Skip to main content
Access keys NCBI Homepage MyNCBI Homepage Main Content Main Navigation
Matrix Biol. Author manuscript; available in PMC 2015 Apr 1.
Published in final edited form as:
PMCID: PMC4039620
NIHMSID: NIHMS551045
PMID: 24361527

Heparan sulfate 3-O-sulfation: A rare modification in search of a function

Abstract

Many protein ligands bind to heparan sulfate, which results in their presentation, protection, oligomerization or conformational activation. Binding depends on the pattern of sulfation and arrangement of uronic acid epimers along the chains. Sulfation at the C3 position of glucosamine is a relatively rare, yet biologically significant modification, initially described as a key determinant for binding and activation of antithrombin and later for infection by Type I Herpes Simplex virus. In mammals, a family of seven heparan sulfate 3-O-sulfotransferases installs sulfate groups at this position and constitutes the largest group of sulfotransferases involved in heparan sulfate formation. However, to date very few proteins or biological systems have been described that are influenced by 3-O-sulfation. This review describes our current understanding of the prevalence and structure of 3-O-sulfation sites, expression and substrate specificity of the 3-O-sulfotransferase family and the emerging roles of 3-O-sulfation in biology.

Keywords: Heparan sulfate, sulfation, 3-O-sulfotransferases, antithrombin, growth factors

2. Introduction

Heparan sulfate is a type of sulfated glycosaminoglycan found covalently attached to a small set of extracellular matrix and plasma membrane proteoglycans. Its origin is ancient, having emerged during metazoan evolution, and its utility is evident, as only minor changes in composition have occurred over more than 500 million years of evolution (Medeiros et al., 2000). Many so-called heparin-binding proteins are known, many of which bind to heparan sulfate under physiological conditions and modulate cell division and differentiation, tissue morphogenesis and architecture, and organismal physiology (Bishop et al., 2007; Ori et al., 2008; Xu and Esko, 2014). Binding to heparan sulfate can have many effects on the protein ligand, ranging from simple presentation and/or stabilization to induction of conformational change, receptor-ligand interactions and protein oligomerization as a prelude to signaling. Thus, much interest exists in understanding the rules that guide the selective engagement of proteins with heparan sulfate chains.

Heparan sulfate is a linear polysaccharide composed of alternating glucosamine and uronic acid residues (Fig. 1) (Esko and Selleck, 2002). During polymerization of the chains, several classes of sulfotransferases install sulfate groups at various positions, including C2 of the uronic acid and N-, C6 and C3 of the glucosamine units. These reactions occur substoichiometrically in segments of variable size along the chain resulting in highly heterogeneous products with variable sulfation. The addition of the 3-O-sulfate group to glucosamine units is a relatively rare modification, present in only a limited number of chains or absent entirely (de Agostini et al., 2008; Marcum et al., 1986a; Pejler et al., 1987a). It is also one of the last modifications in biosynthesis (Zhang et al., 2001a; Zhang et al., 2001b), meaning that the substrates for the 3-O-sulfotransferases (Hs3sts) are sulfated oligosaccharides that have already been modified at other positions by the N-, 2-O- and 6-O-sulfotransferases and by the C5 epimerase (Kusche et al., 1988).

An external file that holds a picture, illustration, etc.
Object name is nihms551045f1.jpg
Structure of the antithrombin-binding pentasaccharide found in heparin

The individual residues of the pentasaccharide are numbered relative to the N-sulfo-glucosamine-3-sulfate residue at position 0. The central 3-O-sulfate group generated by Hs3st-1 (shown in red) and the 6-O-sulfate group at residue −2 (shown in blue) account for the majority of the binding energy of antithrombin to heparin. Note that the 6-O-sulfate group on the 3-O-sulfated glucosamine residue is dispensable and that residue −2 can be N-sulfated.

Binding to heparan sulfate depends on complementarity between positively charged amino acids in the protein ligands and the negatively charged sulfate groups and uronic acid epimers in the polysaccharide chain (Lindahl and Li, 2009; Xu and Esko, 2014). In general, two types of interactions occur, those that depend on overall sulfation and those based on specific types or arrangements of sulfated residues and uronic acid epimers. The former group consists of ligands that can accommodate different arrangements of sulfated sugars, presumably because of conformational flexibility in the binding site that can fit different orientations of sulfate and carboxyl groups in the chains. The latter group consists of ligands that rely on specific subsets of sulfate groups or the spacing of sulfated domains, and ligands that depend on a specific sequence of sulfated sugar residues (some containing a 3-O-sulfated glucosamine) for optimal binding. The paucity of known proteins that are influenced by 3-O-sulfation is rather surprising given the large family of Hs3sts involved in heparan sulfate biosynthesis. This review focuses on 3-O-sulfation, in particular the origin, substrate specificity, protein structure, and expression pattern of the Hs3st family, and the small family of proteins described to date whose activity is influenced by 3-O-sulfation.

3. Prevalence of 3-O-sulfation

The prevalence of 3-O-sulfation in natural heparan sulfates is largely unknown. Reasons for this lack of information include difficulty in obtaining large quantities of heparan sulfate for structural analyses and lack of technology to quantitate the 3-O-sulfate group. Historically, the 3-O-sulfate group was discovered by while searching for enzymes that remove sulfate groups from heparin. An enzyme in the urine was discovered that released the 3-O-sulfate group from synthetic radioactive N-sulfoglucosamine-3-sulfate (Leder, 1980) and from the non-reducing end of heparin fragments (Lindahl et al., 1980). We now know this enzyme as arylsulfatase G and its deficiency results in a lysosomal storage disease in mice in which heparan sulfate accumulates (Kowalewski et al., 2012). Since Leder’s initial study, chemical analyses, NMR and mass spectrometry have confirmed the presence of 3-O-sulfate groups in heparin (Lindahl et al., 1980; Meyer et al., 1981; Yamada et al., 1993) and in some preparations of heparan sulfate (Edge and Spiro, 1990; Marcum et al., 1986a; Pejler et al., 1987a). These latter studies indicate that the prevalence of 3-O-sulfation varies based on the source of heparan sulfate. Heparan sulfate from endothelial cells contains about one 3-O-sulfate group per 100 disaccharides (Marcum et al., 1986a). Five to ten percent of disaccharides from heparan sulfate derived from Reichert’s basement membrane contains 3-O-sulfate (Pejler et al., 1987a), whereas basement membrane heparan sulfate from Engelbreth-Holm-Swarm mouse tumor does not contain any 3-O-sulfate (Pejler et al., 1987a). Heparan sulfate from follicular fluid contains about 6 percent 3-O-sulfated glucosamine units (de Agostini et al., 2008). Certain animal species produce copious amounts of 3-O-sulfated heparin. For example, the clam, Anomalocardia brasiliana, produces heparin with one 3-O-sulfate for every 5 disaccharides (Pejler et al., 1987b). Many tissues express one or more isozyme of Hs3st, suggesting the presence of 3-O-sulfate groups, but very few of these tissues have been analyzed structurally to date.

4. Sites of modification by Hs3sts

The Hs3sts represent the largest gene family among all heparan sulfate sulfotransferases (Liu and Pedersen, 2007). Vertebrates generally have seven isozymes of Hs3st divided into two subgroups based on the homology of the sulfotransferase domain (Liu and Pedersen, 2007). Zebrafish has one additional Hs3st. Hs3st−2, −3a, −3b, −4 and −6 form one subgroup, sharing greater than 80 percent sequence identity in the sulfotransferase domain (Lawrence et al., 2007). This group is often referred to as “gD-type” Hs3sts because all members of the subfamily can generate binding sites for glycoprotein gD of Type I Herpes simplex virus (O'Donnell C et al., 2006; Shukla et al., 1999; Tiwari et al., 2005; Xia et al., 2002; Xu et al., 2005). Hs3st-1 and −5 form the other subgroup, sharing 71 percent identity in the sulfotransferase domain (Xia et al., 2002). These two sulfotransferases have in common the capacity to generate a binding site for antithrombin and thus are designated “AT-type” sulfotransferases. In vertebrates, the AT-type and gD-type subgroups share about 60 percent amino acid identity in the sulfotransferase domain (Lawrence et al., 2007).

Based on the large number of Hs3sts and the observation that they probably act after other sulfotransferases, one might assume that they show selectivity for substrates. Indeed, Hs3st-1 preferentially modifies sites in which a glucuronic acid devoid of 2-O-sulfate resides to the non-reducing side of the target glucosamine unit (Position −1, Fig 1) (Table 1) (Liu et al., 1996; Mochizuki et al., 2008; Shworak et al., 1997; Zhang et al., 1999). The enzyme will tolerate iduronic acid in this position, but 2-O-sulfation specifically prevents its action (Nguyen et al., 2012; Zhang et al., 2001b). In contrast, Hs3st−2, −3, −4 and −6 (gD-types) preferentially modify sites in which position −1 is 2-sulfo-iduronic acid (Lawrence et al., 2007; Liu et al., 1999a; Meissen et al., 2009; Mochizuki et al., 2008; Wu et al., 2004; Xu et al., 2005). Hs3st-5 modifies sites irrespective of 2-O-sulfation and consequently can produce both AT and gD type modifications (Chen et al., 2003; Chen and Liu, 2005; Duncan et al., 2004; Liu et al., 1999a; Mochizuki et al., 2003; Mochizuki et al., 2008; Xia et al., 2002; Xu et al., 2008).

Table 1

Oligosaccharides containing 3-O-sulfated glucosamine units

OligosaccharideEnzymeReferences
Disaccharides
GlcA-aMan3S (G0M3)Hs3st-
1,−5
(Chen et al., 2003; Liu et al., 1996; Shworak et al., 1997)
GlcA-aMan3S6S (G0M9)Hs3st-
1,−5
(Chen et al., 2003; Chen and Liu, 2005; Duncan et al., 2004;
Liu et al., 1996; Shworak et al., 1997; Xu et al., 2008; Zhang et al., 2001b)
IdoA-aMan3S (I0M3)Hs3st-
1,−5
(Chen et al., 2003; Zhang et al., 2001b)
IdoA-aMan3S6S (I0M9)Hs3st-
1,−5
(Chen et al., 2003; Xu et al., 2008; Zhang et al., 2001b)
GlcA2S-aMan3S (G2M3)Hs3st-2(Liu et al., 1999a)
IdoA2S-aMan3S (I2M3)Hs3st-
2,−3a,−
3b,−5,−
6
(Chen et al., 2003; Chen and Liu, 2005; Duncan et al., 2004; Liu et al., 1999a; Xia et al., 2002; Xu et al., 2008; Xu et al., 2005)
IdoA2S-aMan3S6S (I2M9)Hs3st-
5,−6
(Chen et al., 2003; Chen and Liu, 2005; Duncan et al., 2004; Xia et al., 2002; Xu et al., 2008; Xu et al., 2005)
Δ4,5UA-GlcNS3S6S (D0S9)Hs3st-5(Mochizuki et al., 2003; Mochizuki et al., 2008)
Δ4,5UA2S-GlcNS3S (D2S3)Hs3st-
2,−3,−
4,−5
(Lawrence et al., 2007; Mochizuki et al., 2003; Mochizuki et al., 2008; Wu et al., 2004)
Δ4,5UA2S-GlcNS3S6S (D2S9)Hs3st-
2,−3,−
4,−5
(Lawrence et al., 2007; Mochizuki et al., 2003; Mochizuki et al., 2008; Wu et al., 2004)
Tetrasaccharides
IdoA-GlcNAc6S-GlcA-aMan3S (I0A6-G0M3)AT-
type
(Bienkowski and Conrad, 1985)
IdoA-GlcNAc6S-GlcA-aMan3S6S (I0A6-G0M9)AT-
type
(Bienkowski and Conrad, 1985)
GlcA-GlcNS3S6S-IdoA-GlcNS6S (G0S9-I0S6)Hs3st-
3a
(Nguyen et al., 2012)
IdoA-GlcNS3S6S-IdoA-GlcNS6S (I0S9-I0S6)Hs3st-
3a
(Nguyen et al., 2012)
Δ4,5UA-GlcNAc6S-GlcA-GlcNS3S (D0A6-G0S3)Hs3st-1(Yamada et al., 1993; Zhang et al., 1999)
Δ4,5UA-GlcNAc6S-GlcA-GlcNS3S6S (D0A6-G0S9)Hs3st-1(Xu and Esko, 2014; Zhang et al., 1999)
Δ4,5UA-GlcNS6S-GlcA-GlcNS3S6S(D0S6-G0S9)AT-
type
(Yamada et al., 1993)
Δ4,5UA2S-GlcNS-IdoA2S-GlcN3S(D2S0-I2H3)Hs3st-
3a
(Liu et al., 1999b)
Δ4,5UA2S-GlcNS-IdoA2S-GlcN3S6S(D2S0-I2H9)Hs3st-
3a
(Liu et al., 1999b)
Pentasaccharides
GlcNAc6S-GlcA-GlcNS3S6S-IdoA2S-GlcNS (A6-
G0S9-I2S0)
Hs3st-1(Kuberan et al., 2003)
Hexasaccharides
IdoA-GlcNS6S-GlcA-GlcNS3S6S-IdoA2S-
GlcNS6S (I0S6-G0S9-I2S6)
AT-
type
(Linhardt et al., 1982)
GlcA-GlcNS3S6S-IdoA-GlcNS3S6S-GlcA-
GlcNS6S (G0S9-I0S9-G0S6)
Hst3st-
3a
(Nguyen et al., 2012)
Δ4,5UA2S-GlcNS6S-IdoA-GlcNAc6S-GlcA-
GlcNS3S (D2S6-I0A6-G0S3)
AT-
type
(Tsuda et al., 1996)
Δ4,5UA2S-GlcNAc6S-GlcA-GlcNS3S6S-IdoA-
GlcNS6S (D2A6-G0S9-I0S6)
AT-
type
(Desai et al., 1993)
Δ4,5UA2S-GlcNS6S-IdoA-GlcNAc6S-GlcA-
GlcNS3S6S (D2S6-I0A6-G0S9)
AT-
type
(Tsuda et al., 1996; Xu and Esko, 2014)
Δ4,5UA-GlcNAc-GlcA-GlcNS-IdoA2S-GlcNS3S
(D0A0-G0S0-I2S3)
Hs3st-4(Wu et al., 2004)
Δ4,5UA-GlcNAc-GlcA-GlcNS-IdoA2S-GlcNS3S6S
(D0A0-G0S0-I2S9)
Hs3st-4(Wu et al., 2004)
Heptasaccharides
GlcNAc6S-GlcA-GlcNS3S6S-IdoA2S-GlcNS6S-
GlcA-aMan (A6-G0S9-I2S6-G0M0)
Hs3st-1(Pempe et al., 2012; Xu et al., 2011)
GlcNS6S-GlcA-GlcNS3S6S-IdoA2S-GlcNS6S-
GlcA-aMan (S6-G0S9-I2S6-G0M0)
Hs3st-1(Xu et al., 2011)
Octasaccharides
IdoA2S-GlcNS6S-IdoA-GlcNAc6S-GlcA-
GlcNS3S6S-IdoA2S-aMan6S (I2S6-I0A6-G0S9-
I2M6)
AT-
type
(Casu et al., 1981)
IdoA-GlcNS6S-GlcA-GlcNS3S6S-IdoA2S-
GlcNS6S-IdoA2S-aMan (I0S6-G0S9-I2S6-I2M0)
AT-
type
(Linhardt et al., 1982)
IdoA-GlcNAc6S-GlcA-GlcNS3S6S-IdoA2S-
GlcNS6S-IdoA2S-aMan6S (I0A6-G0S9-I2S6-I2M6)
AT-
type
(Casu et al., 1981; Thunberg et al., 1982)
GlcA-GlcNS3S6S-IdoA±2S-GlcNS6S-IdoA±2S-
GlcNS6S-GlcA-aMan (G0S9-I2/0S6-I2/0S6-G0M0)
Hs3st-1(Pempe et al., 2012)
Δ4,5UA2S-GlcNS-IdoA-GlcNAc6S-GlcA-GlcNS3S-
IdoA2S-GlcNS (D2S0-I0A6-G0S3-I2S0)
AT-
type
(Ototani et al., 1982)
Δ4,5UA-GlcNAc6S-GlcA-GlcNS3S6S-IdoA2S-
GlcNS6S-IdoA2S-GlcNS6S (D0A6-G0S9-I2S6-
I2S6)
AT-
type
(Guerrini et al., 2006; Guerrini et al., 2008; Ototani and Yosizawa, 1981)
Δ4,5UA2S-GlcNS6S-IdoA-GlcNAc6S-GlcA-
GlcNS3S6S-IdoA2S-GlcNS6S (D2S6-I0A6-G0S9-
I2S6)
AT-
type
(Desai et al., 1993; Guerrini et al., 2006; Guerrini et al., 2008)
Δ4,5UA-GlcNS-IdoA2S-GlcNAc-UA2S-GlcNS-
IdoA2S-GlcN3S6S(D0S0-I2A0-U2S0-I2H9)
Hs3st-
3a
(Liu et al., 2002)
Δ4,5UA-GlcNAc6S-GlcA-GlcNS3S6S-IdoA2S-
GlcNS6S-IdoA-GlcNAc6S (D0A6-G0S9-
I2S6-I0A6)
AT-
type
(Guerrini et al., 2008)
Δ4,5UA2S-GlcNS6S-GlcA-GlcNAc6S-GlcA-
GlcNS3S6S-IdoA2S-GlcNS6S (D2S6-G0A6-G0S9-
I2S6)
AT-
type
(Guerrini et al., 2008)
Δ4,5UA-GlcNAc6S-GlcA-GlcNS3S6S-IdoA2S-
GlcNS3S6S-IdoA2S-GlcNS6S (D0A6-G0S9-I2S9-
I2S6)
AT-,
gD-
type
(Guerrini et al., 2013)
Δ4,5UA2S-GlcNS6S-IdoA2S-GlcNS6S-IdoA2S-
GlcNS3S6S-IdoA2S-GlcNS6S (D2S6-I2S6-I2S9-
I2S6)
Hs3st-2(Copeland et al., 2008)
Nonasaccharides
GlcA-GlcNAc6S-GlcA-GlcNS3S6S-IdoA2S-
GlcNS6S-IdoA2S-GlcNS6S-GlcA (G0A6-G0S9-
I2S6-I2S6-G0)
Hs3st-1(Xu et al., 2012b)
GlcNS6S-GlcA-GlcNS3S6S-IdoA±2S-GlcNS6S-
IdoA±2S-GlcNS6S-GlcA-aMan (S6-G0S9-I2/0S6-
I2/0S6-G0M0)
Hs3st-1(Pempe et al., 2012)
Decasaccharides
Δ4,5UA2S-GlcNS6S-IdoA-GlcNAc6S-GlcA-
GlcNS3S6S-IdoA2S-GlcNS6S-IdoA2S-GlcNS6S
(D2S6-I0A6-G0S9-I2S6-I2S6)
AT-
type
(Desai et al., 1993)
GlcA-GlcNS3S6S-GlcA-GlcNS3S6S-GlcA-
GlcNS3S6S-GlcA-GlcNS3S6S-GlcA-aMan (G0S9-
G0S9-G0S9-G0S9-G0M0)
Hs3st-
1, −5
(Liu et al., 2010)
IdoA2S-GlcNS6S-IdoA-GlcNS6S-GlcA-
GlcNS3S6S-IdoA2S-GlcNS6S-IdoA2S-GlcNS6S
(I2S6-I0S6-G0S9-I2S6-I2S6)
AT-
type
(Linhardt et al., 1982)
GlcA-GlcNS3S6S-IdoA±2S-GlcNS6S-IdoA±2S-
GlcNS6S-IdoA±2S-GlcNS6S-GlcA-aMan (G0S9-
I2/0S6-I2/0S6-I2/0S6-G0M0)
Hs3st-1(Pempe et al., 2012)
GlcA-GlcNS3S6S-GlcA±2S-GlcNS6S-GlcA±2S-
GlcNS6S-GlcA±2S-GlcNS6S-GlcA-aMan (G0S9-
G2/0S6-G2/0S6-G2/0S6-G0M0)
Hs3st-1(Pempe et al., 2012)
Dodecasaccharides
GlcA-GlcNS3S6S-IdoA±2S-GlcNS6S-IdoA±2S-
GlcNS6S-IdoA±2S-GlcNS6S-IdoA±2S-GlcNS6S-
GlcA-aMan (G0S9-I2/0S6-I2/0S6-I2/0S6-I2/0S6-
G0M0)
Hs3st-1(Pempe et al., 2012)
GlcA-GlcNS3S6S-GlcA±2S-GlcNS6S-GlcA±2S-
GlcNS6S-GlcA±2S-GlcNS6S-GlcA±2S-GlcNS6S-
GlcA-aMan (G0S9-G2/0S6-G2/0S6-G2/0S6-G2/0S6-
G0M0)
Hs3st-1(Pempe et al., 2012)
Δ4,5UA-GlcNAc6S-GlcA-GlcNS3S6S-IdoA2S-
GlcNS6S-IdoA-GlcNAc6S-GlcA-GlcNS3S6S-
IdoA2S-GlcNS6S (D0A6-G0S9-I2S6-I0A6-G0S9-
I2S6)
AT-
type
(Viskov et al., 2013)

Only oligosaccharides from natural sources or that were generated with purified Hs3st isozymes are included in this table. Two methods are generally used to generate heparin/heparan sulfate oligosaccharides from higher order polysaccharides: Heparin lyases create a 4, 5-unsaturated double bond in the non-reducing end uronic acid (indicated by Δ4,5UA). The stereochemistry of the uronic acid is lost after digestion. Nitrous acid decomposition results in the formation of anhydromannose from N-sulfoglucosamine at pH 1 and from N-unsubstituted glucosamine at pH 4.5 (indicated by aMan).

Heparin lyases, which depolymerize heparin and heparan sulfate into disaccharides by beta-eliminative cleavage of the hexosaminic linkages, have the interesting property of generating “resistant” tetrasaccharide products bearing 3-O-sulfated glucosamine at the reducing end (Shriver et al., 2000; Yamada et al., 1993). A recent study of heparin lyase II showed that the resistance to digestion of the bond between positions −2 and −1 appears to arise from failure to bind the substrate caused by steric hindrance due to the proximity of Asn405 in the active site of the enzyme and the 3-OH group of the bound glucosamine residue (Zhao et al., 2011). Three tetrasaccharides produced by lyase treatment of heparan sulfate modified by Hs3st-1 consist of D0A6-G0S3, D0A6-G0S9 and D0A0-G0M3 (Zhang et al., 1999)1. The tetrasaccharide structures produced by Hs3st-3 have been identified as D2S0-I2H3 and D2S0-I2H9, suggesting that Hs3st-3 modifies N-unsubstituted glucosamines (Liu et al., 1999b). This conclusion is not without controversy. The evidence suggesting an N-unsubstituted glucosamine derives from studies using nitrous acid to depolymerize the chains. At pH 4, nitrous acid treatment results in cleavage only at N-unsubstituted glucosamine residues (Shively and Conrad, 1976). Thus, the sensitivity of the tetrasaccharide to pH 4 nitrous acid suggested that the 3-O-sulfate group was located on an N-unsubstituted glucosamine (Liu et al., 1999b). A later study of Hs3st-3-modified octasaccharides enriched for gD binding came to a similar conclusion (Liu et al., 2002). However, in the original study the tetrasaccharide was also ar sensitive to low pH nitrous acid, which only reacts with N- sulfoglucosamine (Liu, Shriver et al. 1999), suggesting that the pH selectivity of nitrous acid might be altered by 3-O-sulfation (Liu et al., 1999b). Recent NMR experiments identified a persistent hydrogen bond within the AT-pentasaccharide between the internal glucosamine sulfamate NH and the adjacent 3-O-sulfate group, suggesting that the 3-O sulfation may confer certain structural constraints that could affect chemical and enzymatic sensitivity (Langeslay et al., 2012). Recently, we observed that resistant tetrasaccharides derived from CHO cells expressing Hs3st-3 consisted of structures containing GlcNS3S without evidence of N-unsubstituted 3-O-sulfated glucosamine (R.L. and J.D.E., unpublished results). Resistant tetrasaccharides have also been observed in heparan sulfate modified in vitro by Hs3st −2, −4 and −5 (Lawrence et al., 2007; Mochizuki et al., 2008). Limited analytical work on hexasaccharides derived from heparan sulfate modified by Hs3st-4 following partial digest with heparin lyases suggested two structures, D0A0-G0S0-I2S3 and D0A0-G0S0-I2S9 (Wu et al., 2004). Thus, the idea that 3-O-sulfation can occur on unsubstituted glucosamine residues may be incorrect.

Analysis of resistant tetrasaccharides undoubtedly oversimplifies the complexity of domains bearing 3-O-sulfate. Because of the cleavage pattern of the lyases, all Hs3st-modified oligosaccharides studied thus far have the 3-O-sulfate group at the reducing end of the fragment. Thus, there is little information about the oligosaccharide sequence towards the reducing side of the 3-O-sulfation site (i.e. at the +1, +2 and +3 positions). To investigate this question, Hs3st-3 activity has been measured against libraries of defined oligosaccharides (Nguyen et al., 2012). These studies showed that active substrates contained an iduronic acid residue to the reducing side of the acceptor, whereas substrates containing glucuronic acid and 2-sulfo-iduronic acid at this position were inactive. In this respect, the specificity of Hs3st-3 differs from that of Hs3st-1, which permits a 2-sulfo-iduronic acid at the +1 position (Fig. 1). More extensive oligosaccharide libraries are needed to better define the acceptor specificity of the Hs3sts. Recently, Kowalewski et al. reported that arylsulfatase G catalyzes the removal of the 3-O-sulfate groups during lysosomal degradation (Kowalewski et al., 2012). Analysis of the non-reducing ends of the chains that accumulate in the liver showed that heparinase digestion released a significant amount of acetylated trisaccharide with the structure S3-U0A0. Thus the chains that accumulate in different tissues from the mutant could be used to obtain additional information about the structure on the reducing side of the 3-O-sulfated unit.

The possibility that the Hs3sts might also exhibit some promiscuity should also be considered. Thus, under some conditions, a gD-type 3-O-sulfotransferase might create AT-type heparan sulfate and vice versa. Glomerular epithelial cells produce AT-type heparan sulfate and express Hs3st-1 and-3, but none of the other Hs3sts, including Hs3st-5. Nevertheless, when the cells were derived from Hs3st-1−/− mice, low levels of antithrombin-binding heparan sulfate were detected (Girardin et al., 2005). Similarly, small amounts of AT-binding heparan sulfate were produced using recombinant Hs3st-3 in vitro (Girardin et al., 2005).

5. Insights into substrate specificity from structural studies of Hs3sts

X-ray crystal structures have been solved for the sulfotransferase domains of Hs3st-1, −3 and −5 (Edavettal et al., 2004; Moon et al., 2004; Xu et al., 2008). These structures, accompanied by extensive mutational studies, form the basis of our understanding of the working mechanism of this enzyme family and provide insights into the substrate specificity of the enzymes. The overall fold of the Hs3st sulfotransferase domain is very similar to the sulfotransferase domain of N-deacetylase-N-sulfotransferases (Ndst), consistent with ~30 percent sequence identity of these enzymes (Edavettal et al., 2004). All of the sulfotransferases utilize the high-energy sulfate donor, 3'-phosphoadenosine-5'-phosphosulfate (PAPS). Thus, not surprisingly Hs3st and Ndst1 share almost identical structure in the binding site for PAPS, and they both use a conserved glutamate (E184 in Hs3st-3, Fig. 2A) as the catalytic base (Edavettal et al., 2004; Kakuta et al., 2003; Kakuta et al., 1999). The major difference between Hs3st and Ndst is in the heparan sulfate-binding cleft, where significantly more positively charged amino acid residues are present in Hs3st. The difference likely reflects the distinct substrate selectivity of Hs3st and Ndst; Ndsts act on N-acetylated domains of the chains and initiate polymer modification, whereas Hs3sts usually act late requiring a pattern of sulfate groups in the target site (Edavettal et al., 2004).

An external file that holds a picture, illustration, etc.
Object name is nihms551045f2.jpg
Structural interactions of 3-O-sulfotransferases with oligosaccharide substrates

(A) Stereo diagram of the interaction between Hs3st-3 and the tetrasaccharide D2S6-I2S6 (PDB code 1T8U). Side chains of critical amino acid residues that mediate substrate binding are shown as thick grey sticks. Possible hydrogen bonds between side chains and tetrasaccharide are pictured as black broken lines. The C3-OH group of the glucosamine residue to be sulfated is shown as a red sphere. The sulfate donor mimic, 3’-phosphoadenosine 5’-phosphate (PAP) is shown, and the protein backbone is shown as a cartoon in the background. (B) Superposition of Hs3st-3 bound tetrasaccharide (D2S6-I2S6; green, PDB code 1T8U) and heptasaccharide (A6-G0S6-I2S6-G0M0; white, PDB code 3UAN). The surface of Hs3st-1 is rendered in gray. Residues in tetrasaccharides are labeled as T-1 to T-4. Residues in the heptasaccharide are labeled as H-1 to H-6 (the anhydromannose residue is not shown). (C) Role of K259 in Hs3st-3 substrate recognition. The tetrasaccharide and heptasaccharide are colored and numbered as in (B). Putative hydrogen bonds between K259 and the 2-O-sulfate group of T-1 and the carboxylate of T-3 are shown in black dashed lines. The distance between the 2-O-sulfate of T-1 and carboxylate of T-3 are shown as a yellow broken line. The distance between the 2-O-sulfate of T-1 and the carboxylate of H-4 are shown as a red broken line. (D-F) Role of gate residues in substrate recognition. The gate residues are shown as large spheres and labeled in white, and the tetrasaccharide is shown with the C3-OH group shown as a red sphere. The molecular surface of each enzyme is shown in grey. The distance between gates residues is pictured as black broken lines. The tetrasaccharide is modeled into Hs3st-1 (PDB code 1VKJ) and Hs3st-5 (PDB code 3BD9) by superimposing their structures on that of the Hs3st-3/tetrasaccharide complex (PDB code 1T8U) using PyMol (Schrödinger).

Perhaps the most interesting information to glean from structural studies of the Hs3sts is how the isozymes distinguish subtle differences in the target substrates, which in turn allows the enzymes to generate distinct products. Co-crystallization of Hs3st-3, 3’-phosphoadenosine-5’-phosphate (PAP) and a tetrasaccharide substrate has revealed how enzyme structure can influence substrate specificity (Moon et al., 2004). An extensive hydrogen-bonding network was observed in the substrate-binding cleft involving K161, R166, K215, Q255, K368 and R370, the acceptor glucosamine (sulfation-site, sugar residue T-2 in the tetrasaccharide substrate, Fig. 2A) and neighboring uronic acids (T-1 and T-3, Fig. 2A). All six residues are completely conserved in all Hs3st family members and across species. Ndst1 lacks three of the six residues, K161, R166 and Q255, suggesting that they contribute to the recognition of substrates unique to Hs3sts. However, the conservation of these residues across the Hs3st isozymes suggests that they do not dictate substrate specificity across the Hs3sts.

A comparison of Hs3st-1 and Hs3st-3 co-crystal structures with oligosaccharide substrates revealed that the substrates orient differently in the two enzymes (Moon et al., 2012). The orientation of the acceptor glucosamine (position 0, T-2 and H-3 in the hexasaccharide substrate, Fig. 2B) and non-reducing end uronic acid (position −1, T-1 and H-2) were nearly identical in the two crystal structures, but the 2-sulfo-iduronic acid unit at the reducing side of the acceptor glucosamine (position +1, T-3 and H-4) was found in a chair conformation in Hs3st-1 (H-4, Fig 2B), whereas in Hs3st-3 it was found in a skew boat conformation (T-3, Fig 2B). The chair conformation preferred by Hs3st-1 approximates the carboxyl group of 2-sulfo-iduronic acid unit (H-4) and the C2-OH of the uronic acid on the non-reducing side of the acceptor glucosamine (H-2). If the iduronic acid H-2 in the heptasaccharide were 2-O-sulfated, charge repulsion would result between the sulfate and carboxylate groups due to their close proximity (Fig. 2C, red dashed line indicates a distance of only 2.8 Å). In contrast, Hs3st-3 induces a skew boat conformation of the 2-sulfo-iduronic acid T-3 (Fig. 2C), which provides more distance between the carboxylate and the sulfate groups (Fig. 2C, yellow dashed line indicates a distance of 3.2 Å). Importantly, Hs3st-3 also places a lysine residue (K259) between the 2-O-sulfate group of residue T-1 and the carboxylate group of residue T-3, which would help neutralize the negative charges and thus prevent charge repulsion (Fig 3C). \K259 is conserved among all gD-type Hs3sts, suggesting that it might also participate in other gD-type Hs3sts in the selection of substrates containing 2-sulfo-iduronic acid (Lawrence et al., 2007; Liu et al., 1999a; Mochizuki et al., 2008; Wu et al., 2004; Xu et al., 2005). Consistent with this idea, Hs3st-1 contains an uncharged asparagine residue in place of K259 (N167, Fig. 2C), which might explain why Hs3st-1 will not act on substrates bearing 2-sulfo-iduronic acid at the −1 position, but will tolerate either glucuronic or iduronic acid (Mochizuki et al., 2008; Zhang et al., 2001b; Zhang et al., 1999).

An external file that holds a picture, illustration, etc.
Object name is nihms551045f3.jpg
Phylogenetic analysis of Hs3sts

The number of Hs3st isozymes identified in each species was deduced from existing publications, NCBI protein database searches and comparison of key amino acids that define substrate and catalytic sites. The phylogenetic tree depicts their evolutionary relationship (adapted from (Medeiros et al., 2000)). The proposed origin of Hs3st is indicated with an arrow.

When the crystal structures of Hs3st-1, −3 and −5 were superimposed, it was observed that two residues in Hs3st-1, H271 and E88, form a narrow opening that would contact the polysaccharide substrate at multiple sites including the two sugar residues to the non-reducing side of the sulfation site (+2 and +3 positions) (Fig. 2D) (Moon et al., 2012; Xu et al., 2008). The corresponding residues in Hs3st-3 (G182 and G365) and Hs3st-5 (S120 and A306) are much smaller and form a significantly wider opening (Fig. 2E, F). The nature of these “gate” residues seems to regulate the substrate specificities of Hs3st-1 and Hs3st-5 in a substantial way. Mutation of these residues to glycine in Hs3st-1, or to histidine and glutamic acid in Hs3st-5, had little or no effect on catalytic activity, but had a profound effect on substrate specificity. Remarkably, interchanging these two residues in Hs3st-1 and Hs3st-5 was sufficient to convert the substrate specificity of one enzyme into another.

Based on these studies, it is clear that substrate specificity is determined by amino acids distal to the catalytic site. Also, the evidence strongly suggests that Hs3sts recognize at least a pentasaccharide motif spanning from one sugar unit to the reducing end of the sulfation site, to three sugar units to the non-reducing side. Larger oligosaccharides may confer additional specificity. The diversification of the Hs3st family suggests that there may be distinct substrate specificities even within a subgroup of Hs3sts.

6. Evolutionary origin of the Hs3sts

As mentioned above, Homo and other mammals express seven Hs3sts, whereas Danio (zebrafish) expresses eight. In comparison to vertebrates, Hs3st isozymes in invertebrates are much less numerous. Drosophila (fruit fly) and Caenorhabditis (nematode) both have one gD and one AT-type transferase (Kamimura et al., 2004; Tecle et al., 2013), whereas Strongylocentrotus (sea urchin) and Planaria (flatworm) have only a single Hs3st related most closely to an AT-type transferase (Fig. 3). The Nematosetella (sea anemone), a member of the Cnidarian phylum, has two Hs3sts whereas Hydra has only one, but all three share ~50 percent amino acid sequence identity to human Hs3sts. In contrast, Trichoplax, a placozoan generally considered as a sister clade to bilaterians and Cnidarians, do not have any Hs3st homologs, but homologs of all other heparan sulfate sulfotransferases appear to be present. Porifera (sponges) do not express any homologs of the sulfotransferases, suggesting that sponges lack heparan sulfate. Based on this information, we propose that the primordial Hs3st originated early in eumetazoan evolution (excluding Porifera), probably first emerging in a common ancestor of bilaterians and cnidarians (Fig. 3).

Although Hs3sts appear to be widely distributed in nature, demonstration of 3-O-sulfate groups in heparan sulfate derived from non-vertebrate species other than Anomalocardia brasiliana (clam) is lacking. Attreed et al. (Attreed et al., 2012) showed that the single chain antibody (HS4C3), which reacts with a 3-O-sulfated determinant containing 2-O- and 6-O-sulfate groups in heparan sulfate and heparin (Ten Dam et al., 2006), binds to specific cellular targets in C. elegans, suggesting the presence of 3-O-sulfated motifs. We have detected 3-O-sulfated disaccharides and tetrasaccharides in heparan sulfate from zebrafish embryos. These structures were diminished by specific Hs3st morpholinos. Furthermore, expression of cDNAs encoding the different zebrafish Hs3sts in CHO cells results in expression of both AT- and gD-type heparan sulfate. (R.L., H. Joseph Yost, Adam Cadwallader and J.D.E, unpublished findings). The context of the 3-O-sulfate groups in these organisms remains unknown.

7. Expression patterns of the Hs3sts

The expression of Hs3st genes is exquisitely controlled in a spatiotemporal manner in vertebrates, befitting a family of seven isozymes (Table 2). Human HS3ST-1, −3a and −3b transcripts are widely expressed in many organs, whereas expression of HS3ST-2 and −4 has been primarily detected in the brain (Lawrence et al., 2007; Mochizuki et al., 2008; Shworak et al., 1999). HS3ST-5 is expressed in skeletal muscle (Xia et al., 2002). In the mouse, Hs3st-6 is expressed predominantly in the liver and kidney with lower expression in the heart, brain, lung and testis (Xu et al., 2005). At least one Hs3st gene is expressed in nearly every cell line examined thus far and many express multiple Hs3st genes simultaneously (Deligny et al., 2010; Girardin et al., 2005; Vanpouille et al., 2007). Notable exceptions are Chinese hamster ovary cell line and Engelbreth-Holm-Swarm mouse tumor, which synthesize heparan sulfate devoid of 3-O-sulfation (Pejler et al., 1987a; Zhang et al., 2001b).

Table 2

Human HS3ST expression

GeneTissueReference
HS3ST-1Heart, brain, lung, kidney, spleen,
stomach, small intestine, colon, testis
(Mochizuki et al., 2008; Shworak et al., 1999)
HS3ST-2Brain(Mochizuki et al., 2008; Shworak et al., 1999)
HS3ST-3aHeart, placenta, lung, liver, kidney,
spleen, stomach, small intestine, colon,
testis
(Mochizuki et al., 2008; Shworak et al., 1999)
HS3ST-3bHeart, placenta, lung, liver, kidney,
spleen, stomach, colon, testis, pancreas,
skeletal muscle
(Mochizuki et al., 2008; Shworak et al., 1999)
HS3ST-4Brain(Mochizuki et al., 2008; Shworak et al., 1999)
HS3ST-5Brain, spinal cord, skeletal muscle(Mochizuki et al., 2003; Mochizuki et al., 2008; Xia et al., 2002)
HS3ST-6Liver, kidney(Xu et al., 2005)

Particular attention has been focused on Hs3st expression in the nervous system. Several Hs3st genes are spatially regulated in the mouse cerebrum and cerebellum throughout development (Yabe et al., 2005). Whole mount in situ hybridization demonstrated unique expression patterns of each of the Hs3st genes in the Zebrafish brain (Cadwallader and Yost, 2006). Studies of a transgenic mouse, in which the human placental alkaline phosphatase (hPLAP) was inserted next to the start codon of Hs3st-2, showed expression in trigeminal and dorsal root ganglia during development (Hasegawa and Wang, 2008). Expression of Hs3st-2 and −4 appeared associated with a subset of neuronal cells within the trigeminal ganglia (Lawrence et al., 2007). In D. melanogaster and C. elegans, the two Hs3st genes are expressed in specific cells, including small subsets of neurons (Kamimura et al., 2004; Tecle et al., 2013). The spatially and temporally restricted pattern of expression of Hs3st genes in the nervous system suggests a neurological role for 3-O-sulfation. A mild neuronal patterning defect in C. elegans was seen after loss of function of the enzymes (discussed in section 8.6), but to date no such role has been demonstrated in other organisms.

There is a notable lack of information about the factors that regulate the expression of the Hs3st genes (e.g. transcription factors, promoter and enhancer sequences, etc.). This should be a fruitful area of investigation based on the differential expression of the isozymes across different tissues and during development and the observation that external stimuli can affect the expression of the Hs3st genes. For example, Hs3st-2 expression and 3-O-sulfation in the pineal gland varies with Circadian rhythm and is inducible by light (Borjigin et al., 2003; Kuberan et al., 2004). In intestinal epithelium, Hs3st-1 expression increases in response to IL-4 and IL-13 (Takeda et al., 2010). In contrast, stimulation of human microvascular endothelial cells with TNF-a or LPS diminishes expression of HS3ST-1 mRNA (Krenn et al., 2008). Finally, Hs3st-3 was upregulated in urothelium in response to IL-6 (Wood et al., 2011). In most of these examples, it remains unknown if sulfotransferase protein levels, enzyme activity or heparan sulfate structure were also affected. When F9 mouse embryonal carcinoma cells were differentiated into parietal endoderm using retinoic acid, cAMP and theophylline, the expression of Hs3st1 increased more than 100-fold with a corresponding increase in production of heparan sulfate that binds to antithrombin (Zhang et al., 1998).

8. Biochemical and physiological implications of 3-O-sulfation

Hundreds of heparan sulfate-binding proteins have been identified and much effort has been expended to determine the structural features in heparan sulfate that mediate binding (Xu and Esko, 2014). Most studies have overlooked the effect of 3-O-sulfation on binding, generally because of the lack of available material for study. For this reason only six proteins have been demonstrated biochemically to be affected by 3-O-sulfation.

8.1 Antithrombin

Antithrombin is the best characterized protein in which 3-O sulfate groups play a crucial role in binding to heparin/heparan sulfate and it serves as the paradigm for how one can unravel the structural context of relevant, biologically active 3-O-sulfated glucosamine residues (Lindahl et al., 1980). Determining the location of these rare groups in a sea of differentially sulfated domains is a formidable problem, compounded by the fact that obtaining multimilligram amounts of heparan sulfate from any source is quite difficult and costly. The analysis of the antithrombin binding site was greatly simplified because of its enrichment in heparin, which is available in kilogram quantities (Atha et al., 1985; Höök et al., 1976; Hopwood et al., 1976; Lindahl et al., 1979; Lindahl et al., 1983; Lindahl et al., 1980; Riesenfeld et al., 1981). Thus, it was possible to partially cleave the chains into smaller oligosaccharides and to obtain sufficient material for separating fragments into high and low affinity fractions, disaccharide analysis, chemical degradation and NMR studies (Yamada et al., 1993). The antithrombin binding motif was subsequently identified in heparan sulfate expressed by endothelial cells, fibroblasts and in other tissues (De Agostini et al., 1990; Marcum et al., 1986a; Marcum et al., 1986b; Marcum et al., 1983; Marcum and Rosenberg, 1985). It should be pointed out that all of these studies focused on sequences that bind to antithrombin, but in fact heparan sulfate chains contain 3-O-sulfate groups at sites that do not bind with high affinity to antithrombin. In the absence of a ligand like antithrombin, the nature of many of these sites remains uncharacterized.

Antithrombin is an inhibitor of the proteases thrombin, factor IXa and factor Xa and plays a central role in hemostasis. It also plays less appreciated roles as an inhibitor of inflammation and angiogenesis (O'Reilly et al., 1999; Wiedermann Ch and Romisch, 2002). Heparin binding to antithrombin induces a conformational change that increases its catalytic activity by several orders of magnitude (Huntington et al., 1996; Rosenberg and Damus, 1973). Indeed, heparin is routinely used as an anticoagulant in the clinic based on its ability to activate antithrombin.

Much effort went into identifying the minimum active heparin structure, which turns out to be a pentasaccharide with the structure A6-G0S9-I2S6 or A6-G0S3-I2S6 bearing a critical 3-O-sulfate on the middle N-sulfo-glucosamine residue (Fig. 1) (Choay et al., 1983; Thunberg et al., 1982). Based on genetic studies in mice, Hs3st-1 is responsible for the overwhelming majority of antithrombin-binding structures (HajMohammadi et al., 2003). Hs3st-1 is generally thought to be a limiting factor in generating antithrombin binding sites (Zhang et al., 1998). However, in the presence of excess Hs3st-1, the precursor structures become limiting. Evidence also suggests that non-precursor saccharides can inhibit Hs3st-1, which presumably compete with the substrate for binding to the enzyme (Razi and Lindahl, 1995).

The presence of N-acetylglucosamine-6-sulfate and glucuronic acid at the −2 and −1 positions of the pentasaccharide sequence suggests that the binding site occurs in the transition zones between highly sulfated “NS” domains and unmodified “NA” domains (Lindahl et al., 1983; Zhang et al., 2001a). Biochemical studies demonstrated that the 3-O-sulfate group at position 0 and 6-O-sulfate group on the N-acetylglucosamine residue at position −2 account for more than half of the binding energy released when antithrombin interacts with the pentasaccharide (Atha et al., 1985; Atha et al., 1987). The absence of the 3-O-sulfate group decreases affinity by several orders of magnitude (KD = 30 nM versus 500 µM) and reduces both the conformation change induced in antithrombin and the inhibition of Factor Xa (Atha et al., 1985; Atha et al., 1987). The co-crystal structures of antithrombin and bound oligosaccharide suggest that while several basic and polar residues in antithrombin make salt bridges and hydrogen bonds with the pentasaccharide, the most pivotal interaction that stabilizes the complex is the one between residue K114 and the 3-O-sulfate group (Jin et al., 1997; Li et al., 2004). This interaction orients K114 in an optimal conformation to form hydrogen bonds and salt bridges with sugar residues at position +1 and +2 (Fig. 4) (Richard et al., 2009).

An external file that holds a picture, illustration, etc.
Object name is nihms551045f4.jpg
Stereo diagram of pentasaccharide-antithrombin interaction

(PDB code: 2GD4). Heparan sulfate binding residues of antithrombin (salmon) and the pentasaccharide (grey) are shown in sticks, and the antithrombin protein backbone is shown in cartoon (salmon). Putative salt bridges and hydrogen bonds stabilized by the 3- O-sulfate group and K114 are shown as broken lines. Sugar residues in the pentasaccharide are labeled as in Figure 1.

The literature frequently describes antithrombin-binding heparin as “the pentasaccharide structure,” suggesting that only a unique pentasaccharide can bind to antithrombin. This notion is overly simplistic in that other oligosaccharides can bind and activate antithrombin. Some of these structures have higher affinity and result in more robust activation of antithrombin than the minimal pentasaccharide. These include an identical pentasaccharide with a second 3-O-sulfate group on the N-sulfo-glucosamine at position +2 or 3-O-sulfated heparan sulfate devoid of iduronic acid or 2-sulfo-iduronic acid at the + 1 position (Chen et al., 2007; Guerrini et al., 2013; Guerrini et al., 2008; Zhang et al., 2001b).

Only about one-third of typical unfractionated heparin preparations have the ability to bind and activate antithrombin (Höök et al., 1976; Lam et al., 1976). In most tissues examined to date, the percentage of heparan sulfate chains bearing AT-type sequences is quite low. For example, only five percent of heparan sulfate chains from bovine aortic endothelial cells, human umbilical vein endothelial cells and rat microvascular endothelial cells binds with high affinity to antithrombin (Kojima et al., 1992; Marcum et al., 1986a; Mertens et al., 1992). In several tissues examined in the rat, approximately 5–10 percent of heparan sulfate chains contained a high affinity-binding site (Horner, 1990). The brain and spleen have antithrombin binding sites on ~20 percent of the chains (Horner, 1990). Surprisingly, over 50 percent of heparan sulfate chains found in follicular fluid bind with high affinity and activate antithrombin (de Agostini et al., 2008). The production of anticoagulant heparan sulfate proteoglycans in granulosa cells peaks prior to ovulation, linking AT-type heparan sulfate production to the menstrual cycle (Princivalle et al., 2001). Heparin/heparan sulfate with low affinity for antithrombin also contains 3-O-sulfation, apparently in a context that does not support antithrombin activation (Shworak et al., 1997). The purpose of 3-O-sulfation catalyzed by Hs3st-1 in contexts other than antithrombin-binding sites is currently unknown.

Mutations that affect heparin binding to antithrombin underscore the importance of heparan sulfate-induced activation of antithrombin in vivo. Several mutations that interrupt the heparin-binding site of antithrombin have been identified in human patients, some with thrombotic phenotypes (Gandrille et al., 1990; Koide et al., 1984; Lane et al., 1996). Accordingly, a knock-in mouse with a point mutation (R48C) corresponding to a human variant produces antithrombin with reduced affinity for heparin (Dewerchin et al., 2003). This mouse suffers from severe spontaneous thrombosis beginning as early as birth and is unresponsive to heparin. As described above, AT-type 3-O-sulfation unquestionably endows heparan sulfate with high inhibitory activity towards Factor Xa, yet 3-O-sulfation seems dispensable in maintaining hemostatic tone in vivo based on inactivation of Hs3st-1 in the mouse (HajMohammadi et al., 2003). Surprisingly, Hs3st-1−/− mice do not exhibit a prothrombic phenotype despite having drastically reduced levels of AT-type heparan sulfate (HajMohammadi et al., 2003). Two explanations for this finding have been proposed. First, the other 3-O-sulfotransferases (in particular Hs3st-5) may create sufficient AT-type heparan sulfate to compensate for loss of Hs3st-1. Other gD-type 3-O-sulfotransferases can create AT-type heparan sulfate as well, albeit much less efficiently than Hs3st-1 (Girardin et al., 2005; HajMohammadi et al., 2003). Second, studies have shown that suboptimal heparan sulfate structures, including those without 3-O-sulfation, may activate antithrombin sufficiently to maintain hemostasis (Nordenman et al., 1978; Richard et al., 2009; Scully et al., 1988; Streusand et al., 1995).

Antithrombin also has anti-inflammatory and anti-angiogenic properties that may be influenced by 3-O-sulfation (Wiedermann Ch and Romisch, 2002). Hs3st-1−/− mice are hypersensitive to LPS-induced inflammation and antithrombin has a protective effect in the presence of AT-type heparan sulfate (Shworak et al., 2010). The details and mechanism of this effect have not yet been elucidated. Cleaved and latent forms of antithrombin lose their anticoagulant properties but have anti-angiogenic properties (O'Reilly et al., 1999). These forms of antithrombin retain their preference for 3-O-sulfated heparan sulfate and binding to heparan sulfate influences the anti-angiogenic activity (Schedin-Weiss et al., 2008; Zhang et al., 2005). Hs3st-1−/− mice experience intrauterine growth retardation and spontaneous eye degeneration (HajMohammadi et al., 2003), suggesting that other undiscovered ligands may require 3-O-sulfated sequences created by Hs3st-1.

8.2 Herpes Simplex virus glycoprotein D

Glycoprotein D (gD), an envelope glycoprotein of Herpes Simplex virus, facilitates viral fusion by interacting with cellular entry receptors that include 3-O-sulfated heparan sulfate (Shukla et al., 1999). Wild type CHO cells are resistant to HSV-1 infection, but transgenic expression of Hs3st−2, −3a, −3b, −4, −5 and −6 allows HSV infection (O'Donnell C et al., 2006; Shukla et al., 1999; Tiwari et al., 2005; Xia et al., 2002; Xu et al., 2005). Primary human corneal fibroblasts also produce 3-O-sulfated heparan sulfate, which renders them susceptible to HSV-1 entry (Tiwari et al., 2006). The affinity of gD for 3-O-sulfated heparan sulfate is relatively low; the addition of 3-O-sulfate lowers the Kd of gD for heparan sulfate from 43 µM to 2 µM. Nevertheless, 3-O-sulfation results in a dramatic increase of susceptibility to viral infection (Shukla et al., 1999). These findings demonstrate an important principle, namely that 3-O-sulfation can have a profound biological effect without inducing a high affinity binding site for the protein per se.

Recombinant gD has been used to purify a 3-O-sulfated octasaccharide (Liu et al., 2002). The reported structure contained a 3-O-sulfated, N-unsubstituted glucosamine, but this structure should be reevaluated based on the arguments presented earlier (see Section 4) (Liu et al., 1999b). Other octasaccharides derived from heparin can also interact with gD and have been proposed as potentially clinically relevant inhibitors of infection (Copeland et al., 2008; Hu et al., 2011). The binding site for heparan sulfate has been mapped to the N-terminal portion of gD by mutagenesis and crystallography (Carfi et al., 2001; Yoon et al., 2003). Co-crystallization studies are needed to determine how 3-O-sulfation promotes binding of gD to heparan sulfate and facilitates membrane fusion.

8.3 Fibroblast growth factors (FGFs)

There is indirect evidence for a role of 3-O-sulfation in the binding of FGF7 and FGF receptors (FGFR) to heparan sulfate. Heparin was fractionated over antithrombin to obtain high and low affinity material. Only the fraction of heparin with high affinity for antithrombin bound to FGFR1, supported FGF1 or FGF2 binding to FGFR1, and facilitated FGF1-induced DNA synthesis (McKeehan et al., 1999; Ye et al., 2001). This fraction also protected FGF7 from proteolysis better than unfractionated heparin (Ye et al., 2001). Fractionation of size-defined heparin oligosaccharides on immobilized FGF7 produced high affinity octasaccharides that also had anticoagulant activity (Luo et al., 2006). These octasaccharides also enhanced FGF7-stimulated DNA synthesis and intracellular signaling in mouse keratinocytes (Luo et al., 2006). Further information about the structure of the binding sequence or the structure of the complex is not yet available.

In Zebrafish, left-right patterning during early development is driven in part by cilia-dependent fluid flow in Kupffer’s vesicle (Essner et al., 2005). Two Hs3sts independently influence cilia function in the Kupffer’s vesicle (Neugebauer et al., 2013). Morpholino knockdown of Hs3st-5 resulted in decreased cilia length, which showed synergistic effects upon loss of FGF8. FGF8 is a known heparin binding protein, but it remains to be seen if binding of FGF8 to heparan sulfate is influenced by 3-O-sulfation. On the other hand, knockdown of Hs3st-6 results in normal cilia length but impaired cilia movement, which correlated with disruption of dynein organization. Hs3st-3z, −5, −6 and −7 are coexpressed in the cells making up Kupffer’s vesicle, but are not able to compensate for loss of either Hs3st-5 or −6. Likewise, knockdown of Hs3st-3z or −7 did not reproduce the left-right patterning defect. These findings demonstrate distinct roles for multiple 3-O-sulfotransferases coexpressed in the same cells. Furthermore, they demonstrate that individual members of each subgroup of Hs3st may create unique 3-O sulfated sequences with distinct biological properties.

8.4 Cyclophilin B

Cyclophilin B stimulates lymphocyte adhesion and migration upon binding to cell surface heparan sulfate (Allain et al., 2002). Cyclophilin B binds to heparin octasaccharides with high affinity (Kd = 16 nM) (Vanpouille et al., 2007). Like gD, the heparan sulfate binding site for cyclophilin B has been suggested to contain a 3-O-sulfated N-unsubstituted glucosamine (Vanpouille et al., 2007). siRNA mediated knockdown of Hs3st-3a in Jurkat T cells resulted in the loss of cyclophilin B binding to the cell surface and loss of ERK phosphorylation. HeLa cells express Hs3st-2, −3a, −3b, −5 and −6 but do not support cyclophilin binding (Vanpouille et al., 2007). This finding suggests the possibility that heparan sulfate on HeLa cells lacks other structural features (e.g. appropriate N- or O-sulfation) that are required for cyclophilin B binding (Deligny et al., 2010).

8.5 Stabilin

Clearance of circulating heparin occurs primarily in liver sinusoidal endothelial cells by stabilin-1 and −2 in a manner that appears to be enhanced by 3-O sulfation (Borjigin et al., 2003; Pempe et al., 2012). Stabilin-1 and −2 overexpressing cells took up Hs3st-1-modified oligosaccharide more efficiently than an identical non-3-O-sulfated oligosaccharide. Antithrombin was able to inhibit this effect, suggesting that the antithrombin binding sequence can also bind to stabilin. In addition, mice cleared Hs3st-1-modified oligosaccharides faster than oligosaccharides devoid of 3-O-sulfation. Thus, while not required for clearance, 3-O-sulfation facilitates the removal of soluble heparin and presumably heparan sulfate through stabilin-1 and −2.

9. 3-O-Sulfation in development and disease

9.1 3-O-sulfation in development

Evidence is emerging that 3-O-sulfation can modulate various developmental processes in model organisms. As described above, studies of left-right asymmetry implicate Hs3st-5 in FGF8-mediated signaling in Zebrafish. Recent studies show that expression of Hs3st-7 is required for normal development of the Zebrafish heart (Samson et al., 2013). Morpholino knockdown of Hs3st-7 in the developing Zebrafish resulted in disorganization of cardiac contractile apparatus with reduced ventricular contraction. In normal Zebrafish, Hs3st-7 expression somehow restricts the expression of bone morphogenetic protein 4 (BMP4) to the atrioventricular junction. Upon loss of Hs3st-7, BMP4 expression expands throughout the atrium and ventricle and causes loss of cardiac contractility. Although previous studies showed that BMP7 can interact with heparan sulfate (Irie et al., 2003; Midorikawa et al., 2003), this is the first report indicating a role for 3-O-sulfation in binding and sequestration.

Leukemia inhibitory factor (LIF) plays a role in maintaining stem cells by activating signaling through signal transduction and activator of transcription 3 (STAT3). In tissue culture, withdrawal of LIF results in differentiation of mouse embryonic stem cells and a concurrent upregulation of Hs3st-5 (Hirano et al., 2012). That cell surface 3-O-sulfation increased was also suggested by cell surface binding of a single chain antibody (HS4C3) with propensity to bind 3-O-sulfated heparan sulfate (Ten Dam et al., 2006). Interestingly, overexpression of Hs3st-5 induced differentiation of mouse embryonic stem cells even in the presence of LIF. Hs3st-5 presumably acts by increasing 3-O-sulfation of cell surface proteoglycans, which in turn alters signaling. Indeed, overexpression of Hs3st-5 resulted in redistribution of Fas, a member of the tumor necrosis factor receptor family, to lipid rafts on the cell surface. Activation of Fas signaling was inferred by activation of Caspase-3 and degradation of Nanog. Whether 3-O-sulfated heparan sulfate binds to Fas or one or more signaling receptors on the cell surface remains unknown.

C. elegans expresses two 3-O-sulfotransferases, hst-3.1 and hst-3.2 that group phylogenetically with AT-type and gD-type Hs3sts, respectively (Tecle et al., 2013). Loss of function of either 3-O-sulfotransferase resulted in aberrant branching of the hermaphrodite-specific neuron. Furthermore, an extra branching phenotype of the AIY interneuron induced by overexpression of the extracellular adhesion molecule kal-1 depends on the expression of hst-3.2. Interestingly, reexpression of hst-3.2 specifically in the AIY interneuron was not sufficient to restore the phenotype, whereas pan-neuronal or muscle specific rescue of hst-3.2 expression restored the branching phenotype. This finding indicates that 3-O-sulfated heparan sulfate acts in trans (i.e. from a neighboring cell). These results also demonstrate that 3-O-sulfation can influence neuronal patterning in development and suggest kal-1 as a candidate heparin-binding protein that is influenced by 3-O-sulfation.

The effect of 3-O-sulfation has also been investigated in Drosophila. Like C. elegans, Drosophila expresses two 3-O-sulfotransferases (Hs3st-A and -B) that cluster by sequence homology with the mammalian AT-type and gD-type 3-O-sulfotransferases, respectively (Kamimura et al., 2004) (Fig. 3). Loss of either 3-O-sulfotransferase resulted in embryonic or larval lethality. Tissue specific loss of Hs3st-B in Drosophila has been suggested to cause significant structural defects similar to those seen by inactivation of Notch and Notch target genes. These results are surprising because none of the Notch ligands appear to bind to heparan sulfate and obvious Notch phenotypes have not been detected in mice lacking Hs3st-1 or −2 (although this may be explained as compensation by the other 3-O-sulfotransfersases) or in C. elegans lacking all 3-O-sulfation. Conceivably, the dependence of Notch signaling on 3-O-sulfation may be restricted to Drosophila.

9.2 Diseases related to 3-O-sulfation

Aberrant expression of the 3-O-sulfotransferases as a result of DNA hypermethylation is emerging as a common theme in cancer biology. CpG islands located upstream of the transcription start site and within the first exon of the HS3ST-2 gene are hypermethylated in breast, colon, lung and pancreatic cancers (Miyamoto et al., 2003). In the tumor samples studied, HS3ST-2 was essentially silenced as a result of hypermethylation, an effect that was reversed by treatment with a DNA-methyltransferase inhibitor. Another epigenetic study of heparan sulfate biosynthetic enzymes in chondrosarcoma cells showed that HS3ST-1, −2, −3a and −6 were all hypermethylated (Bui et al., 2010). Reversing methylation boosted expression of these HS3STs 1.5- to 2.7-fold and resulted in decreased cell proliferation, increased cell adhesion and decreased cell migration. Hypermethylation of the HS3ST-2 gene has also been detected in B-cell, T-cell and myeloid malignancies, cervical cancer and its immediate precursor, cervical intraepithelial neoplasia (Jiang et al., 2009; Martin-Subero et al., 2009; Shivapurkar et al., 2007). Aberrant methylation of HS3ST-2 may prove useful as a biomarker for early detection of prostate and cervical cancer (Mahapatra et al., 2012; Shivapurkar et al., 2007). Restoration of 3-O-sulfation suppresses tumor growth suggesting that activation of Hs3sts could be a useful therapeutic target in cancer patients. Whether alterations in Hs3st expression in these systems affect the structure of heparan sulfate has not been determined, nor have the proteins influenced by 3-O-sulfation been identified. It remains unclear if DNA methylation is a general mechanism that regulates Hs3st expression under normal physiological conditions.

Pathogens other than HSV-1 may have evolved to co-opt 3-O-sulfation. Linkage analysis revealed an association between Hs3st-3a/Hs3st-3b and sensitivity to parasitemia induced by Plasmodium falciparum (Atkinson et al., 2012). A genome-wide association study linked mother to infant transmission of HIV to Hs3st-3a expression and Hs3st-3b expression is upregulated in patients with HIV-associated dementia (Boven et al., 2007; Joubert et al., 2010). Finally, Hs3st-3a expression may suppress Hepatitis B virus replication in hepatocytes through an unidentified mechanism (Zhang et al., 2010). These interesting results deserve additional consideration to define the relevant ligands and receptors that interact with 3-O-sulfated heparan sulfate.

At least eight lysosomal enzymes are responsible for the degradation of heparan sulfate. Recently, arylsulfatase G was identified as the enzyme that removes 3-O-sulfate groups from non-reducing end N-sulfoglucosamine-3-sulfate (Kowalewski et al., 2012). Like many other heparan sulfate degradative enzymes, loss of arylsulfatase G in mice results in lysosomal accumulation of heparan sulfate and characteristics typical of other mucopolysaccharidoses. Whether loss of 3-O-sulfatase results in a Sanfilippo type syndrome in humans awaits discovery of a patient lacking the enzyme. The 3-O-sulfatase is unable to remove 3-O-sulfate from N-unsubstituted glucosamine-3-sulfate, a putative product of gD-type Hs3sts. If any of the Hs3sts generate N-unsubstituted glucosamine-3 -sulfate or, then there must be another so far unidentified 3-O-sulfatase.

10. Conclusions

Any discussion of heparan sulfate-protein interactions must address the question of specificity (Kreuger et al., 2006; Lindahl and Li, 2009; Xu and Esko, 2014). Antithrombin is clearly unusual in that its biological activity critically depends on high affinity binding to a 3-O-sulfated pentasaccharide. The requirement for the 3-O-sulfate group is so great, that in its absence, the affinity of the interaction drops several orders of magnitude. Stated in another way, the affinity of the interaction is of sufficient magnitude to drive a conformational change adequate to activate antithrombin. Can we expect other ligands to show similar characteristics? Many, but not all, heparan sulfate binding proteins show an “analog” response with respect to charge, i.e. affinity and biological activity increases as the degree of sulfation increases. Thus, one could argue that the presence of a 3-O-sulfate group in the context of other modifications merely increases the local charge density providing another opportunity for non-specific interaction. However, it is difficult to explain the evolution of seven 3-O-sulfotransferases differing in substrate specificity and tissue expression to merely increase charge density. Clearly, the information emerging from genetic studies in model organisms indicates selective and crucial interactions occur between some ligands and 3-O-sulfated heparan sulfate. Future studies should focus on characterizing the physical interaction of these ligands with 3-O-sulfated oligosaccharides and expanding the repertoire of known 3-O-sulfate dependent proteins.

The large number of Hs3st genes and their complex spatial and temporal expression provides a system that has the potential to control ligand binding and signaling in multiple tissues and at different stages of development. To date very few proteins have been described whose activity is enhanced by or dependent upon 3-O sulfation. We also know little about the chemical context in which 3-O-sulfation takes place and how the biosynthetic process is orchestrated to generate the precursor structures required by the Hs3sts. To a large extent, technical issues have limited progress. First, the small amounts of heparan sulfate available from defined sources makes classical fractionation schemes, like that used to define the antithrombin-binding site in heparin, technically demanding. Second, most methods to characterize sites of 3-O-sulfation have been based on a “bottom-up” approach using degradative techniques. Third, structural studies have been limited by the lack of defined standards, which also limits characterization of the sulfotransferases. Fourth, few researchers have focused on 3-O-sulfation, probably due to its rarity and the various analytical and synthetic barriers described above.

Although this assessment of the field may appear gloomy, we believe that the field has turned a corner. New “top-down” analytical methods based on tandem mass spectrometry and ion-mobility methods hold great promise for providing much greater contextual information about the structures of 3-O-sulfation sites (Kailemia et al., 2013; Meissen et al., 2009). Chemical and chemoenzymatic methods have been recently developed that have the capacity to produce multimilligram quantities of defined oligosaccharides, providing much needed reagents to aid in the discovery of new ligands and to characterize the binding sites in the enzymes and in the protein ligands (Arungundram et al., 2009; Xu et al., 2011; Xu et al., 2012). Genetic studies of model organisms focused on 3-O-sulfation have begun to yield interesting insights into the biological function of the Hs3sts and new ligands to study. Inactivation of the genes in cells by silencing methods, TALEN technology and gene targeting in embryonic stem cells provide new tools to study the biological implications of altering 3-O-sulfation in vitro and in vivo. These advances should fuel a resurgence of interest in 3-O-sulfation and eventually lead to a more complete understanding of this relatively rare and mystifying modification.

Highlights

  • 3-O-sulfation of heparan sulfate is a relatively rare modification
  • The 3-O-sulfotranferases constitutes the largest family of heparan sulfate sulfotransferases
  • 3 -O-sulfotranferases bind to different precursor sequences
  • Technical difficulties have made it difficult to determine the structure of 3-O-sulfated sequences
  • Very few have been described that bind to 3-O-sulfated sequences

Acknowledgements

We thank Drs. Jian Liu, H. Joseph Yost, Joseph Zaia and Ulf Lindahl for their critical reading of the manuscript and helpful comments.

Funding Sources: This work was supported by grants GM93131 and HL107150 (to J.D.E.) and by training grant T32CA067754 (to B.E.T.) from the National Institutes of Health and by grant 13BGIA14150008 from the American Heart Association (to D.X.)

Abbreviations

Hs3stheparan sulfate glucosamine 3-O-sulfotransferase
Ndstheparan sulfate N-deacetylase/N- sulfotransferase
PAPS3'-phosphoadenosine-5'-phosphosulfate
PAP3'-phosphoadenosine-5'-phosphate
HSV-1Type 1 Herpes simplex virus
gDglycoprotein D of Type 1 Herpes simplex virus
ATantithrombin
FGFfibroblast growth factor
FGFRfibroblast growth factor receptor
BMPbone morphogenetic protein
LIFleukemia inhibitory factor
STAT3signal transduction and activator of transcription 3
HIVhuman immunodeficiency virus
hPLAPhuman placental alkaline phosphatase

Footnotes

Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final citable form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

Contributions: B.E.T., D.X., R.L., and J.D.E. wrote and edited the paper.

1To simplify the representation of constituent disaccharides, we use a disaccharide structure code (DSC, Lawrence, R., H. Lu, R.D. Rosenberg, J.D. Esko, and L. Zhang. 2008. Disaccharide structure code for the easy representation of constituent oligosaccharides from glycosaminoglycans. Nat Methods. 5:291–292.) In DSC, a uronic acid is designated as U, G, I or D for an unspecified hexuronic acid, D-glucuronic acid, L-iduronic acid or Δ4,5-unsaturated uronic acid, respectively. The N substituent is either H, A, S or R for hydrogen, acetate, sulfate or some other substituent, respectively. The presence and location of ester-linked sulfate groups are depicted by the number of the carbon atom on which the sulfate group is located or by 0 if absent. For example, I2S6 refers to a disaccharide composed of 2-sulfoiduronic acid-N-sulfoglucosamine-6-sulfate, whereas D2S6 refers to a similarly structured disaccharide that instead has a Δ4,5-double bond in the uronic acid. M refers to anhydromannose. The presence of 3-O- and 6-O sulfate on the same hexosamine is indicated by the number 9.

References

  • Allain F, Vanpouille C, Carpentier M, Slomianny MC, Durieux S, Spik G. Interaction with glycosaminoglycans is required for cyclophilin B to trigger integrin-mediated adhesion of peripheral blood T lymphocytes to extracellular matrix. Proc Natl Acad Sci USA. 2002;99:2714–2719. [PMC free article] [PubMed] [Google Scholar]
  • Arungundram S, Al-Mafraji K, Asong J, Leach FE, 3rd, Amster IJ, Venot A, Turnbull JE, Boons GJ. Modular synthesis of heparan sulfate oligosaccharides for structure-activity relationship studies. J Am Chem Soc. 2009;131:17394–17405. [PMC free article] [PubMed] [Google Scholar]
  • Atha DH, Lormeau JC, Petitou M, Rosenberg RD, Choay J. Contribution of monosaccharide residues in heparin binding to antithrombin III. Biochemistry. 1985;24:6723–6729. [PubMed] [Google Scholar]
  • Atha DH, Lormeau JC, Petitou M, Rosenberg RD, Choay J. Contribution of 3-O- and 6-O-sulfated glucosamine residues in the heparin-induced conformational change in antithrombin III. Biochemistry. 1987;26:6454–6461. [PubMed] [Google Scholar]
  • Atkinson A, Garnier S, Afridi S, Fumoux F, Rihet P. Genetic variations in genes involved in heparan sulphate biosynthesis are associated with Plasmodium falciparum parasitaemia: a familial study in Burkina Faso. Malaria journal. 2012;11:108. [PMC free article] [PubMed] [Google Scholar]
  • Attreed M, Desbois M, van Kuppevelt TH, Bulow HE. Direct visualization of specifically modified extracellular glycans in living animals. Nat Methods. 2012;9:477–479. [PMC free article] [PubMed] [Google Scholar]
  • Bishop JR, Schuksz M, Esko JD. Heparan sulphate proteoglycans fine-tune mammalian physiology. Nature. 2007;446:1030–1037. [PubMed] [Google Scholar]
  • Borjigin J, Deng J, Sun X, De Jesus M, Liu T, Wang MM. Diurnal pineal 3-O-sulphotransferase 2 expression controlled by beta-adrenergic repression. J Biol Chem. 2003;278:16315–16319. [PubMed] [Google Scholar]
  • Boven LA, Noorbakhsh F, Bouma G, van der Zee R, Vargas DL, Pardo C, McArthur JC, Nottet HS, Power C. Brain-derived human immunodeficiency virus-1 Tat exerts differential effects on LTR transactivation and neuroimmune activation. J Neurovirol. 2007;13:173–184. [PubMed] [Google Scholar]
  • Bui C, Ouzzine M, Talhaoui I, Sharp S, Prydz K, Coughtrie MW, Fournel-Gigleux S. Epigenetics: methylation-associated repression of heparan sulfate 3-O-sulfotransferase gene expression contributes to the invasive phenotype of H-EMC-SS chondrosarcoma cells. Faseb J. 2010;24:436–450. [PubMed] [Google Scholar]
  • Cadwallader AB, Yost HJ. Combinatorial expression patterns of heparan sulfate sulfotransferases in zebrafish: I. The 3-O-sulfotransferase family. Dev Dyn. 2006;235:3423–3431. [PubMed] [Google Scholar]
  • Carfi A, Willis SH, Whitbeck JC, Krummenacher C, Cohen GH, Eisenberg RJ, Wiley DC. Herpes simplex virus glycoprotein D bound to the human receptor HveA. Mol Cell. 2001;8:169–179. [PubMed] [Google Scholar]
  • Chen J, Duncan MB, Carrick K, Pope RM, Liu J. Biosynthesis of 3-O-sulfated heparan sulfate: unique substrate specificity of heparan sulfate 3-O-sulfotransferase isoform 5. Glycobiology. 2003;13:785–794. [PubMed] [Google Scholar]
  • Chen J, Jones CL, Liu J. Using an enzymatic combinatorial approach to identify anticoagulant heparan sulfate structures. Chem Biol. 2007;14:986–993. [PMC free article] [PubMed] [Google Scholar]
  • Chen J, Liu J. Characterization of the structure of antithrombin-binding heparan sulfate generated by heparan sulfate 3-O-sulfotransferase 5. Biochim Biophys Acta. 2005;1725:190–200. [PubMed] [Google Scholar]
  • Choay J, Petitou M, Lormeau JC, Sinay P, Casu B, Gatti G. Structure-activity relationship in heparin: a synthetic pentasaccharide with high affinity for antithrombin III and eliciting high anti-factor Xa activity. Biochem Biophys Res Commun. 1983;116:492–499. [PubMed] [Google Scholar]
  • Copeland R, Balasubramaniam A, Tiwari V, Zhang F, Bridges A, Linhardt RJ, Shukla D, Liu J. Using a 3-O-sulfated heparin octasaccharide to inhibit the entry of herpes simplex virus type 1. Biochemistry. 2008;47:5774–5783. [PMC free article] [PubMed] [Google Scholar]
  • De Agostini A, Watkins SC, Slayter HS, Youssoufian H, Rosenberg RD. Localization of anticoagulantly active heparan sulfate proteoglycans in vascular endothelium: Antithrombin binding on cultured endothelial cells and perfused rat aorta. J.Cell Biol. 1990;111:1293–1304. [PMC free article] [PubMed] [Google Scholar]
  • de Agostini AI, Dong JC, de Vantery Arrighi C, Ramus MA, Dentand-Quadri I, Thalmann S, Ventura P, Ibecheole V, Monge F, Fischer AM, HajMohammadi S, Shworak NW, Zhang L, Zhang Z, Linhardt RJ. Human follicular fluid heparan sulfate contains abundant 3-O-sulfated chains with anticoagulant activity. J Biol Chem. 2008;283:28115–28124. [PMC free article] [PubMed] [Google Scholar]
  • Deligny A, Denys A, Marcant A, Melchior A, Mazurier J, van Kuppevelt TH, Allain F. Synthesis of heparan sulfate with cyclophilin B-binding properties is determined by cell type-specific expression of sulfotransferases. J Biol Chem. 2010;285:1701–1715. [PMC free article] [PubMed] [Google Scholar]
  • Dewerchin M, Herault JP, Wallays G, Petitou M, Schaeffer P, Millet L, Weitz JI, Moons L, Collen D, Carmeliet P, Herbert JM. Life-threatening thrombosis in mice with targeted Arg48-to-Cys mutation of the heparin-binding domain of antithrombin. Circ Res. 2003;93:1120–1126. [PubMed] [Google Scholar]
  • Duncan MB, Chen J, Krise JP, Liu J. The biosynthesis of anticoagulant heparan sulfate by the heparan sulfate 3-O-sulfotransferase isoform 5. Biochim Biophys Acta. 2004;1671:34–43. [PubMed] [Google Scholar]
  • Edavettal SC, Lee KA, Negishi M, Linhardt RJ, Liu J, Pedersen LC. Crystal structure and mutational analysis of heparan sulfate 3-O-sulfotransferase isoform 1. J Biol Chem. 2004;279:25789–25797. [PubMed] [Google Scholar]
  • Edge AS, Spiro RG. Characterization of novel sequences containing 3-O-sulfated glucosamine in glomerular basement membrane heparan sulfate and localization of sulfated disaccharides to a peripheral domain. J Biol Chem. 1990;265:15874–15881. [PubMed] [Google Scholar]
  • Esko JD, Selleck SB. Order out of chaos: Assembly of ligand binding sites in heparan sulfate. Annu Rev Biochem. 2002;71:435–471. [PubMed] [Google Scholar]
  • Essner JJ, Amack JD, Nyholm MK, Harris EB, Yost HJ. Kupffer's vesicle is a ciliated organ of asymmetry in the zebrafish embryo that initiates left-right development of the brain, heart and gut. Development. 2005;132:1247–1260. [PubMed] [Google Scholar]
  • Gandrille S, Aiach M, Lane DA, Vidaud D, Molho-Sabatier P, Caso R, de Moerloose P, Fiessinger JN, Clauser E. Important role of arginine 129 in heparin-binding site of antithrombin III. Identification of a novel mutation arginine 129 to glutamine. J Biol Chem. 1990;265:18997–19001. [PubMed] [Google Scholar]
  • Girardin EP, Hajmohammadi S, Birmele B, Helisch A, Shworak NW, de Agostini AI. Synthesis of anticoagulantly active heparan sulfate proteoglycans by glomerular epithelial cells involves multiple 3-O-sulfotransferase isoforms and a limiting precursor pool. J Biol Chem. 2005;280:38059–38070. [PubMed] [Google Scholar]
  • Guerrini M, Elli S, Mourier P, Rudd TR, Gaudesi D, Casu B, Boudier C, Torri G, Viskov C. An unusual antithrombin-binding heparin octasaccharide with an additional 3-O-sulfated glucosamine in the active pentasaccharide sequence. Biochem J. 2013;449:343–351. [PubMed] [Google Scholar]
  • Guerrini M, Guglieri S, Casu B, Torri G, Mourier P, Boudier C, Viskov C. Antithrombin-binding octasaccharides and role of extensions of the activepentasaccharide sequence in the specificity and strength of interaction. Evidence for very high affinity induced by an unusual glucuronic acid residue. J Biol Chem. 2008;283:26662–26675. [PMC free article] [PubMed] [Google Scholar]
  • HajMohammadi S, Enjyoji K, Princivalle M, Christi P, Lech M, Beeler D, Rayburn H, Schwartz JJ, Barzegar S, De Agostini AI, Post MJ, Rosenberg RD, Shworak NW. Normal levels of anticoagulant heparan sulfate are not essential for normal hemostasis. J Clin Invest. 2003;111:989–999. [PMC free article] [PubMed] [Google Scholar]
  • Hasegawa H, Wang F. Visualizing mechanosensory endings of TrkC-expressing neurons in HS3ST-2-hPLAP mice. J Comp Neurol. 2008;511:543–556. [PMC free article] [PubMed] [Google Scholar]
  • Hirano K, Sasaki N, Ichimiya T, Miura T, Van Kuppevelt TH, Nishihara S. 3-O-sulfated heparan sulfate recognized by the antibody HS4C3 contribute to the differentiation of mouse embryonic stem cells via Fas signaling. PLoS ONE. 2012;7:e43440. [PMC free article] [PubMed] [Google Scholar]
  • Höök M, Bjork I, Hopwood J, Lindahl U. Anticoagulant activity of heparin: separation of high-activity and low-activity heparin species by affinity chromatography on immobilized antithrombin. FEBS Lett. 1976;66:90–93. [PubMed] [Google Scholar]
  • Hopwood J, Höök M, Linker A, Lindahl U. Anticoagulant activity of heparin: isolation of antithrombin-binding sites. FEBS Lett. 1976;69:51–54. [PubMed] [Google Scholar]
  • Horner AA. Rat heparan sulphates. A study of the antithrombin-binding properties of heparan sulphate chains from rat adipose tissue, brain, carcase, heart, intestine, kidneys, liver, lungs, skin and spleen. Biochem J. 1990;266:553–559. [PMC free article] [PubMed] [Google Scholar]
  • Hu YP, Lin SY, Huang CY, Zulueta MM, Liu JY, Chang W, Hung SC. Synthesis of 3-O-sulfonated heparan sulfate octasaccharides that inhibit the herpes simplex virus type 1 host-cell interaction. Nat Chem. 2011;3:557–563. [PubMed] [Google Scholar]
  • Huntington JA, Olson ST, Fan BQ, Gettins PGW. Mechanism of heparin activation of antithrombin. Evidence for reactive center loop preinsertion with expulsion upon heparin binding. Biochemistry. 1996;35:8495–8503. [PubMed] [Google Scholar]
  • Irie A, Habuchi H, Kimata K, Sanai Y. Heparan sulfate is required for bone morphogenetic protein-7 signaling. Biochem Biophys Res Commun. 2003;308:858–865. [PubMed] [Google Scholar]
  • Jiang Y, Dunbar A, Gondek LP, Mohan S, Rataul M, O'Keefe C, Sekeres M, Saunthararajah Y, Maciejewski JP. Aberrant DNA methylation is a dominant mechanism in MDS progression to AML. Blood. 2009;113:1315–1325. [PMC free article] [PubMed] [Google Scholar]
  • Jin L, Abrahams JP, Skinner R, Petitou M, Pike RN, Carrell RW. The anticoagulant activation of antithrombin by heparin. Proc Natl Acad Sci USA. 1997;94:14683–14688. [PMC free article] [PubMed] [Google Scholar]
  • Joubert BR, Lange EM, Franceschini N, Mwapasa V, North KE, Meshnick SR. A whole genome association study of mother-to-child transmission of HIV in Malawi. Genome Med. 2010;2:17. [PMC free article] [PubMed] [Google Scholar]
  • Kailemia MJ, Li L, Xu Y, Liu J, Linhardt RJ, Amster IJ. Structurally informative tandem mass spectrometry of highly sulfated natural and chemoenzymatically synthesized heparin and heparan sulfate glycosaminoglycans. Mol Cell Proteomics. 2013;12:979–990. [PMC free article] [PubMed] [Google Scholar]
  • Kakuta Y, Li L, Pedersen LC, Pedersen LG, Negishi M. Heparan sulphate N-sulphotransferase activity: reaction mechanism and substrate recognition. Biochem Soc Trans. 2003;31:331–334. [PubMed] [Google Scholar]
  • Kakuta Y, Sueyoshi T, Negishi M, Pedersen LC. Crystal structure of the sulfotransferase domain of human heparan sulfate N-deacetylase/N-sulfotransferase 1. J Biol Chem. 1999;274:10673–10676. [PubMed] [Google Scholar]
  • Kamimura K, Rhodes JM, Ueda R, McNeely M, Shukla D, Kimata K, Spear PG, Shworak NW, Nakato H. Regulation of Notch signaling by Drosophila heparan sulfate 3-O sulfotransferase. J Cell Biol. 2004;166:1069–1079. [PMC free article] [PubMed] [Google Scholar]
  • Koide T, Odani S, Takahashi K, Ono T, Sakuragawa N. Antithrombin III Toyama: replacement of arginine-47 by cysteine in hereditary abnormal antithrombin III that lacks heparin-binding ability. Proc Natl Acad Sci U S A. 1984;81:289–293. [PMC free article] [PubMed] [Google Scholar]
  • Kojima T, Leone CW, Marchildon GA, Marcum JA, Rosenberg RD. Isolation and characterization of heparan sulfate proteoglycans produced by cloned rat microvascular endothelial cells. J Biol Chem. 1992;267:4859–4869. [PubMed] [Google Scholar]
  • Kowalewski B, Lamanna WC, Lawrence R, Damme M, Stroobants S, Padva M, Kalus I, Frese MA, Lubke T, Lullmann-Rauch R, D'Hooge R, Esko JD, Dierks T. Arylsulfatase G inactivation causes loss of heparan sulfate 3-O-sulfatase activity and mucopolysaccharidosis in mice. Proc Natl Acad Sci U S A. 2012;109:10310–10315. [PMC free article] [PubMed] [Google Scholar]
  • Krenn EC, Wille I, Gesslbauer B, Poteser M, van Kuppevelt TH, Kungl AJ. Glycanogenomics: a qPCR-approach to investigate biological glycan function. Biochem Biophys Res Commun. 2008;375:297–302. [PubMed] [Google Scholar]
  • Kreuger J, Spillmann D, Li JP, Lindahl U. Interactions between heparan sulfate and proteins: the concept of specificity. J Cell Biol. 2006;174:323–327. [PMC free article] [PubMed] [Google Scholar]
  • Kuberan B, Lech M, Borjigin J, Rosenberg RD. Light-induced 3-O-sulfotransferase expression alters pineal heparan sulfate fine structure. A surprising link to circadian rhythm. J Biol Chem. 2004;279:5053–5054. [PubMed] [Google Scholar]
  • Kusche M, Bäckström G, Riesenfeld J, Petitou M, Chaoy J, Lindahl U. Biosynthesis of heparin. O-sulfation of the antithrombin-binding region. J Biol Chem. 1988;263:15474–15484. [PubMed] [Google Scholar]
  • Lam LH, Silbert JE, Rosenberg RD. The separation of active and inactive forms of heparin. Biochem Biophys Res Commun. 1976;69:570–577. [PubMed] [Google Scholar]
  • Lane DA, Kunz G, Olds RJ, Thein SL. Molecular genetics of antithrombin deficiency. Blood reviews. 1996;10:59–74. [PubMed] [Google Scholar]
  • Langeslay DJ, Young RP, Beni S, Beecher CN, Mueller LJ, Larive CK. Sulfamate proton solvent exchange in heparin oligosaccharides: evidence for a persistent hydrogen bond in the antithrombin-binding pentasaccharide Arixtra. Glycobiology. 2012;22:1173–1182. [PubMed] [Google Scholar]
  • Lawrence R, Lu H, Rosenberg RD, Esko JD, Zhang L. Disaccharide structure code for the easy representation of constituent oligosaccharides from glycosaminoglycans. Nat Methods. 2008;5:291–292. [PubMed] [Google Scholar]
  • Lawrence R, Yabe T, Hajmohammadi S, Rhodes J, McNeely M, Liu J, Lamperti ED, Toselli PA, Lech M, Spear PG, Rosenberg RD, Shworak NW. The principal neuronal gD-type 3-O-sulfotransferases and their products in central and peripheral nervous system tissues. Matrix Biol. 2007;26:442–455. [PMC free article] [PubMed] [Google Scholar]
  • Leder IG. A novel 3-O sulfatase from human urine acting on methyl-2-deoxy-2-sulfamino-alphs-D-glucopyranoside 3-sulfate. Biochem Biophys Res Commun. 1980;94:1183–1189. [PubMed] [Google Scholar]
  • Li W, Johnson DJ, Esmon CT, Huntington JA. Structure of the antithrombin-thrombin-heparin ternary complex reveals the antithrombotic mechanism of heparin. Nat Struct Mol Biol. 2004;11:857–862. [PubMed] [Google Scholar]
  • Lindahl U, Backstrom G, Höök M, Thunberg L, Fransson LA, Linker A. Structure of the antithrombin-binding site in heparin. Proc.Natl.Acad.Sci.U.S.A. 1979;76:3198–3202. [PMC free article] [PubMed] [Google Scholar]
  • Lindahl U, Backstrom G, Thunberg L. The antithrombin-binding sequence in heparin. Identification of an essential 6-O-sulfate group. J Biol Chem. 1983;258:9826–9830. [PubMed] [Google Scholar]
  • Lindahl U, Backstrom G, Thunberg L, Leder IG. Evidence for a 3-O-sulfated D-glucosamine residue in the antithrombin-binding sequence of heparin. Proc Natl Acad Sci USA. 1980;77:6551–6555. [PMC free article] [PubMed] [Google Scholar]
  • Lindahl U, Li JP. Interactions between heparan sulfate and proteins-design and functional implications. Int Rev Cell Mol Biol. 2009;276:105–159. [PubMed] [Google Scholar]
  • Liu J, Pedersen LC. Anticoagulant heparan sulfate: structural specificity and biosynthesis. Appl Microbiol Biotechnol. 2007;74:263–272. [PMC free article] [PubMed] [Google Scholar]
  • Liu J, Shriver Z, Pope RM, Thorp SC, Duncan MB, Copeland RJ, Raska CS, Yoshida K, Eisenberg RJ, Cohen G, Linhardt RJ, Sasisekharan R. Characterization of a heparan sulfate octasaccharide that binds to herpes simplex virus type 1 glycoprotein D. J Biol Chem. 2002;277:33456–33467. [PubMed] [Google Scholar]
  • Liu J, Shworak NW, Fritze LMS, Edelberg JM, Rosenberg RD. Purification of heparan sulfate D-glucosaminyl 3-O-sulfotransferase. J Biol Chem. 1996;271:27072–27082. [PubMed] [Google Scholar]
  • Liu J, Shworak NW, Sinay P, Schwartz JJ, Zhang L, Fritze LM, Rosenberg RD. Expression of heparan sulfate D-glucosaminyl 3-O-sulfotransferase isoforms reveals novel substrate specificities. J Biol Chem. 1999a;274:5185–5192. [PubMed] [Google Scholar]
  • Liu JA, Shriver Z, Blaiklock P, Yoshida K, Sasisekharan R, Rosenberg RD. Heparan sulfate D-glucosaminyl 3-O-sulfotransferase-3A sulfates N-unsubstituted glucosamine residues. J Biol Chem. 1999b;274:38155–38162. [PubMed] [Google Scholar]
  • Luo Y, Ye S, Kan M, McKeehan WL. Structural specificity in a FGF7-affinity purified heparin octasaccharide required for formation of a complex with FGF7 and FGFR2IIIb. J Cell Biochem. 2006;97:1241–1258. [PubMed] [Google Scholar]
  • Mahapatra S, Klee EW, Young CY, Sun Z, Jimenez RE, Klee GG, Tindall DJ, Donkena KV. Global methylation profiling for risk prediction of prostate cancer. Clin Cancer Res. 2012;18:2882–2895. [PubMed] [Google Scholar]
  • Marcum JA, Atha DH, Fritze LM, Nawroth P, Stern D, Rosenberg RD. Cloned bovine aortic endothelial cells synthesize anticoagulantly active heparan sulfate proteoglycan. J Biol Chem. 1986a;261:7507–7517. [PubMed] [Google Scholar]
  • Marcum JA, Conway EM, Youssoufian H, Rosenberg RD. Anticoagulantly active heparin-like molecules from cultured fibroblasts. Exp.Cell Res. 1986b;166:253–258. [PubMed] [Google Scholar]
  • Marcum JA, Fritze L, Galli SJ, Karp G, Rosenberg RD. Microvascular heparin-like species with anticoagulant activity. Am.J.Physiol. 1983;245:H725–H733. [PubMed] [Google Scholar]
  • Marcum JA, Rosenberg RD. Heparinlike molecules with anticoagulant activity are synthesized by cultured endothelial cells. Biochem Biophys Res Commun. 1985;126:365–372. [PubMed] [Google Scholar]
  • Martin-Subero JI, Ammerpohl O, Bibikova M, Wickham-Garcia E, Agirre X, Alvarez S, Bruggemann M, Bug S, Calasanz MJ, Deckert M, Dreyling M, Du MQ, Durig J, Dyer MJ, Fan JB, Gesk S, Hansmann ML, Harder L, Hartmann S, Klapper W, Kuppers R, Montesinos-Rongen M, Nagel I, Pott C, Richter J, Roman-Gomez J, Seifert M, Stein H, Suela J, Trumper L, Vater I, Prosper F, Haferlach C, Cruz Cigudosa J, Siebert R. A comprehensive microarray-based DNA methylation study of 367 hematological neoplasms. PLoS One. 2009;4:e6986. [PMC free article] [PubMed] [Google Scholar]
  • McKeehan WL, Wu XC, Kan M. Requirement for anticoagulant heparan sulfate in the fibroblast growth factor receptor complex. J Biol Chem. 1999;274:21511–21514. [PubMed] [Google Scholar]
  • Medeiros GF, Mendes A, Castro RAB, Baú EC, Nader HB, Dietrich CP. Distribution of sulfated glycosaminoglycans in the animal kingdom: widespread occurrence of heparin-like compounds in invertebrates. Biochim.Biophys.Acta Gen.Subj. 2000;1475:287–294. [PubMed] [Google Scholar]
  • Meissen JK, Sweeney MD, Girardi M, Lawrence R, Esko JD, Leary JA. Differentiation of 3-O-sulfated heparin disaccharide isomers: identification of structural aspects of the heparin CCL2 binding motif. J Am Soc Mass Spectrom. 2009;20:652–657. [PMC free article] [PubMed] [Google Scholar]
  • Mertens G, Cassiman JJ, Van den Berghe H, Vermylen J, David G. Cell surface heparan sulfate proteoglycans from human vascular endothelial cells. Core protein characterization and antithrombin III binding properties. J.Biol.Chem. 1992;267:20435–20443. [PubMed] [Google Scholar]
  • Meyer B, Thunberg L, Lindahl U, Larm O, Leder IG. The antithrombin-binding sequence of heparin studied by n.m.r. spectroscopy. Carbohydr Res. 1981;88:C1–C4. [PubMed] [Google Scholar]
  • Midorikawa Y, Ishikawa S, Iwanari H, Imamura T, Sakamoto H, Miyazono K, Kodama T, Makuuchi M, Aburatani H. Glypican-3, overexpressed in hepatocellular carcinoma, modulates FGF2 and BMP-7 signaling. Int J Cancer. 2003;103:455–465. [PubMed] [Google Scholar]
  • Miyamoto K, Asada K, Fukutomi T, Okochi E, Yagi Y, Hasegawa T, Asahara T, Sugimura T, Ushijima T. Methylation-associated silencing of heparan sulfate D-glucosaminyl 3-O-sulfotransferase-2 (3-OST-2) in human breast, colon, lung and pancreatic cancers. Oncogene. 2003;22:274–280. [PubMed] [Google Scholar]
  • Mochizuki H, Yoshida K, Gotoh M, Sugioka S, Kikuchi N, Kwon YD, Tawada A, Maeyama K, Inaba N, Hiruma T, Kimata K, Narimatsu H. Characterization of a heparan sulfate 3-O-sulfotransferase-5, an enzyme synthesizing a tetrasulfated disaccharide. J Biol Chem. 2003;278:26780–26787. [PubMed] [Google Scholar]
  • Mochizuki H, Yoshida K, Shibata Y, Kimata K. Tetrasulfated disaccharide unit in heparan sulfate: enzymatic formation and tissue distribution. J Biol Chem. 2008;283:31237–31245. [PMC free article] [PubMed] [Google Scholar]
  • Moon AF, Edavettal SC, Krahn JM, Munoz EM, Negishi M, Linhardt RJ, Liu J, Pedersen LC. Structural analysis of the sulfotransferase (3-o-sulfotransferase isoform 3) involved in the biosynthesis of an entry receptor for herpes simplex virus 1. J Biol Chem. 2004;279:45185–45193. [PMC free article] [PubMed] [Google Scholar]
  • Moon AF, Xu Y, Woody SM, Krahn JM, Linhardt RJ, Liu J, Pedersen LC. Dissecting the substrate recognition of 3-O-sulfotransferase for the biosynthesis of anticoagulant heparin. Proc Natl Acad Sci U S A. 2012;109:5265–5270. [PMC free article] [PubMed] [Google Scholar]
  • Neugebauer JM, Cadwallader AB, Amack JD, Bisgrove BW, Yost HJ. Differential roles for 3-OSTs in the regulation of cilia length and motility. Development. 2013;140:3892–3902. [PMC free article] [PubMed] [Google Scholar]
  • Nguyen TK, Arungundram S, Tran VM, Raman K, Al-Mafraji K, Venot A, Boons GJ, Kuberan B. A synthetic heparan sulfate oligosaccharide library reveals the novel enzymatic action of d-glucosaminyl 3-O-sulfotransferase-3a. Mol Biosyst. 2012;8:609–614. [PMC free article] [PubMed] [Google Scholar]
  • Nordenman B, Danielsson A, Bjork I. The binding of low-affinity and high-affinity heparin to antithrombin. Fluorescence studies. Eur J Biochem. 1978;90:1–6. [PubMed] [Google Scholar]
  • O'Donnell C D, Tiwari V, Oh MJ, Shukla D. A role for heparan sulfate 3-O-sulfotransferase isoform 2 in herpes simplex virus type 1 entry and spread. Virology. 2006;346:452–459. [PubMed] [Google Scholar]
  • O'Reilly MS, Pirie-Shepherd S, Lane WS, Folkman J. Antiangiogenic activity of the cleaved conformation of the serpin antithrombin. Science. 1999;285:1926–1928. [PubMed] [Google Scholar]
  • Ori A, Wilkinson MC, Fernig DG. The heparanome and regulation of cell function: structures, functions and challenges. Front Biosci. 2008;13:4309–4338. [PubMed] [Google Scholar]
  • Pejler G, B:ackström G, Lindahl U, Paulsson M, Dziadek M, Fujiwara S, Timpl R. Structure and affinity for antithrombin of heparan sulfate chains derived from basement membrane proteoglycans. J Biol Chem. 1987a;262:5036–5043. [PubMed] [Google Scholar]
  • Pejler G, Danielsson, Björk I, Lindahl U, Nader HB, Dietrich CP. Structure and antithrombin-binding properties of heparin isolated from the clams Anomalocardia brasiliana and Tivela mactroides . J Biol Chem. 1987b;262:11413–11421. [PubMed] [Google Scholar]
  • Pempe EH, Xu Y, Gopalakrishnan S, Liu J, Harris EN. Probing structural selectivity of synthetic heparin binding to stabilin protein receptors. J Biol Chem. 2012;287:20774–20783. [PMC free article] [PubMed] [Google Scholar]
  • Princivalle M, Hasan S, Hosseini G, de Agostini AI. Anticoagulant heparan sulfate proteoglycans expression in the rat ovary peaks in preovulatory granulosa cells. Glycobiology. 2001;11:183–194. [PubMed] [Google Scholar]
  • Razi N, Lindahl U. Biosynthesis of heparin/heparan sulfate. The D-glucosaminyl 3-O-sulfotransferase reaction: Target and inhibitor saccharides. J Biol Chem. 1995;270:11267–11275. [PubMed] [Google Scholar]
  • Richard B, Swanson R, Olson ST. The signature 3-O-sulfo group of the anticoagulant heparin sequence is critical for heparin binding to antithrombin but is not required for allosteric activation. J Biol Chem. 2009;284:27054–27064. [PMC free article] [PubMed] [Google Scholar]
  • Riesenfeld J, Thunberg L, Höök M, Lindahl U. The antithrombin-binding sequence of heparin. Location of essential N-sulfate groups. J Biol Chem. 1981;256:2389–2394. [PubMed] [Google Scholar]
  • Rosenberg RD, Damus PS. The purification and mechanism of action of human antithrombin-heparin cofactor. J Biol Chem. 1973;248:6490–6505. [PubMed] [Google Scholar]
  • Samson SC, Ferrer T, Jou CJ, Sachse FB, Shankaran SS, Shaw RM, Chi NC, Tristani-Firouzi M, Yost HJ. 3-OST-7 regulates BMP-dependent cardiac contraction. PLoS Biol. 2013 in press. [PMC free article] [PubMed] [Google Scholar]
  • Schedin-Weiss S, Richard B, Hjelm R, Olson ST. Antiangiogenic forms of antithrombin specifically bind to the anticoagulant heparin sequence. Biochemistry. 2008;47:13610–13619. [PMC free article] [PubMed] [Google Scholar]
  • Scully MF, Ellis V, Kakkar VV. Heparan sulphate with no affinity for antithrombin III and the control of haemostasis. FEBS Lett. 1988;241:11–14. [PubMed] [Google Scholar]
  • Shivapurkar N, Sherman ME, Stastny V, Echebiri C, Rader JS, Nayar R, Bonfiglio TA, Gazdar AF, Wang SS. Evaluation of candidate methylation markers to detect cervical neoplasia. Gynecologic oncology. 2007;107:549–553. [PMC free article] [PubMed] [Google Scholar]
  • Shively JE, Conrad HE. Formation of anhydrosugars in the chemical depolymerization of heparin. Biochemistry. 1976;15:3932–3942. [PubMed] [Google Scholar]
  • Shriver Z, Sundaram M, Venkataraman G, Fareed A, Linhardt R, Biemann K, Sasisekharan R. Cleavage of the antithrombin III binding site in heparin by heparinases and its implication in the generation of low molecular weight heparin. Proc Natl Acad Sci USA. 2000;97:10365–10370. [PMC free article] [PubMed] [Google Scholar]
  • Shukla D, Liu J, Blaiklock P, Shworak NW, Bai XM, Esko JD, Cohen GH, Eisenberg RJ, Rosenberg RD, Spear PG. A novel role for 3-O-sulfated heparan sulfate in herpes simplex virus 1 entry. Cell. 1999;99:13–22. [PubMed] [Google Scholar]
  • Shworak NW, Kobayashi T, de Agostini A, Smits NC. Anticoagulant heparan sulfate to not clot--or not? Prog Mol Biol Transl Sci. 2010;93:153–178. [PubMed] [Google Scholar]
  • Shworak NW, Liu J, Fritze LMS, Schwartz JJ, Zhang LJ, Logeart D, Rosenberg RD. Molecular cloning and expression of mouse and human cDNAs encoding heparan sulfate D-glucosaminyl 3-O-sulfotransferase. J Biol Chem. 1997;272:28008–28019. [PubMed] [Google Scholar]
  • Shworak NW, Liu JA, Petros LM, Zhang LJ, Kobayashi M, Copeland NG, Jenkins NA, Rosenberg RD. Multiple isoforms of heparan sulfate D-glucosaminyl 3-O-sulfotransferase - Isolation, characterization, and expression of human cDNAs and identification of distinct genomic loci. J Biol Chem. 1999;274:5170–5184. [PubMed] [Google Scholar]
  • Streusand VJ, Bjork I, Gettins PG, Petitou M, Olson ST. Mechanism of acceleration of antithrombin-proteinase reactions by low affinity heparin. Role of the antithrombin binding pentasaccharide in heparin rate enhancement. J Biol Chem. 1995;270:9043–9051. [PubMed] [Google Scholar]
  • Takeda K, Hashimoto K, Uchikawa R, Tegoshi T, Yamada M, Arizono N. Direct effects of IL-4/IL-13 and the nematode Nippostrongylus brasiliensis on intestinal epithelial cells in vitro. Parasite immunology. 2010;32:420–429. [PubMed] [Google Scholar]
  • Tecle E, Diaz-Balzac CA, Bulow HE. Distinct 3-O-sulfated heparan sulfate modification patterns are required for kal-1-dependent neurite branching in a context-dependent manner in Caenorhabditis elegans. G3 (Bethesda) 2013;3:541–552. [PMC free article] [PubMed] [Google Scholar]
  • Ten Dam GB, Kurup S, van de Westerlo EM, Versteeg EM, Lindahl U, Spillmann D, van Kuppevelt TH. 3-O-sulfated oligosaccharide structures are recognized by anti-heparan sulfate antibody HS4C3. J Biol Chem. 2006;281:4654–4562. [PubMed] [Google Scholar]
  • Thunberg L, Bäckström G, Lindahl U. Further characterization of the antithrombin-binding sequence in heparin. Carbohydr.Res. 1982;100:393–410. [PubMed] [Google Scholar]
  • Tiwari V, Clement C, Xu D, Valyi-Nagy T, Yue BY, Liu J, Shukla D. Role for 3-O-sulfated heparan sulfate as the receptor for herpes simplex virus type 1 entry into primary human corneal fibroblasts. J Virol. 2006;80:8970–8980. [PMC free article] [PubMed] [Google Scholar]
  • Tiwari V, O'Donnell CD, Oh MJ, Valyi-Nagy T, Shukla D. A role for 3-O-sulfotransferase isoform-4 in assisting HSV-1 entry and spread. Biochem Biophys Res Commun. 2005;338:930–937. [PubMed] [Google Scholar]
  • Vanpouille C, Deligny A, Delehedde M, Denys A, Melchior A, Lienard X, Lyon M, Mazurier J, Fernig DG, Allain F. The heparin/heparan sulfate sequence that interacts with cyclophilin B contains a 3-O-sulfated N-unsubstituted glucosamine residue. J Biol Chem. 2007;282:24416–24429. [PubMed] [Google Scholar]
  • Wiedermann Ch J, Romisch J. The anti-inflammatory actions of antithrombin--a review. Acta Med Austriaca. 2002;29:89–92. [PubMed] [Google Scholar]
  • Wood MW, Breitschwerdt EB, Gookin JL. Autocrine effects of interleukin-6 mediate acute-phase proinflammatory and tissue-reparative transcriptional responses of canine bladder mucosa. Infect Immun. 2011;79:708–715. [PMC free article] [PubMed] [Google Scholar]
  • Wu ZL, Lech M, Beeler DL, Rosenberg RD. Determining heparan sulfate structure in the vicinity of specific sulfotransferase recognition sites by mass spectrometry. J Biol Chem. 2004;279:1861–1866. [PubMed] [Google Scholar]
  • Xia GQ, Chen JH, Tiwari V, Ju WJ, Li JP, Malmström A, Shukla D, Liu J. Heparan sulfate 3-O-sulfotransferase isoform 5 generates both an antithrombin-binding site and an entry receptor for herpes simplex virus, type 1. J Biol Chem. 2002;277:37912–37919. [PubMed] [Google Scholar]
  • Xu D, Esko JD. Demystifying heparan sulfate-binding proteins. Annu Rev Biochem. 2014 in press. [PMC free article] [PubMed] [Google Scholar]
  • Xu D, Moon AF, Song D, Pedersen LC, Liu J. Engineering sulfotransferases to modify heparan sulfate. Nat Chem Biol. 2008;4:200–202. [PMC free article] [PubMed] [Google Scholar]
  • Xu D, Tiwari V, Xia G, Clement C, Shukla D, Liu J. Characterization of heparan sulphate 3-O-sulphotransferase isoform 6 and its role in assisting the entry of herpes simplex virus type 1. Biochem J. 2005;385:451–459. [PMC free article] [PubMed] [Google Scholar]
  • Xu Y, Masuko S, Takieddin M, Xu H, Liu R, Jing J, Mousa SA, Linhardt RJ, Liu J. Chemoenzymatic synthesis of homogeneous ultralow molecular weight heparins. Science. 2011;334:498–501. [PMC free article] [PubMed] [Google Scholar]
  • Xu Y, Pempe EH, Liu J. Chemoenzymatic Synthesis of Heparin Oligosaccharides with both Anti-factor Xa and Anti-factor IIa Activities. J Biol Chem. 2012;287:29054–29061. [PMC free article] [PubMed] [Google Scholar]
  • Yabe T, Hata T, He J, Maeda N. Developmental and regional expression of heparan sulfate sulfotransferase genes in the mouse brain. Glycobiology. 2005;15:982–993. [PubMed] [Google Scholar]
  • Yamada S, Yoshida K, Sugiura M, Sugahara K, Khoo KH, Morris HR, Dell A. Structural studies on the bacterial lyase-resistant tetrasaccharides derived from the antithrombin III-binding of porcine intestinal heparin. J Biol Chem. 1993;268:4780–4787. [PubMed] [Google Scholar]
  • Ye S, Luo YD, Lu WQ, Jones RB, Linhardt RJ, Capila I, Toida T, Kan M, Pelletier H, McKeehan WL. Structural basis for interaction of FGF-1, FGF-2, and FGF-7 with different heparan sulfate motifs. Biochemistry. 2001;40:14429–14439. [PubMed] [Google Scholar]
  • Yoon M, Zago A, Shukla D, Spear PG. Mutations in the N termini of herpes simplex virus type 1 and 2 gDs alter functional interactions with the entry/fusion receptors HVEM, Nectin-2, and 3-O-sulfated heparan sulfate but not with Nectin-1. J Virol. 2003;77:9221–9231. [PMC free article] [PubMed] [Google Scholar]
  • Zhang LJ, Beeler DL, Lawrence R, Lech M, Liu J, Davis JC, Shriver Z, Sasisekharan R, Rosenberg RD. 6-O-sulfotransferase-1 represents a critical enzyme in the anticoagulant heparan sulfate biosynthetic pathway. J Biol Chem. 2001a;276:42311–42321. [PubMed] [Google Scholar]
  • Zhang LJ, Lawrence R, Schwartz JJ, Bai XM, Wei G, Esko JD, Rosenberg RD. The effect of precursor structures on the action of glucosaminyl 3-O-sulfotransferase-1 and the biosynthesis of anticoagulant heparan sulfate. J Biol Chem. 2001b;276:28806–28813. [PubMed] [Google Scholar]
  • Zhang LJ, Schwartz JJ, Miller J, Liu J, Fritze LMS, Shworak NW, Rosenberg RD. The retinoic acid and cAMP-dependent up-regulation of 3-O-sulfotransferase-1 leads to a dramatic augmentation of anticoagulantly active heparan sulfate biosynthesis in F9 embryonal carcinoma cells. J Biol Chem. 1998;273:27998–28003. [PubMed] [Google Scholar]
  • Zhang LJ, Yoshida K, Liu J, Rosenberg RD. Anticoagulant heparan sulfate precursor structures in F9 embryonal carcinoma cells. J Biol Chem. 1999;274:5681–5691. [PubMed] [Google Scholar]
  • Zhang W, Swanson R, Izaguirre G, Xiong Y, Lau LF, Olson ST. The heparin-binding site of antithrombin is crucial for antiangiogenic activity. Blood. 2005;106:1621–1628. [PMC free article] [PubMed] [Google Scholar]
  • Zhang Z, Liu X, Chen J, Su H, Luo Q, Ye J, Tang N, Zhang W, Chen W, Ko BC, Huang A. Heparin sulphate D-glucosaminyl 3-O-sulfotransferase 3B1 plays a role in HBV replication. Virology. 2010;406:280–285. [PubMed] [Google Scholar]
  • Zhao W, Garron ML, Yang B, Xiao Z, Esko JD, Cygler M, Linhardt RJ. Asparagine 405 of heparin lyase II prevents the cleavage of glycosidic linkages proximate to a 3-O-sulfoglucosamine residue. FEBS Lett. 2011;585:2461–2466. [PMC free article] [PubMed] [Google Scholar]
-