sinusoids can organize a hematopoietic microenvironment. Cell. 2007;131:324–336.

165. Mendez-Ferrer S, Michurina TV, Ferraro F, et al. Mesenchymal and haematopoietic stem cells form a unique bone marrow niche. Nature. 2010;466:829–834.

166. Ding L, Saunders TL, Enikolopov G, Morrison SJ. Endothelial and perivascular cells maintain haematopoietic stem cells. Nature. 2012;481:457–462.

167. Calvi LM, Adams GB, Weibrecht KW, et al. Osteoblastic cells regulate the haematopoietic stem cell niche. Nature. 2003;425:841–846.

168. Raaijmakers MH, Mukherjee S, Guo S, et al. Bone progenitor dysfunction induces myelodysplasia and secondary leukaemia. Nature. 2010;464:852–857.

169. Visnjic D, Kalajzic Z, Rowe DW, Katavic V, Lorenzo J, Aguila HL. Hematopoiesis is severely altered in mice with an induced osteoblast deficiency. Blood. 2004;103:3258–3264.

170. Butler JM, Nolan DJ, Vertes EL, et al. Endothelial cells are essential for the self-renewal and repopulation of Notch-dependent hematopoietic stem cells. Cell Stem Cell. 2010;6:251–264.

171. Yamazaki S, Ema H, Karlsson G, et al. Nonmyelinating Schwann cells maintain hematopoietic stem cell hibernation in the bone marrow niche. Cell. 2011;147:1146–1158.

172. Trumpp A, Essers M, Wilson A. Awakening dormant haematopoietic stem cells. Nat Rev Immunol. 2010;10:201–209.

173. Wilson A, Laurenti E, Oser G, et al. Hematopoietic stem cells reversibly switch from dormancy to self-renewal during homeostasis and repair. Cell. 2008;135:1118–1129.

174. Kiel MJ, Morrison SJ. Uncertainty in the niches that maintain haematopoietic stem cells. Nat Rev Immunol. 2008;8:290–301.

175. Xie Y, Yin T, Wiegraebe W, et al. Detection of functional haematopoietic stem cell niche using real-time imaging. Nature. 2009;457:97–101.

176. Levesque JP, Winkler IG, Hendy J, et al. Hematopoietic progenitor cell mobilization results in hypoxia with increased hypoxia-inducible transcription factor-1 alpha and vascular endothelial growth factor A in bone marrow. Stem Cells. 2007;25:1954–1965.

177. Parmar K, Mauch P, Vergilio JA, Sackstein R, Down JD. Distribution of hematopoietic stem cells in the bone marrow according to regional hypoxia. Proc Natl Acad Sci USA. 2007;104:5431–5436.

178. McIntyre TM, Prescott SM, Weyrich AS, Zimmerman GA. Cell–cell interactions: leukocyte-endothelial interactions. Curr Opin Hematol. 2003;10:150–158.

179. Okeley NM, Alley SC, Anderson ME, et al. Development of orally active inhibitors of protein and cellular fucosylation. Proc Natl Acad Sci USA. 2013;110:5404–5409.

180. Geng Y, Yeh K, Takatani T, King MR. Three to tango: MUC1 as a ligand for both E-selectin and ICAM-1 in the breast cancer metastatic cascade. Front Oncol. 2012;2:76.

181. Zen K, Liu DQ, Guo YL, et al. CD44v4 is a major E-selectin ligand that mediates breast cancer cell transendothelial migration. PLoS ONE. 2008;3:e1826.

182. Dimitroff CJ, Descheny L, Trujillo N, et al. Identification of leukocyte E-selectin ligands, P-selectin glycoprotein ligand-1 and E-selectin ligand-1, on human metastatic prostate tumor cells. Cancer Res. 2005;65:5750–5760.

183. Dimitroff CJ, Lechpammer M, Long-Woodward D, Kutok JL. Rolling of human bone-metastatic prostate tumor cells on human bone marrow endothelium under shear flow is mediated by E-selectin. Cancer Res. 2004;64:5261–5269.

184. Burdick MM, Bochner BS, Collins BE, Schnaar RL, Konstantopoulos K. Glycolipids support E-selectin-specific strong cell tethering under flow. Biochem Biophys Res Commun. 2001;284:42–49.

185. Burdick MM, Chu JT, Godar S, Sackstein R. HCELL is the major E- and L-selectin ligand expressed on LS174T colon carcinoma cells. J Biol Chem. 2006;281:13899–13905.

186. Gout S, Morin C, Houle F, Huot J. Death receptor-3, a new E-Selectin counter-receptor that confers migration and survival advantages to colon carcinoma cells by triggering p38 and ERK MAPK activation. Cancer Res. 2006;66:9117–9124.

187. Tremblay PL, Auger FA, Huot J. Regulation of transendothelial migration of colon cancer cells by E-selectin-mediated activation of p38 and ERK MAP kinases. Oncogene. 2006;25:6563–6573.

188. Martin-Satue M, de Castellarnau C, Blanco J. Overexpression of alpha(1,3)-fucosyltransferase VII is sufficient for the acquisition of lung colonization phenotype in human lung adenocarcinoma HAL-24Luc cells. Br J Cancer. 1999;80:1169–1174.

189. Martin-Satue M, Marrugat R, Cancelas JA, Blanco J. Enhanced expression of alpha(1,3)-fucosyltransferase genes correlates with E-selectin-mediated adhesion and metastatic potential of human lung adenocarcinoma cells. Cancer Res. 1998;58:1544–1550.

190. Iwai K, Ishikura H, Kaji M, et al. Importance of E-selectin (ELAM-1) and sialyl Lewis(a) in the adhesion of pancreatic carcinoma cells to activated endothelium. Int J Cancer. 1993;54:972–977.

191. Brodt P, Fallavollita L, Bresalier RS, Meterissian S, Norton CR, Wolitzky BA. Liver endothelial E-selectin mediates carcinoma cell adhesion and promotes liver metastasis. Int J Cancer. 1997;71:612–619.

192. Khatib AM, Kontogiannea M, Fallavollita L, Jamison B, Meterissian S, Brodt P. Rapid induction of cytokine and E-selectin expression in the liver in response to metastatic tumor cells. Cancer Res. 1999;59:1356–1361.

193. Biancone L, Araki M, Araki K, Vassalli P, Stamenkovic I. Redirection of tumor metastasis by expression of E-selectin in vivo. J Exp Med. 1996;183:581–587.

194. Mannori G, Santoro D, Carter L, Corless C, Nelson RM, Bevilacqua MP. Inhibition of colon carcinoma cell lung colony formation by a soluble form of E-selectin. Am J Pathol. 1997;151:233–243.

195. Borsig L, Wong R, Feramisco J, Nadeau DR, Varki NM, Varki A. Heparin and cancer revisited: mechanistic connections involving platelets, P-selectin, carcinoma mucins, and tumor metastasis. Proc Natl Acad Sci USA. 2001;98:3352–3357.

196. Borsig L, Wong R, Hynes RO, Varki NM, Varki A. Synergistic effects of L- and P-selectin in facilitating tumor metastasis can involve non-mucin ligands and implicate leukocytes as enhancers of metastasis. Proc Natl Acad Sci USA. 2002;99:2193–2198.

197. Nakamura S, Ohtani H, Watanabe Y, et al. In situ expression of the cell adhesion molecules in inflammatory bowel disease Evidence of immunologic activation of vascular endothelial cells. Lab Invest. 1993;69:77–85.

198. Kakiuchi Y, Tsuji S, Tsujii M, et al. Cyclooxygenase-2 activity altered the cell-surface carbohydrate antigens on colon cancer cells and enhanced liver metastasis. Cancer Res. 2002;62:1567–1572.

199. Kim YJ, Borsig L, Han HL, Varki NM, Varki A. Distinct selectin ligands on colon carcinoma mucins can mediate pathological interactions among platelets, leukocytes, and endothelium. Am J Pathol. 1999;155:461–472.

200. Nieswandt B, Hafner M, Echtenacher B, Mannel DN. Lysis of tumor cells by natural killer cells in mice is impeded by platelets. Cancer Res. 1999;59:1295–1300.

201. Barthel SR, Gavino JD, Wiese GK, Jaynes JM, Siddiqui J, Dimitroff CJ. Analysis of glycosyltransferase expression in metastatic prostate cancer cells capable of rolling activity on microvascular endothelial (E)-selectin. Glycobiology. 2008;18:806–817.

202. Fukuda M, Hiraoka N, Yeh JC. C-type lectins and sialyl Lewis X oligosaccharides Versatile roles in cell–cell interaction. J Cell Biol. 1999;147:467–470.

203. Zipin A, Israeli-Amit M, Meshel T, et al. Tumor–microenvironment interactions: the fucose-generating FX enzyme controls adhesive properties of colorectal cancer cells. Cancer Res. 2004;64:6571–6578.

204. Dube DH, Bertozzi CR. Glycans in cancer and inflammation—potential for therapeutics and diagnostics. Nat Rev Drug Discov. 2005;4:477–488.

205. Hoff SD, Matsushita Y, Ota DM, et al. Increased expression of sialyl-dimeric LeX antigen in liver metastases of human colorectal carcinoma. Cancer Res. 1989;49:6883–6888.

206. Nakamori S, Furukawa H, Hiratsuka M, et al. Expression of carbohydrate antigen sialyl Le(a): a new functional prognostic factor in gastric cancer. J Clin Oncol. 1997;15:816–825.

207. Nakamori S, Kameyama M, Imaoka S, et al. Involvement of carbohydrate antigen sialyl Lewis(x) in colorectal cancer metastasis. Dis Colon Rectum. 1997;40:420–431.

208. Renkonen R, Mattila P, Majuri ML, et al. In vitro experimental studies of sialyl Lewis x and sialyl Lewis a on endothelial and carcinoma cells: crucial glycans on selectin ligands. Glycoconj J. 1997;14:593–600.

209. Sakuma K, Aoki M, Kannagi R. Transcription factors c-Myc and CDX2 mediate E-selectin ligand expression in colon cancer cells undergoing EGF/bFGF-induced epithelial–mesenchymal transition. Proc Natl Acad Sci USA. 2012;109:7776–7781.

210. Greenburg G, Hay ED. Epithelia suspended in collagen gels can lose polarity and express characteristics of migrating mesenchymal cells. J Cell Biol. 1982;95:333–339.

211. Greenburg G, Hay ED. Cytodifferentiation and tissue phenotype change during transformation of embryonic lens epithelium to mesenchyme-like cells in vitro. Dev Biol. 1986;115:363–379.

212. Greenburg G, Hay ED. Cytoskeleton and thyroglobulin expression change during transformation of thyroid epithelium to mesenchyme-like cells. Development. 1988;102:605–622.

213. Kalluri R. EMT: when epithelial cells decide to become mesenchymal-like cells. J Clin Invest. 2009;119:1417–1419.

214. Kalluri R, Weinberg RA. The basics of epithelial–mesenchymal transition. J Clin Invest. 2009;119:1420–1428.

215. Aktas B, Tewes M, Fehm T, Hauch S, Kimmig R, Kasimir-Bauer S. Stem cell and epithelial–mesenchymal transition markers are frequently overexpressed in circulating tumor cells of metastatic breast cancer patients. Breast Cancer Res. 2009;11:R46.

216. Blick T, Hugo H, Widodo E, et al. Epithelial mesenchymal transition traits in human breast cancer cell lines parallel the CD44(hi/)CD24 (lo/−) stem cell phenotype in human breast cancer. J Mammary Gland Biol Neoplasia. 2010;15:235–252.

217. Brabletz T, Jung A, Spaderna S, Hlubek F, Kirchner T. Opinion: migrating cancer stem cells—an integrated concept of malignant tumour progression. Nat Rev Cancer. 2005;5:744–749.

218. Kong D, Banerjee S, Ahmad A, et al. Epithelial to mesenchymal transition is mechanistically linked with stem cell signatures in prostate cancer cells. PLoS ONE. 2010;5:e12445.

219. Kong D, Wang Z, Sarkar SH, et al. Platelet-derived growth factor-D overexpression contributes to epithelial–mesenchymal transition of PC3 prostate cancer cells. Stem Cells. 2008;26:1425–1435.

220. Morel AP, Lievre M, Thomas C, Hinkal G, Ansieau S, Puisieux A. Generation of breast cancer stem cells through epithelial–mesenchymal transition. PLoS ONE. 2008;3:e2888.

221. Santisteban M, Reiman JM, Asiedu MK, et al. Immune-induced epithelial to mesenchymal transition in vivo generates breast cancer stem cells. Cancer Res. 2009;69:2887–2895.

222. Thomson S, Buck E, Petti F, et al. Epithelial to mesenchymal transition is a determinant of sensitivity of non-small-cell lung carcinoma cell lines and xenografts to epidermal growth factor receptor inhibition. Cancer Res. 2005;65:9455–9462.

223. Zhau HE, Odero-Marah V, Lue HW, et al. Epithelial to mesenchymal transition (EMT) in human prostate cancer: lessons learned from ARCaP model. Clin Exp Metastasis. 2008;25:601–610.

224. Pinho SS, Reis CA, Gartner F, Alpaugh ML. Molecular plasticity of E-cadherin and sialyl Lewis X expression, in two comparative models of mammary tumorigenesis. PLoS ONE. 2009;4.

225. Goetz DJ, Ding H, Atkinson WJ, et al. A human colon carcinoma cell line exhibits adhesive interactions with P-selectin under fluid flow via a PSGL-1-independent mechanism. Am J Pathol. 1996;149:1661–1673.

226. Hanley WD, Burdick MM, Konstantopoulos K, Sackstein R. CD44 on LS174T colon carcinoma cells possesses E-selectin ligand activity. Cancer Res. 2005;65:5812–5817.

227. Hanley WD, Napier SL, Burdick MM, Schnaar RL, Sackstein R, Konstantopoulos K. Variant isoforms of CD44 are P- and L-selectin ligands on colon carcinoma cells. FASEB J. 2006;20:337–339.

228. Napier SL, Healy ZR, Schnaar RL, Konstantopoulos K. Selectin ligand expression regulates the initial vascular interactions of colon carcinoma cells: the roles of CD44v and alternative sialofucosylated selectin ligands. J Biol Chem. 2007;282:3433–3441.

229. Thomas SN, Schnaar RL, Konstantopoulos K. Podocalyxin-like protein is an E-/L-selectin ligand on colon carcinoma cells: comparative biochemical properties of selectin ligands in host and tumor cells. Am J Physiol Cell Physiol. 2009;296:C505–C513.

230. Fernandez-Rodriguez J, Dwir O, Alon R, Hansson GC. Tumor cell MUC1 and CD43 are glycosylated differently with sialyl-Lewis a and x epitopes and show variable interactions with E-selectin under physiological flow conditions. Glycoconj J. 2001;18:925–930.

231. Zhang K, Baeckstrom D, Brevinge H, Hansson GC. Secreted MUC1 mucins lacking their cytoplasmic part and carrying sialyl-Lewis a and x epitopes from a tumor cell line and sera of colon carcinoma patients can inhibit HL-60 leukocyte adhesion to E-selectin-expressing endothelial cells. J Cell Biochem. 1996;60:538–549.

232. Alcaide P, King SL, Dimitroff CJ, Lim YC, Fuhlbrigge RC, Luscinskas FW. The 130-kDa glycoform of CD43 functions as an E-selectin ligand for activated Th1 cells in vitro and in delayed-type hypersensitivity reactions in vivo. J Invest Dermatol. 2007;127:1964–1972.

233. Matsumoto M, Atarashi K, Umemoto E, et al. CD43 functions as a ligand for E-Selectin on activated T cells. J Immunol. 2005;175:8042–8050.

234. Shirure VS, Reynolds NM, Burdick MM. Mac-2 binding protein is a novel e-selectin ligand expressed by breast cancer cells. PLoS ONE. 2012;7:e44529.

235. Sako D, Chang XJ, Barone KM, et al. Expression cloning of a functional glycoprotein ligand for P-selectin. Cell. 1993;75:1179–1186.

236. Veldman GM, Bean KM, Cumming DA, Eddy RL, Sait SN, Shows TB. Genomic organization and chromosomal localization of the gene encoding human P-selectin glycoprotein ligand. J Biol Chem. 1995;270:16470–16475.

237. Moore KL, Stults NL, Diaz S, et al. Identification of a specific glycoprotein ligand for P-selectin (CD62) on myeloid cells. J Cell Biol. 1992;118:445–456.

238. Li FG, Erickson HP, James JA, Moore KL, Cummings RD, McEver RP. Visualization of P-selectin glycoprotein ligand-1 as a highly extended molecule and mapping of protein epitopes for monoclonal antibodies. J Biol Chem. 1996;271:6342–6348.

239. Laszik Z, Jansen PJ, Cummings RD, Tedder TF, McEver RP, Moore KL. P-selectin glycoprotein ligand-1 is broadly expressed in cells of myeloid, lymphoid, and dendritic lineage and in some nonhematopoietic cells. Blood. 1996;88:3010–3021.

240. Moore TM, Khimenko P, Adkins WK, Miyasaka M, Taylor AE. Adhesion molecules contribute to ischemia and reperfusion-induced injury in the isolated rat lung. J Appl Physiol. 1995;78:2245–2252.

241. Frenette PS, Denis CV, Weiss L, et al. P-Selectin glycoprotein ligand 1 (PSGL-1) is expressed on platelets and can mediate platelet–endothelial interactions in vivo. J Exp Med. 2000;191:1413–1422.

242. Kappelmayer J, Kiss A, Karaszi E, Veszpremi A, Jako J, Kiss C. Identification of P-selectin glycoprotein ligand-1 as a useful marker in acute myeloid leukaemias. Br J Haematol. 2001;115:903–909.

243. Li F, Wilkins PP, Crawley S, Weinstein J, Cummings RD, McEver RP. Post-translational modifications of recombinant P-selectin glycoprotein ligand-1 required for binding to P- and E-selectin. J Biol Chem. 1996;271:3255–3264.

244. Dimitroff CJ, Lee JY, Schor KS, Sandmaier BM, Sackstein R. Differential L-selectin binding activities of human hematopoietic cell L-selectin ligands, HCELL and PSGL-1. J Biol Chem. 2001;276:47623–47631.

245. Brockhausen I. Mucin-type O-glycans in human colon and breast cancer: glycodynamics and functions. EMBO Rep. 2006;7:599–604.

246. Aigner S, Ramos CL, Hafezi-Moghadam A, et al. CD24 mediates rolling of breast carcinoma cells on P-selectin. FASEB J. 1998;12:1241–1251.

247. Myung JH, Gajjar KA, Pearson RM, Launiere CA, Eddington DT, Hong S. Direct measurements on CD24-mediated rolling of human breast cancer MCF-7 cells on E-selectin. Anal Chem. 2011;83:1078–1083.

248. Shimodaira K, Nakayama J, Nakamura N, Hasebe O, Katsuyama T, Fukuda M. Carcinoma-associated expression of core 2 beta-1,6-N-acetylglucosaminyltransferase gene in human colorectal cancer: role of O-glycans in tumor progression. Cancer Res. 1997;57:5201–5206.

249. St Hill CA, Farooqui M, Mitcheltree G, et al. The high affinity selectin glycan ligand C2-O-sLex and mRNA transcripts of the core 2 beta-1,6-N-acetylglucosaminyltransferase (C2GnT1) gene are highly expressed in human colorectal adenocarcinomas. BMC Cancer. 2009;9:79.

250. Ogawa J, Inoue H, Koide S. Expression of alpha-1,3-fucosyltransferase type IV and VII genes is related to poor prognosis in lung cancer. Cancer Res. 1996;56:325–329.

251. Hiller KM, Mayben JP, Bendt KM, et al. Transfection of alpha(1,3)fucosyltransferase antisense sequences impairs the proliferative and tumorigenic ability of human colon carcinoma cells. Mol Carcinog. 2000;27:280–288.

252. Weston BW, Hiller KM, Mayben JP, et al. Expression of human alpha(1,3)fucosyltransferase antisense sequences inhibits selectin-mediated adhesion and liver metastasis of colon carcinoma cells. Cancer Res. 1999;59:2127–2135.

253. Recchi MA, Harduin-Lepers A, Boilly-Marer Y, Verbert A, Delannoy P. Multiplex RT-PCR method for the analysis of the expression of human sialyltransferases: application to breast cancer cells. Glycoconj J. 1998;15:19–27.

254. Recchi MA, Hebbar M, Hornez L, Harduin-Lepers A, Peyrat JP, Delannoy P. Multiplex reverse transcription polymerase chain reaction assessment of sialyltransferase expression in human breast cancer. Cancer Res. 1998;58:4066–4070.

255. Bernacki RJ, Kim U. Concomitant elevations in serum sialytransferase activity and sialic acid content in rats with metastasizing mammary tumors. Science. 1977;195:577–580.

256. Varki A. Selectin ligands: will the real ones please stand up? J Clin Invest. 1997;100:S31–S35.

257. Munro JM. Endothelial–leukocyte adhesive interactions in inflammatory diseases. Eur Heart J. 1993;14(Suppl. KSuppl. K):72–77.

258. Perez P, Garcia-Acha I, Duran A. Effect of papulacandin B on the cell wall and growth of Geotrichum lactis. J Gen Microbiol. 1983;129:245–250.

259. Cabib E, Kang MS, Au-Young J. Chitin synthase from Saccharomyces cerevisiae. Methods Enzymol. 1987;138:643–649.

260. Lehrman MA. Biosynthesis of N-acetylglucosamine-P-P-dolichol, the committed step of asparagine-linked oligosaccharide assembly. Glycobiology. 1991;1:553–562.

261. Saul R, Chambers JP, Molyneux RJ, Elbein AD. Castanospermine, a tetrahydroxylated alkaloid that inhibits beta-glucosidase and beta-glucocerebrosidase. Arch Biochem Biophys. 1983;221:593–597.

262. Goss PE, Baptiste J, Fernandes B, Baker M, Dennis JW. A phase I study of swainsonine in patients with advanced malignancies. Cancer Res. 1994;54:1450–1457.

263. Bischoff J, Kornfeld R. The effect of 1-deoxymannojirimycin on rat liver alpha-mannosidases. Biochem Biophys Res Commun. 1984;125:324–331.

264. Platt FM, Neises GR, Dwek RA, Butters TD. N-butyldeoxynojirimycin is a novel inhibitor of glycolipid biosynthesis. J Biol Chem. 1994;269:8362–8365.

265. Schwientek T, Yeh JC, Levery SB, et al. Control of O-glycan branch formation Molecular cloning and characterization of a novel thymus-associated core 2 beta1, 6-N-acetylglucosaminyltransferase. J Biol Chem. 2000;275:11106–11113.

266. Hindsgaul O, Kaur KJ, Srivastava G, et al. Evaluation of deoxygenated oligosaccharide acceptor analogs as specific inhibitors of glycosyltransferases. J Biol Chem. 1991;266:17858–17862.

267. Gloster TM, Zandberg WF, Heinonen JE, Shen DL, Deng L, Vocadlo DJ. Hijacking a biosynthetic pathway yields a glycosyltransferase inhibitor within cells. Nat Chem Biol. 2011;7:174–181.

268. Grabenhorst E, Nimtz M, Costa J, Conradt HS. In vivo specificity of human alpha1,3/4-fucosyltransferases III–VII in the biosynthesis of LewisX and Sialyl LewisX motifs on complex-type N-glycans Coexpression studies from bhk-21 cells together with human beta-trace protein. J Biol Chem. 1998;273:30985–30994.

269. Zandberg WF, Kumarasamy J, Pinto BM, Vocadlo DJ. Metabolic inhibition of sialyl-Lewis X biosynthesis by 5-thiofucose remodels the cell surface and impairs selectin-mediated cell adhesion. J Biol Chem. 2012;287:40021–40030.

270. Rillahan CD, Antonopoulos A, Lefort CT, et al. Global metabolic inhibitors of sialyl- and fucosyltransferases remodel the glycome. Nat Chem Biol. 2012;8:661–668.

271. Izumi M, Wada K, Yuasa H, Hashimoto H. Synthesis of bisubstrate and donor analogues of sialyltransferase and their inhibitory activities. J Org Chem. 2005;70:8817–8824.

272. Burkart MD, Vincent SP, Duffels A, Murray BW, Ley SV, Wong CH. Chemo-enzymatic synthesis of fluorinated sugar nucleotide: useful mechanistic probes for glycosyltransferases. Bioorg Med Chem. 2000;8:1937–1946.

273. Van Dorst JA, Tikkanen JM, Krezdorn CH, et al. Exploring the substrate specificities of alpha-2,6- and alpha-2,3-sialyltransferases using synthetic acceptor analogues. Eur J Biochem. 1996;242:674–681.

274. Hinou H, Sun XL, Ito Y. Systematic syntheses and inhibitory activities of bisubstrate-type inhibitors of sialyltransferases. J Org Chem. 2003;68:5602–5613.

275. Kuan SF, Byrd JC, Basbaum C, Kim YS. Inhibition of mucin glycosylation by aryl-N-acetyl-alpha-galactosaminides in human colon cancer cells. J Biol Chem. 1989;264:19271–19277.

276. Zhuang D, Grey A, Harris-Brandts M, Higgins E, Kashem MA, Dennis JW. Characterization of O-linked oligosaccharide biosynthesis in cultured cells using paranitrophenyl alpha-D-GalNAc as an acceptor. Glycobiology. 1991;1:425–433.

277. Gouyer V, Leteurtre E, Delmotte P, et al. Differential effect of GalNAcalpha-O-bn on intracellular trafficking in enterocytic HT-29 and Caco-2 cells: correlation with the glycosyltransferase expression pattern. J Cell Sci. 2001;114:1455–1471.

278. Huet G, Hennebicq-Reig S, de Bolos C, et al. GalNAc-alpha-O-benzyl inhibits NeuAcalpha2-3 glycosylation and blocks the intracellular transport of apical glycoproteins and mucus in differentiated HT-29 cells. J Cell Biol. 1998;141:1311–1322.

279. Huet G, Kim I, de Bolos C, et al. Characterization of mucins and proteoglycans synthesized by a mucin-secreting HT-29 cell subpopulation. J Cell Sci. 1995;108(Pt 3Pt 3):1275–1285.

280. Neville DC, Field RA, Ferguson MA. Hydrophobic glycosides of N-acetylglucosamine can act as primers for polylactosamine synthesis and can affect glycolipid synthesis in vivo. Biochem J. 1995;307(Pt 3Pt 3):791–797.

281. Marathe DD, Buffone Jr A, Chandrasekaran EV, et al. Fluorinated per-acetylated GalNAc metabolically alters glycan structures on leukocyte PSGL-1 and reduces cell binding to selectins. Blood. 2010;115:1303–1312.

282. Barthel SR, Antonopoulos A, Cedeno-Laurent F, et al. Peracetylated 4-fluoro-glucosamine reduces the content and repertoire of N- and O-glycans without direct incorporation. J Biol Chem. 2011;286:21717–21731.

283. Brown JR, Fuster MM, Whisenant T, Esko JD. Expression patterns of alpha 2,3-sialyltransferases and alpha 1,3-fucosyltransferases determine the mode of sialyl Lewis X inhibition by disaccharide decoys. J Biol Chem. 2003;278:23352–23359.

284. Sarkar AK, Brown JR, Esko JD. Synthesis and glycan priming activity of acetylated disaccharides. Carbohydr Res. 2000;329:287–300.

285. Sarkar AK, Fritz TA, Taylor WH, Esko JD. Disaccharide uptake and priming in animal-cells—inhibition of sialyl-Lewis-X by acetylated Gal-beta-1-]4glcnac-beta-O-naphthalenemethanol. Proc Natl Acad Sci USA. 1995;92:3323–3327.

286. Sarkar AK, Rostand KS, Jain RK, Matta KL, Esko JD. Fucosylation of disaccharide precursors of sialyl LewisX inhibit selectin-mediated cell adhesion. J Biol Chem. 1997;272:25608–25616.

287. Fuster MM, Brown JR, Wang L, Esko JD. A disaccharide precursor of sialyl Lewis X inhibits metastatic potential of tumor cells. Cancer Res. 2003;63:2775–2781.

288. Brown JR, Yang F, Sinha A, et al. Deoxygenated disaccharide analogs as specific inhibitors of beta 1-4-galactosyltransferase 1 and selectin-mediated tumor metastasis. J Biol Chem. 2009;284:4952–4959.

289. Kerkela E, Hakkarainen T, Makela T, et al. Transient proteolytic modification of mesenchymal stromal cells increases lung clearance rate and targeting to injured tissue. Stem Cells Transl Med. 2013;2:510–520.

290. Nystedt J, Anderson H, Tikkanen J, et al. Cell surface structures influence lung clearance rate of systemically infused mesenchymal stromal cells. Stem Cells. 2013;31:317–326.

291. Perkey E, Miller RA, Garcia GG. Ex vivo enzymatic treatment of aged CD4 T cells restores cognate T cell helper function and enhances antibody production in mice. J Immunol. 2012;189:5582–5589.

292. Kumar R, Burns EA. Age-related decline in immunity: implications for vaccine responsiveness. Expert Rev Vaccines. 2008;7:467–479.

293. Haynes L, Eaton SM. The effect of age on the cognate function of CD4+ T cells. Immunol Rev. 2005;205:220–228.

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

You can't read the all page of ebook, please click here login for view all page.
Reset
3.144.100.237