the expression and regulation of hyaluronan synthases …165994/fulltext01.pdf · the expression...

63
Comprehensive Summaries of Uppsala Dissertations from the Faculty of Medicine 955 _____________________________ _____________________________ The Expression and Regulation of Hyaluronan Synthases and Their Role in Glycosaminoglycan Synthesis BY JONAS BRINCK ACTA UNIVERSITATIS UPSALIENSIS UPPSALA 2000

Upload: phungtuyen

Post on 06-Jun-2018

221 views

Category:

Documents


0 download

TRANSCRIPT

Comprehensive Summaries of Uppsala Dissertationsfrom the Faculty of Medicine 955

_____________________________ _____________________________

The Expression and Regulation of Hyaluronan Synthases and Their Role

in Glycosaminoglycan Synthesis

BY

JONAS BRINCK

ACTA UNIVERSITATIS UPSALIENSISUPPSALA 2000

2

Dissertation for the Degree of Doctor of Philosophy (Faculty of Medicine) in MedicalBiochemistry presented at Uppsala University in 2000

ABSTRACT

Brinck, J. 2000. The Expression and Regulation of Hyaluronan Synthases and Their Role inGlycosaminoglycan Synthesis. Acta Universitatis Upsaliensis. Comprehensive Summaries ofUppsala Dissertations from the Faculty of Medicine 955. 63 pp. Uppsala. ISBN 91-554-4806-2.

The glycosaminoglycan hyaluronan is an essential component of the extracellular matrix in allhigher organisms, affecting cellular processes such as migration, proliferation and differentiation.Hyaluronan is synthesized by a plasma membrane bound hyaluronan synthase (HAS) which existsin three genetic isoforms. This thesis focuses on the understanding of the hyaluronan biosynthesisby studies on the expression and regulation of the HAS proteins.

In order to characterize the structural and functional properties of the HAS isoforms wedeveloped a method to solubilize HAS protein(s) while retaining enzymatic activity. The partiallypurified HAS protein is, most likely, not asscociated covalently with other components. Cellstransfected with cDNAs for HAS1, HAS2 and HAS3 were studied and all three HAS isozymeswere able to synthesize high molecular weight hyaluronan chains in intact cells. The regulation ofthe hyaluronan chain length involves cell specific elements as well as external stimulatory factors.HAS3 transfected cells with high hyaluronan production exhibit reduced migration capacity andreduced amounts of a cell surface hyaluronan receptor molecule (CD44) compared to wild-typecells.

The three HAS isoforms were studied and shown to be differentially expressed and regulated inresponse to external stimuli. Platelet derived growth factor (PDGF-BB) and transforming growthfactor (TGF-ß1) are important regulators of HAS at both the transcriptional and translational level.The HAS2 isoform is the isoform most susceptible to external regulation.

The role of the UDP-glucose dehydrogenase in mammalian glycosaminoglycan biosynthesiswas assessed. The enzyme is essential for hyaluronan, heparan sulfate and chondroitin sulfatebiosynthesis, but does not exert a rate-limiting effect.

Key words: hyaluronan, glycosaminoglycans, CD44, growth factors, UDP-glucose dehydrogenase.

Jonas Brinck, Department of Medical Biochemistry and Microbiology, Biomedical Centre, Box582, SE-751 23 Uppsala, Sweden

© Jonas Brinck 2000

ISSN 0282-7476ISBN 91-554-4806-2

Printed in Sweden by Lindbergs Grafiska, Uppsala 2000

3

non scholae sed vitae discimusSeneca († 65)

4

This thesis is based on the following papers, which are referred to in the text by their Roman

numerals:

I Asplund, T., Brinck, J*., Suzuki, M., Briskin, M. and Heldin, P. (1998) Characterization

of hyaluronan synthase from a human glioma cell line. Biochim Biophys Acta. 1380:377-

388* The two first authors contributed equally to the work.

II Brinck, J. and Heldin, P. (1999) Expression of recombinant hyaluronan synthase (HAS)

isoforms in CHO cells reduces cell migration and cell surface CD44. Exp Cell Res.

252:342-351

III Jacobson, A., Brinck, J., Briskin, M., Spicer, A. and Heldin, P. (2000) Expression of

human hyaluronan synthases in response to external stimuli. Biochem J. 348:29-35

IV Brinck, J., Roman, E*., Heldin, P. and Kusche-Gullberg, M. (2000) The role of UDP-

glucose dehydrogenase in mammalian glycosaminoglycan biosynthesis. Manuscript.* The two first authors contributed equally to the work

Published manuscripts have been reprinted with permission from the respective journals.

5

TABLE OF CONTENTS

BACKGROUND………………………………………………………………… 9

1 The extracellular matrix – the social context of the cells…………………………….…. 9

2 Glycosaminoglycans……………………………………………………………….…… 9

2.1 Hyaluronan (HA)………………………………………………………….…… 10

2.1.1 Structure and properties…………………………………………………10

2.1.2 Tissue distribution……………………………………………….….… 10

2.1.3 Functions…………………………………………………………….… 11

2.2 Sulfated glycosaminoglycans……………………………………………………11

2.2.1 Structures and properties…………………………………………….… 11

2.2.2 Functions……………………………………………………………… 12

3 HA binding proteins…………………………………………………………………… 13

3.1 Link module containing proteins………………………………………………. 13

3.1.1 Hyalectans………………………………………………………………14

3.1.2 CD44……………………………………………………………………14

3.1.2.1 Structure and ligand binding…………………………………… 14

3.1.2.2 CD44 and HA in cell migration…………………………………16

3.1.3 LYVE-1………………………………………………………………… 17

3.2 BX7B containing proteins…………………………………………………….… 17

3.3 Novel HA binding proteins…………………………………………………..… 18

4 Biosynthesis of glycosaminoglycans………………………………………………….…18

4.1 UDP-sugars…………………………………………………………………… 19

4.1.1 Synthesis of UDP-sugar pools……………………………………….…19

4.1.2 Regulatory role………………………………………………………… 21

4.2 Biosynthesis of HA………………………………………………………………22

4.2.1 Hyaluronan synthase (HAS) gene family……………………………… 22

4.2.2 HAS protein………………………………………………………….… 23

4.2.2.1 Enzyme properties…………………………………………..… 23

4.2.2.2 Kinetics…………………………………………………..…… 25

4.2.3 Regulation of HA biosynthesis………………………………………… 26

4.2.3.1 Turnover of HAS protein……………………………………… 26

4.2.3.2 Effects of growth factors……………………………………… 27

4.2.3.3 Involvement of protein kinases…………………………..…… 28

4.2.4 Biosynthetic directionality……………………………………………… 28

6

4.3 Biosynthesis of sulfated glycosaminoglycans…………………………….…… 29

4.3.1 Linkage region………………………………………………………… 30

4.3.2 Polymerization and modification……………………………….……… 30

5 Degradation of HA………………………………………………………………………31

5.1 Tissue turnover………………………………………………………….……… 31

5.2 Hyaluronidases………………………………………………………….……… 31

PRESENT INVESTIGATION…………………………………………..……… 33

6 Aims of the study…………………………………………………………….………… 33

7 HAS enzyme can be solubilized in an active form (Paper I)………………….………… 34

7.1 Solubilization and purification………………………………………..………… 34

7.2 Antibodies towards HAS protein……………………………………….……… 35

8 Is HA synthesized from the reducing end? (Paper I)…………………………….………35

9 Expression of HAS affects cell behavior (Paper II)……………………………..……… 37

9.1 Synthesis and accumulation of HA……………………………………..……… 37

9.2 HA chain length………………………………………………………………… 38

9.3 Cell migration…………………………………………………………..……… 39

10 HAS isoforms show specificity in gene regulation (Paper III)………………………… 39

10.1 Effects of culture conditions…………………………………………..……… 39

10.2 Effects of external stimuli………………………………………………………40

11 The role of UDP-glucose dehydrogenase in mammalian glycosaminoglycan

biosynthesis (Paper IV) ……………………………………………………….……… 40

12 Conclusions…………………………………………………………………………… 42

13 Future perspectives……………………………………………………………….…… 43

ACKNOWLEDGEMENTS…………………………………………………….… 45

REFERENCES………………………………………………………….…….… 46

7

Abbreviations

aa amino acids

CHO Chinese hamster ovary (cells)

CS chondroitin sulfate

DS dermatan sulfate

EGF epidermal growth factor

GAG glycosaminoglycan

Gal D-galactose

GalNAc N-acetyl D-galactosamine

Glc D-glucose

GlcA D-glucuronic acid

GlcNAc N-acetyl D-glucosamine

GPI glycosylphospatidylinositol

HA hyaluronan

HAS hyaluronan synthase

HS heparan sulfate

IdoA L-iduronic acid

KS keratan sulfate

PDGF platelet derived growth factor

PMA phorbol 12-myristate 13-acetate

TGF transforming growth factor

TSG-6 protein product of tumor necrosis factor-stimulated gene-6

UDP uridine 5'-diphosphate

xlHAS-rs Xenapus laevis hyaluronan synthase related sequence

Xyl D-xylose

8

9

BACKGROUND

1 The extracellular matrix – the social context of the cellsIt is a well-known fact that the human body is composed of different kinds of cells carrying out the

specialized functions for which they have been designed. Cells with similiar or complementary

functions assemble into tissues and organs that constitute functional entities and make the biological

organism Homo sapiens “work”. One difference between the human body and single cell

organisms such as bacteria is the fact that the individual cells in the body exist in a self-created

social environment, in contrast to the bacteria where the cell wall constitutes the barrier towards the

exterior world. The biological name for this social habitat is the extracellular matrix.

The cells synthesize the components of the extracellular matrix to fit their needs, which

means that structure and function of the matrix is almost as specialized and unique as the cells

themselves. It is the extracellular matrix that gives the tissues their tensile strength or elasticity, and

serves as a scaffold for cell adhesion and cell movement. The components of the matrix also help to

transmit signals that regulate cellular functions and are able to enhance or diminish these signals.

The extracellular matrix consists of a wide variety of protein and sugar molecules entangled

in a complex network that is constantly turned over. Among these are the linear carbohydrate

polymers, glycosaminoglycans, with essential functions both in the matrix and attached to the

surfaces of cells. It is of great importance to understand their metabolism since imbalances or

disorders in this system eventually will lead to disease. The main topic of this thesis is to give new

insights to the synthesis and regulation of the glycosaminoglycan hyaluronan.

2 GlycosaminoglycansThe members of the glycosaminoglycan (GAG) family are essential components in all higher

organisms and are found both on the surfaces of cells and in the extracellular matrix. GAGs are

unbranched carbohydrate polymers consisting of repeating disaccharide units that can be subject to

extensive modifications, thereby rendering highly diverse polysaccharide chains. The individiual

members of the GAG family are defined according to their disaccharide structure and modifications.

10

2.1 Hyaluronan (HA)2.1.1 Structure and properties

The GAG hyaluronan (HA)(reviewed in (106)) was first isolated from the vitreous body of the eye

by Meyer and Palmer in the 1930’s (123). HA consists of repeating disaccharide units of [GlcAß1-

3GlcNAcß1-4]n, where n can be up to twentyfive-thousand. The contour length of an HA chain of

Mr 4 x 106 is 10 µm (44). The HA chain is polymerized by a plasma membrane bound enzyme and

is not subjected to any type of covalent modification during its synthesis.

Nuclear magnetic resonance studies of the shape of the HA chain by Scott (165) have

shown the existence of internal hydrogen bounds that stabilize the chain in a stiffened helical

conformation. HA chains have the capacity to self-aggregate in aqueous solutions which can be

visualized in electron microscopy of rotary-shadowed preparations of HA (166). The aggregation of

two anti-parallel HA molecules is promoted by hydrogen bonds between the acetoamido group on

one chain and the carboxylate on the other (167). Preparations of high molecular weight HA in

concentrations of 5-10 mg/ml are therefore viscous.

Figure 1. An HA pericellular matrix excluding fixed red blood cells from the plasma membrane of a

Chinese hamster ovary cell.

2.1.2 Tissue distribution

HA is synthesized by almost all members of the animal kingdom as well as by certain bacteria and

algae viruses (106)(26). HA is found mainly in the extracellular space where it accumulates, but the

polymer can also be bound to the cell surface or be located intracellularly around in the nucleus and

in the lysosomes (153)(36)(108). The largest storage of HA in the human body is in the skin,

constituting about 50 % of the total HA supply (152).

11

Table 1 HA concentrations in various human organs and fluids

Organ or fluid Concentration (µg/g)

aqueous humour 0.3-2.2

brain 35-115

dermis 200

plasma (serum) 0.01-0.1

synovial fluid 1400-3600

thoracic lymph 8.5-18

umbilical cord 4100

urine 0.1-0.3

vitreous body 140-340

Data taken from (48).

2.1.3 Function

High molecular weight HA have been ascribed an essential structural role in the extracellular matrix

of the connective tissue where the polymers create a mesh-like structure (106)(166). The viscous

hydrated HA gel provides resistance to compressive forces and acts as a biological lubricant. During

embryonic development the deposite of HA creates a cell free area by expansion of the extracellular

space which facilitates cell migration (202). Besides being a structural element, HA has been

reported to be involved in several specific processes such as lympocyte extravasation (32), cancer

metastasis (171) and angiogenesis (158) probably through interactions with HA binding proteins.

Since the HA polymer itself does not exhibit any structural diversity, its function is in part

due to the chain length. The inductive role of HA in angiogenesis could be ascribed to HA

oligosaccharides (4-25 disaccharides)(221) whereas high molecular weight HA exerted an

inhibitory effect (43). HA fragments have also been reported to evoke an inflammatory response in

macrophages (121)(80).

2.2 Sulfated glycosaminoglycans2.2.1 Structure and properties

The sulfated GAGs can be subdivided into three groups based on the disaccharide composition:

heparan sulfate (HS)/heparin, chondroitin sulfate (CS)/dermatan sulfate (DS) and keratan sulfate

(KS) (Table 2)(96). During their synthesis in the Golgi, the chains are heavily modified by

12

deacetylation, C5-epimerization, N- and O-sulfations generating a high degree of diversity (see

section 4.3). The chain lengths vary between 5-70 kDa.

Table 2 The structure of the sulfated GAGs

GAG disaccharide unit modifications

heparan sulfate (HS)/

heparin

[GlcA/IdoAß/α1-4GlcNAcα1-4] N-deacetylation, N-sulfation

C5-epimerization of GlcA

C2-sulfation on GlcA/IdoA

C3,C6-sulfation on GlcNAc

chondroitin sulfate (CS)/

dermatan sulfate (DS)

[GlcA/IdoAß/α1-3GalNAcß1-4] C5-epimerization of GlcA

C2-sulfation on GlcA/IdoA

C4,C6-sulfation on GlcNAc

keratan sulfate (KS) [Galß1-4GlcNAcß1-3] C6-sulfation on Gal/GlcNAc

The sulfated GAG chains appear as proteolglycans, i.e. covalently linked to proteins, where the

protein core can be of a different nature: transmembrane, glycosylphosphatidylinositol (GPI)-

anchored or soluble. A proteoglycan can have several GAG chains of different types attached to it as

exemplified by aggrecan, which consists of a 200 kDa protein core and about 130 CS and KS

chains, with a total Mr of about 3 x 106 (66).

The expression pattern of proteoglycans is specific; heparin is synthesized only by

connective tissue mast cells in the form of the proteoglycan serglycin whereas HS is found on

several types of proteoglycans on virtually all cells in the body (96). The structural diversity of the

GAG chains is also precisely regulated as shown by recent studies with monoclonal antibodies

directed against structurally different epitopes of HS (212)(90). Different structures of the kidney

and muscle basal lamina were investigated and shown to contain a subset of different HS species.

2.2.2 Functions

The proteoglycan superfamily consists of over 30 members with a wide variety of biological

functions (82). They play an important role in the extracellular matrix organization, influence cell

growth and tissue maturation, and participate in the regulation of matrix turn-over by the binding

and inactivation of protease inhibitors. Proteoglycans act also as biological sieves by being

structural components in the glomerular basal laminae.

Many of the specific functions of the proteoglycans are due to the highly modified diverse

saccharide sequences of the GAG chains. For example, the anticoagulant activity of heparin has

13

been known for a long time (77). A unique pentameric structure present only in a fraction of the

total heparin has the capacity to interact with antithrombin-III, thus forming a potent inhibitor of

thrombin and thereby blood coagulation (12). Furthermore, proteoglycans are important modulators

of growth factor activities (159). The binding of growth factors to the GAGs protect ligands from

proteolytic degradation and GAGs have therefore been suggested to act as a growth factor storage

facility. HS has been shown to be required for binding of basic fibroblast growth factor to its

receptor and the growth factor mediated effects (228)(149).

3 HA binding proteinsHA is unique in the GAG family by lacking covalent linkages to proteins thus not being able to

form proteoglycans. The only exception to this rule is the ester linkage between the C-terminal

aspartic acid of the heavy chains of pre-α-trypsin inhibitor/inter-α-trypsin inhibitor and the C6-

hydroxyl group of an internal GlcNAc residue in the HA chain (234), although the functional role

of this linkage is not clear.

The fact that HA can form several specific interactions with proteins, serves as indirect evidence that

HA is more than a structural component of the extracellular matrix. These HA binding proteins are

mainly found extracellularly, but can also be bound to the cell surface or be located intracellularly.

New HA binding proteins are currently under characterization and the list is likely to get longer.

Below is some of the best characterized HA binding proteins presented, subdivided according to

their HA binding motif.

3.1 Link module containing proteinsThe Link module, about 100 amino acids (aa), is the structural HA binding motif shared by the Link

module superfamily, which comprise extracellular matrix molecules (link protein, hyaluronectin,

versican, aggrecan, neurocan, brevican, TSG-6) and the cell surface receptors (CD44 and LYVE-1)

(25).The three-dimensional structure of the human TSG-6 Link module, expressed in Escherichia

coli, has been resolved by nuclear magnetic resonance spectroscopy (100). It comprises two α-

helices and two triple-stranded anti-parallel ß-sheets arranged around a large hydrophobic core,

identical to the fold of the C-type lectin domain. The recombinant expressed Link module of TSG-6

has the capacity to bind HA (100), compared to CD44 which requires additional 60 aa of the

receptor in order to acquire the correct fold and HA binding capacity (6). The Link module of TSG-

6 can also bind CS at a position which is overlapping with the HA binding surface (136).

14

3.1.1 Hyalectans

The hyalectans are a family of proteoglycans interacting with HA and lectins and have currently four

members: versican, aggrecan, neurocan and brevican (83). They share a common tridomain structure

consisting of an N-terminus that binds HA, a central domain that carries GAG chains and a C-

terminal that binds to lectins (82).

Versican is the largest member of the hyalectan family with the core protein of 265-370 kDa

which is modified by 10-30 CS/DS chains. Versican binds HA with a Kd of about 4 nM (109) but

can also bind simple monosaccharide sugars as well as HS and heparin (209). The versican gene

can be upregulated by a variety of growth factors such as platelet derived growth factor (PDGF),

epidermal growth factor (EGF) and tumor necrosis factor (TNF) (195). Versican has been

implicated in regulation of neural crest cell migration by exerting a barrier role for the axonal

outgrowth (103).

Aggrecan consists of a core protein of about 220 kDa modified by CS and KS chains (up

to 130). Aggrecan secreted by chondrocytes aggregates extracellularly but can also bind to the cell

surface (94)(176). Aggrecan, link protein and HA forms a ternary complex (66) which, inside

cartilage, occupies a large hydrodynamic volume making the tissue elastic to compressive forces.

The affinity of the aggrecan link complex for HA is quite high (Kd≤1 nM).

Neurocan (protein core 140 kDa) was first cloned from rat brain (150) and is synthesized

by neurons, and probably astrocytes, in the brain (218). Neurocan binds neural adhesion molecules

Ng-CAM and N-CAM with high affinity and can thereby inhibit their homophilic interactions and

block neurite outgrowth (58)(51). Neurocan is developmentally regulated, and its interactions with

its ligands may be confined to restricted areas and a relatively brief stage in development (126).

Brevican is the smallest member of the hyalectans (protein core 100 kDa). The molecule can

be synthesized in a soluble extracellular form or as a GPI-anchored protein (168), which localizes to

the plasma membrane. This latter splice variant is deprived of its lectin binding domain, but still has

the HA binding domain and sites for glycosylation. Brevican is one of the most abundant CS

containing proteoglycans in the adult brain and its expression is highly specific (224). Brevican has

been reported to inhibit neurite outgrowth in vitro thus maybe controlling infiltration of axons and

dendrites in vivo (222).

3.1.2 CD44

3.1.2.1 Structure and ligand binding

CD44 is the best characterized HA receptor to date and its structure and function have been covered

in several excellent reviews (113)(95)(112). Besides HA, the CD44 receptor can bind a number of

other extracellular matrix components such as collagen I (42), fibronectin (89), HS, CS and

heparin (174). CD44 is widely expressed, and can be found on most hematopoetic cells,

15

keratinocytes, chondrocytes, many epithelial cell types and some endothelial and neural cells (112).

The many functions attributed to the receptor include T-cell extravasation (32), myelopoeisis (129)

and tumor migration (59)(237), although the mechanisms behind this are not always clear.

CD44 is a type-1 transmembrane glycoprotein and the standard isoform (CD44H; originally

found on hematopoetic cells (23)) consists of ~270 aa. The amino terminal part (~180 aa) of the

extracellular domain is characterized as the conserved region and exhibits 85 % homology between

mammalian species. This region harbors the Link module, five N-glycosylation sites and a BX7B

motif (see section 3.2). The proximal part of the extracellular domain (non-conserved region) is

characterized by alternative splicing of at least ten variant exons of the CD44 gene which allows

insertion of extra aa (87)(200)(237). The alternative splicing occurs only in particular cell types and

under certain conditions and is therefore considered to be precisely regulated. This proximal region

also contains O-glycosylation sites and sites for attachment of GAG chains. The CD44 protein,

including the splice variants, have been shown to be decorated with KS, HS and CS chains

(14)(72)(188). The transmembrane domain (21 aa) is 100 % conserved between mammalian

species. The cytoplasmic domain (~72 aa) seems to be able to interact with several intracellular

proteins such as ankyrin (91) and the ERM-protein family (ezrin, radixin and moesin) (206)(160)

thus potentially creating a linkage between the receptor and the actin cytoskeleton. The cytoplasmic

domain can also be phosphorylated at several tyrosine and serine residues but only two serine

residues are phosphorylated in vivo (130).

CD44 is the principle cell surface receptor for HA (1) by binding of HA to the Link module

(100)(4); the contribution of the BX7B motif (225) to HA binding is probably minor (112). The

smallest HA molecule that can bind to CD44 is a hexasaccharide to decasaccharide (210)(2)(189).

However, the existence of CD44 molecules on the cell surface does not automatically convey HA

binding but depends on the ”activation state” of the cell (113)(95). Three CD44 activation states

have been described: 1) CD44+ cells that bind HA constitutively, 2) CD44+ cells that can be induced

(by monoclonal antibodies or cytokines) to bind HA and 3) CD44+ cells that do not bind HA at all.

A lot of work has been put in to clarify what regulates the binding of HA to CD44. N-

glycosylation of CD44 seems to inhibit HA binding, since active HA binding cells have less N-

glycosylation than non-binding cells and treatment of these non-binding cells with N-glycosylation

inhibitors rendered them active (64)(92). O-glycosylation of CD44 has also been reported to have a

negative effect on HA binding in some cases (24) but not in all (235). Modification of the receptor

by sulfation leads to increased HA binding (118).

The transmembrane and intracellular domain of CD44 contribute to HA binding. Evidence

that the transmembrane domain is involved in dimerization of the receptor by disulfide bonding

(116) and palmitoylation (11) has been presented, and a correlation between ligand binding capacity

and dimerization has been shown (173). The cytoplasmic part of CD44 is important for ligand

16

binding, since cells expressing this truncated form failed to bind HA if they were not first exposed

to CD44-specific antibodies (111). The cytoplasmic part is also involved in heterogeneous

localization and polarization of receptors on the cell surface in pre-B cells and epithelial cells

(131)(216). Taken together, this strengthens the role of receptor-clustering as a positive factor for

HA binding. Phosphorylation of the intracellular domain of CD44 does however not seem to

promote or reglulate ligand binding (138)(208).

3.1.2.2 CD44 and HA in cell migration

Cell migration and cell locomotion involves a series of complex interactions between cytoskeleton,

cell surface receptors and matrix components (105)(170). Several examples from the field of

embryonic development implicate HA and CD44 as positive elements in cell migration. The

migration of cells into the developing chick embryo cornea and heart coincided temporally and

spatially with the synthesis of HA and the migration decreased as the hyaluronidase activity in the

tissues increased (204)(135). During the development of the soft palate, Spicer and co-workers have

shown with studies of HAS knock-out mice that the cells of the cranio-facial mesenchyme are

dependent on HA for migration and normal organization of the tissue (178). The embryonic

development of the cardiac septum and valves were also dependent on HA as shown in HAS2

knock-out mice (16); when the cardiac valve cells were cut out and grown on collagen gels they

could not migrate unless they were rescued with HAS2 expression or addition of exogenous HA. In

another study, Thomas et al. (194) showed that melanoma cells enhanced their migration rates on

HA coated dishes after transfection with CD44 compared to non-transfected cells, while antibodies

against the receptor inhibited migration. Phosphorylation of the intracellular serine residues of

CD44 is important for the CD44 mediated migratory effect, since melanoma cells expressing

phosphorylation mutants of CD44 were not able to migrate (138).

Attempts to explain the mechanism of the CD44 – HA interaction in cell migration have

been made. The deposit of HA in the extracellular matrix leads to an overall expansion of the tissue

(202)(122), thereby creating hydrated routes that facilitates cell migration according to Toole (202).

For example, HA indirectly promoted glioblastoma cell migration in 3-dimensional fibrin gels by

modulation of the fibrin fiber architecture thus increasing the pore sizes in the gels (65). In such a

model, CD44 could be the link between the cell’s actin cytoskeleton and the scaffold that the HA-

gel provide. The mechanism of migration might also comprise a turnover of the extracellular

components and cell surface receptors (105). Fibroblasts, macrophages and chondrocytes have the

capacity to internalize and degrade HA by a CD44 mediated endocytosis (22)(81), although the

amount of HA degraded by the cells of mesenchymal origin seems to be limited (71).

Knudson and Knudson (97) have presented an hypothesis where the assembly of HA

containing pericellular matrices around cells (first visualized by Claris and Fraser (18)) prevent

17

them from forming necessary interactions for cell migration with a decreased migration rate as an

effect. In vitro these matrices consist either of newly synthesized HA chains, still attached to the

HAS protein (69), or of exogeneously added HA which is bound to HA receptors (CD44)(99)(98).

The matrix formation can be promoted by HA binding proteoglycans, which also increase their

density and stability (70). Experimental evidence supporting this hypothesis has been presented

(140)(141) where neural crest cell migration was inhibited by HA matrix promoting chondroitin

sulfate proteoglycans (i.e. aggrecan).

3.1.3 LYVE-1

LYVE-1 is a type-1 transmembrane protein and a new member of the Link module superfamily

being a homologue of the CD44 receptor with an overall similarity to CD44 of 41 % (7). Of the

nine key amino acids that have been shown to be important in HA binding to CD44 (137)(4), only

three are conserved in LYVE-1 (7). The receptor is almost exclusively expressed in lymph vessels

where it co-localizes with HA on the luminal side. LYVE-1 is not, unlike CD44, expressed in blood

vessels. Since the major part of the HA turnover and clearance takes place via the lymphatic system

(see section 5.1) the receptor could be an important regulator of these events.

3.2 BX7B containing proteinsThe α-helical peptide sequence BX7B, where B is a basic aa and X is any non-negatively charged aa,

have been shown to bind HA (226)(225). At least three proteins (RHAMM, IHABP and CD34)

have such a functional HA binding motif, although it exists in several other proteins, including

members of the Link module superfamily. It is not known if the BX7B sequence can confer HA

binding capacity in these contexts.

RHAMM (acronym for Receptor for Hyaluronan-Mediated Motility) was first characterized

by Turley in 1992 from a 3T3 cDNA expression library (63) and contains two adjacent BX7B

motifs which contribute equally to HA binding. The 52 kDa RHAMM protein proved to be similar

to two of the proteins (52 kDa and 58 kDa) previously reported to be members of the Hyaluronan

Receptor Complex (HARC) involved in ras-transformed cell locomotion (207). However, the

cDNA clone isolated (63) was not of full length and additional exons, including splice variants, have

been described (14 exons in total)(40)(217) coding for a protein of 70-73 kDa.

The RHAMM receptor was first reported to be cell surface associated/extracellular (63) but

recent studies revealed that it also exists in the cytoplasm and in the nucleus (233)(39). RHAMM

has been implicated as a regulator of several important cell functions. Overexpression of the

RHAMMv4 (splice variant 4) in murine fibroblasts is transforming and causes spontaneous

metastases in the lung (60). RHAMM has been postulated to act down-stream of Ras on the MAP

kinase - ERK (extracellular-regulated kinase) signaling pathway (233). There is also a correlation

18

between RHAMM overexpression and overexpression of MAP kinase, which could serve as a

prognostic indicator of breast carcinoma progression (217).

In 1998 Hofmann et al. (76) identified IHABP (acronym for Intracellular Hyaluronic Acid

Binding Protein) which has over 90 % similarity to the RHAMM protein and which shares the HA

binding BX7B motifs found in RHAMM. Fieber et al. (45) could later show that RHAMM is a N-

truncated form of IHABP and that the full length RHAMM/IHABP gene encodes 18 exons in total.

IHABP is expressed as a 95 kDa intracellular protein in a wide variety of tissues. The function of

IHABP remains to be elucidated but the intracellular localization gives potentially new biological

roles for HA. The physiological role of RHAMM is under evaluation in the light of the new data

about IHABP (75).

3.3 Novel HA binding proteinsThe liver is important for HA turnover by the uptake and degradation of the HA in liver endothelial

cells (see section 5). The putative receptor(s) responsible for receptor-mediated endocytosis of HA

on these cells have been described but not cloned (107)(120)(227)(236). The receptor reported by

McCourt et al. (120) is possibly a part of a protein complex of three (related) proteins (two 170-

180 kDa and one 270 kDa proteins) and shows functional similarities to the scavenger receptor

family. Polyclonal antisera towards the 270 kDa protein blocked HA binding and degradation in

liver endothelial cells. The receptor reported by Zhou et al. (236) indicate a 300 kDa protein

complex as the functional unit for the receptor. Complete cloning and sequencing remains to be

done for both of these receptors to determine if they are related and to elucidate the molecular basis

of their HA binding.

Recently, three putative HA binding proteins have been identified on cDNA level by search

in an expressed sequence tag (EST) database (205). Based on the original tissues from which these

cDNAs were derived, namely white fat, bone marrow and osteoblast, they have received the names

WF-HABP, BM-HABP and OE-HABP, respectively. The functional roles of these molecules as

HA binding proteins remain to be assessed.

4 Biosynthesis of glycosaminoglycansThe biosynthesis of the members of the GAG family is done by partly unique pathways and involve

interactions between enzymes in different subcellular compartments. In a first common step, high

energy nucleotide sugar donors are synthesized in the cytoplasm. These UDP-sugars can be

directly utilized by the plasma membrane bound hyaluronan synthases (HAS) for HA

polymerization. The sulfated GAGs are synthesized on a protein core as proteoglycans in the Golgi

and the UDP-sugars must therefore first be transported into the Golgi compartments by an

19

antiporter system. The Golgi biosynthetic machinery include chain polymerization and various

modifications which, in concert, generate unique chain structures.

4.1 UDP-sugars4.1.1 Synthesis of UDP-sugar pools

The main part of the cell’s glycosylation reactions occur in the Golgi apparatus, which includes N-

and O-glycosylation of proteins and lipids, synthesis of ceramide linked oligosaccharides and

glycophospholipid anchors and the polymerization of GAG chains (214). A common feature for

these reactions is the transfer of monosaccharide units to proteins/lipids from high-energy

nucleotide sugar donors, UDP-sugars, by the action of specific glycosyltransferases.

The biosynthetic pathways for UDP-sugar have been studied and are now established (50).

Practically all of them occur in the cytoplasm of the cell and include unique pathways and

interconversion of UDP-sugars from other pathways (Figure 2). The starting material is often the

cytoplasmatic glucose pool, but monosaccharide units salvaged from lysosomal degradation of

glycoproteins and GAGs (117) can also act as precursors. Rome and Hill (157) have shown that

hexosamines in the lysosomes can be reutilized by over 50 % after being transported back into the

cytoplasm by specific transporters. Thus the transporters may give an important contribution to the

UDP-sugar synthesis in the cell.

UDP-GalNAcUDP-GalNAc-1-PGalNAc UDP-GlcNAc

Fru-6-P

GlcN-6-P

GlcNAc-6-P

GlcNAc-1-P

UDP-Glc

UDP-GlcA

UDP-Xylose

UDP-Gal

Gal

Gal-1-P

Glc

Glc-6-P

Glc-1-P GlcN

Figure 2. Schematic view of UDP-sugar biosynthesis.

Adapted from (50).

20

The cellular concentrations of UDP-sugars have been estimated by Phelps and co-workers in bovine

corneal epithelium, endothelium (62) and stroma (61), and in sheep cartilage (55) (Table 3). It must

be emphasized that these are estimates of the total cellular pools and do not specify anything about

concentrations in various subcompartments. The turnover of these UDP-sugar pools is rapid.

Studying freshly isolated chondrosarcoma cells in culture Sweeney et al. (187) could show the half-

life of UDP-Glc/Gal and UDP-GlcA was 12 min and for UDP-Glc/GalNAc was 50 min. The

cellular UDP-sugars pools seem also to be exchangeable since the ratios of UDP-Glc/UDP-Gal

and UDP-GlcNAc/UDP-GalNAc (62)(187) are similar to the free equilibrium ratios reported for

the corresponding C4-epimerases (56).

Table 3 Cellular concentrations of UDP-sugars in cornea tissues and nasal cartilage

Cellular concentration (µM)

UDP-sugar epithelium endothelium stroma cartilage

UDP-Gal 115 28 85 24

UDP-GalNAc 138 82 95

UDP-Glc 279 87 364 73

UDP-GlcA 28

UDP-GlcNAc 368 172 211

UDP-Xyl 85 9 72 7

Data taken from (62)(61)(55).

The UDP-glucose dehydrogenase (UDPGDH) (reviewed in (139)) is a key enzyme in GAG

precursor biosynthesis. The enzyme converts UDP-Glc to UDP-GlcA, by the oxidation of C6

hydroxyl group with concomitant reduction of NAD+ to NADH. GlcA and its C5-epimer, IdoA,

constitutes every other residue in HA, HS/heparin and CS/DS. Also, the only way to synthesize

UDP-Xyl, necessary in the linkage region of GAGs (see section 4.3), is by decarboxylation of

UDP-GlcA. This reaction occurs both in the ER/early Golgi and in the cytoplasm (215)(93). The

UDP-GlcA is also needed as substrate in the glucuronidation of bile salts and detoxification of

xenobiotics in the liver ER (88). The functional unit of UDPGDH is a cytoplasmic complex of six

subunits, “trimers of dimers” (139). The Km-values for the UDPGDH from various vertebrate

tissues are 90-150 µM for NAD+ and 10-50 µM for UDP-glucose (55)(9). Kinetic studies in

corneal stroma and epithelium show that UDP-Xyl can act as an allosteric inhibitor for UDPGDH

(55)(132).

21

To be utilized for the biosynthesis of the sulfated GAGs in the Golgi, UDP-sugars synthesized in

the cytoplasma have to overcome their subcellular mislocation. The transport into the Golgi is

mediated through a group of hydrophobic transmembrane proteins and has been extensively studied

by Hirschberg (reviewed in (74)(73)). The proteins are antiporters, meaning that for each nucleotide

sugar transported into the Golgi, the corresponding nucleoside monophosphate must be transported

back to the cytoplasm. Each nucleotide sugar has its specific transporter and there is little, if any,

non-specificity in the system. The transport does not require ATP but is thought to be driven by the

down-concentration for the nucleoside monophosphate being transported out from the lumen into

the cytosplasm.

Studies on reconstituted antiporters in proteoliposomes have revealed useful information

regarding transport kinetics (127)(128). The Km-values for the UDP-Gal, UDP-Xyl and UDP-GlcA

transporters were determined to be between 2-5 µM which was very close to the Km-values for intact

Golgi vesicles. Preloading the proteoliposomes with UMP, a putative antiporter for the nucleotide

sugars, increased the nucleotide sugar transport by 2-3-fold. Thus the antiporters have the potential

to concentrate the nucleotide sugars between 50- to 100-fold relative to the incubation medium (73).

4.1.2 Regulatory role

The question of whether UDP-sugars exert any regulatory effect on mammalian GAG biosynthesis

is important, but there are few definite answers up to date. Balduini et al. (5) could show a decrease

CS synthesis in cornea when UDP-Xyl was added in vitro. They concluded that the level of

regulation was at the UDPGDH enzyme, presumably due to the negative allosteric effect of UDP-

Xyl. The turn-over of the UDP-sugars is rapid as reported by Mason and co-workers (187). After

stimulation of chondrosarcoma cells with serum or insulin the [35S]sulfate and [3H]leucin

incorporation into macromolecules increased 2-3-fold although the UDP-sugar pool sizes remained

constant. In another study (230), articular explant cultures could increase their UDP-GlcA pool

upon stimulation with serum concomitantly with increased production of GAGs. Thus the increased

demand for UDP-sugars in cells can be met by an increased fluctuation of the pools or by an

expansion of the pool size. They concluded that the UDP-sugars pools are unlikely to become rate-

limiting in GAG biosynthesis (187).

A Madin-Darby canine kidney (MDCK) cell line, 98 % deficient in UDP-Gal transport into

the Golgi lumen, had marked reduced galactosylation of glycoproteins and glycosphingolipids (13)

and reduced amounts of KS (the only GAG containing Gal in the polymer) (201). However, the

biosynthesis of HS and CS in these cells was not affected although Gal is needed in the linkage

region between the core protein and the GAG polymer (201). The authors speculated that the Km-

values for the glycosyltransferases involved in the transfer of Gal in the linkage region are lower

22

than that of the transferases involved in polymerization of KS, thereby explaining the different

effects on KS and HS/CS, although they could not rule out the possibility of subcompartementation

of the two processes.

Several reports, including studies of Km-values of the nucleotide antiporters in

proteolysomes, have suggested that the rate-limiting step of the Golgi located posttranslational

modifications (e.g. glycosylation) is the transport of UDP-sugars into the Golgi by the antiporter

system and not the glycosyltransferase activities (73).

4.2 Biosynthesis of HA4.2.1 Hyaluronan synthase (HAS) gene family

In 1992 van de Rijn and co-workers identified a gene responsible for HA production in group A

Streptococcus (33)(211). It soon became clear that the synthase was part of a hyaluronan synthase

(HAS) operon consisting of three members and that all encapsulated Group A Streptococci

possessed this operon (20). The gene was characterized (31)(35) and received the acronym hasA.

The hasB gene was characterized as a UDPGDH (34), the enzyme that converts UDP-Glc to UDP-

GlcA. The hasC gene encodes a UDP-glucose pyrophosphorylase responsible for the production of

UDP-Glc from glucose 1-phosphate and UTP (19).

The vertebrate HAS proteins belong to a gene family with four members (180). The first

three isoforms termed HAS1, HAS2 and HAS3 were cloned from human and mouse origins in a

short period of time between 1996 and 1998 (85)(172)(219)(177)(182). The fourth member, termed

HAS-related sequence (xlHAS-rs), was identified in Xenopus leavis (African green frog) with no

mammalian orthologue found so far (180). The first HAS gene was actually submitted to the gene

bank already in 1983 (164) but under the name DG42 (Differentially expressed at Gastrulation), a

protein found to be predominantly expressed during X. laevis gastrulation. The function of the

protein was at that point in time unknown. Later, a controversy about its catalytic properties arose

(213) which now has been resolved (see section 4.2.2.1).

The four members of the vertebrate gene family have probably arisen from a common

ancestral gene (180) after three gene duplications with subsequent divergence. Data that support this

hypothesis are common intron-exon boundaries and high sequence similarities between different

orthologues (Figure 3). The HAS gene family seems to have developed early in evolution since the

three first HAS genes are located on three separate autosomes (183). The tight coupling between the

HAS protein and the precursor producing enzymes seen in bacteria is not seen in the vertebrates; the

UDPGDH gene is located on a separate chromosome from the HAS genes (179).

The identification of a HAS gene in 1997 in the Paramecium bursaria Chlorella virus

(PBCV-1) genome (30) was a big surprise. When infecting chlorella-like green algae, a HAS

enzyme is translated from the open reading frame A98R. The origin of the viral HAS gene is still an

23

enigma. The enzyme shows about 30 % similarity on aa level with both vertebrate and bacterial HAS

with a slight predominance for the vertebrate protein. However, the viral genome also carries two

genes for HA precursors synthesis; a UDPGDH and a glutamine:fructose-6-phosphate

amidotransferase for synthesizing UDP-GlcA and glucosamine, respectively (104), which show a

higher degree of homology with their prokaryotic counterparts than their vertebrate ones (26).

However, the three genes do not function as an operon in the algae as they do in bacteria (30).

HAS1

HAS2

HAS3

xl HAS-rs 47

41

53

44

52

58

44

53

67

46

46

44

xl HAS-rs HAS1 HAS2 HAS3

Amino acid sequence similarity (%)

DNA sequence similarity (%)

xl HAS-rs

HAS1

HAS2

HAS3

ancestralHAS gene

Figure 3. The evolution of the vertebrate HAS gene family and the sequence similarities

between its members.

Data taken from (180).

The Gram-negative bacteria Pasteurella multocida is able to synthesize an HA capsule. Utilizing

transposon insertional mutagenesis a new type of HAS enzyme was cloned in the bacterial strain by

DeAngelis (29). This enzyme does not show any major resemblance to the other HAS genes,

deviating by size, predicted membrane topology and polymerization mechanism. DeAngelis has

suggested a classification of the HAS enzymes based on these characteristics (26). As of now, all

the HAS proteins cloned including vertebrate, prokaryotic and viral HAS belong to class I, with the

P. multocida HAS as the only member of class II.

4.2.2 HAS protein

4.2.2.1 Enzyme properties

The structure and topology of the HAS protein is to a large extent unknown. Philipson and

Schwartz could, with rigid experimental data, show that the vertebrate enzyme co-purified with the

plasma membrane fraction of mouse oligodendroglioma cells when fractionated by centrifugation

24

(142). Previously, Dorfmann and co-workers had shown that the bacterial HAS resides in the

membrane in group A Streptococci (119).

The size of the vertebrate HAS proteins is about 65 kDa, deduced from their aa sequences.

In a review of the HAS enzymes (220) a seven transmembrane or membrane associated domain

topology of the enzyme was suggested based on hydrophobicity plots of the aa sequence

(Figure 4); only a very small portion of the polypeptide chain is exposed to the outer surface in this

model. The structure has partly been supported by experimental data putting both the N- and C-

terminus in the cytosol (178). A basic aa cluster in the C-terminal part of the enzyme has been

proposed to participate in the regulation of the shedding of the negatively charged HA chain from

the cell surface. The large intracellular loop has motifs that are highly homologous between different

HAS members (180)(26) indicating this as the catalytic domain. Recently, site-directed mutagenesis

of HAS1 revealed potential sites for ß1-3GlcA and ß1-4GlcNAc transferase activities in the central

loop (229). The suggested topology is shared by both the prokaryotic and eukaryotic members of

the class I HAS family (220) with the difference that the prokaryotic proteins are shorter by a lack

of two C-terminal membrane spanning domains. This leaves only the P. multocida HAS (class II)

with a different structure (30) which probably only has two transmembrane domains or is soluble

(26).

NH2

COOH

Cytoplasm

Extracellular matrix

Figure 4. Suggested topology of the HAS enzyme (class I).

Adapted from (220). Star and circles show positions of aa important in ß1-3GlcA and ß1-4GlcNAc

transferase activities, respectively. Box depicts basic aa cluster suggested to be involved in shedding

of the HA chain from cell surface.

The long lasting enigma about the enzymatic activity of the DG42 protein, now termed Xenopus

laevis HAS1, seems to be resolved. In 1996 Semino et al. (169) claimed that DG42 synthesizes

25

chitin oligomers [GlcNAcß1-4GlcNAc]n n=4-5, while Meyer and Kreil (124) showed HA

polymerization; the dispute was commented on by Ajit Varki (213)). Kimata and co-workers have

recently shown that both activities reside in the purified enzyme when assayed in vitro (229).

The structure and topology of the HAS enzyme has been studied in greater detail in the

bacterial system. Weigel and co-workers have suggested that the HAS protein forms a pore in the

plasma membrane through which the polymer can be translocated concomitantly with chain

elongation (199)(197). The association of about 16 cardiolipin molecules, which induced enzyme

activity in vitro, seems to assist in the formation of a pore like structure. Other lipids have been

shown to enhance enzymatic activity, while others completely abolish it. (197). Less is known about

the vertebrate HAS enzymes and their lipid dependence and association. Several laboratories,

including ours, have solubilized HAS protein from different cell lines with sustained activity using

(preferably) digitonin as detergent (Paper I)(133). Kimata and co-workers have recently shown that

the detergent CHAPS was able to restore activity and kinetics of solubilized HAS protein to that of

the membrane bound enzyme (229).

To synthesize HA the following enzymatic activities are needed: ß1-3GlcA transferase, ß1-

4GlcNAc transferase and HA translocation (through the membrane). Does the HAS protein display

all of these activities by itself or is it aided by an accessory component? Three reports give rigid

evidence supporting the one polypeptide hypothesis. Weigel and co-workers (199) could show that

the size of the functional Streptococcal HAS complex expressed in E. coli was the same as the HAS

protein itself in irradiation studies, thereby excluding additional proteins. DeAngelis (28)

transformed the yeast strain Saccharomyces cerevisiae with the cDNA for Xenopus laevis HAS1

which gave the yeast the capacity to synthesize HA polymers in vitro. Wild-type yeast lacks the

ability to synthesize any GAGs due to lack of UDP-GlcA in the cells. The yeast cell should

therefore completely lack the machinery for GAG synthesis but the HAS polypeptide gives it that

capacity. Kimata and co-workers (229) managed to purify mouse HAS1 to homogeniety and

showed that it could synthesize HA in vitro.

4.2.2.2 Kinetics

The enzyme kinetics of synthases of both prokaryotic (198) and eukaryotic (86) origin have been

studied with regard to their substrates UDP-GlcA and UDP-GlcNAc. The data are summarized in

Table 4. The UDP-GlcNAc utilization of recombinant Streptococcus pyogenes (spHAS) and

Streptococcus equisimilis (seHAS) expressed in E. coli were studied and shown to be sigmoidal

for spHAS in a Michaelis-Menten plot. The indicated cooperativity for UDP-GlcNAc was seen

both for membrane fractions of spHAS (198) and for the solubilized form in the presence of

cardiolipin (197). The utilization of the UDP-GlcA was non-cooperative for both the bacterial

strains. Stoolmiller and Dorfman have also reported a non-hyperbolic plot for UDP-GlcNAc

26

utilization by Streptococcus (185). The enzyme kinetics displayed by spHAS may reflect an

evolutionary adaptation to ensure that cell growth is not imparied by the production of the HA

capsule (198).

The three vertebrate HAS proteins were assessed in a similar manner to the bacterial forms

(86). The enzymes displayed similar Km- and Vmax-values and showed no sign of cooperativity for

either substrate.

Table 4 Enzyme kinetics of prokaryotic and eukaryotic HAS enzymes

Source

Km for UDP-GlcA †

(µM)

Km for UDP-GlcNAc ††

(µM)

HA polymerization rate

(monosaccharide/min)

spHAS 27 125

seHAS 51 30

HAS1 53 501 1256

HAS2 33 331 1014

HAS3 35 223 174

Data taken from (198)(86). Membrane fractions of recombinant Streptococcal and mouse HAS

proteins were used to determine Km-values.† Concentration of UDP-GlcNAc was 500 µM†† Concentration of UDP-GlcA was 100 µM for bacterial HAS and 200 µM for mouse HAS

4.2.3 Regulation of HA biosynthesis

The HA biosynthesis can be regulated on different levels. As described in section 4.1.2 the access to

UDP-sugars can potentially constitute a regulatory mechanism for GAG synthesis. Several studies

have also been conducted on the growth factor stimulation of HA synthesis and the involvement of

protein kinases.

4.2.3.1 Turnover of the HAS protein

The turnover of the HAS proteins have only been studied indirectly and the results are

contradictory. Calabro and Hascall (15) studied HA and CS synthesis in chondrosarcoma cells

during brefaldin A treatment. Brefaldin A is a fungal metabolite that interferes with vesicular

transport causing disassembly of the Golgi complex and could thus indirectly interfere with the

delivery of the newly synthesized HAS proteins to the cell membrane. However, the HA synthesis

remained remarkably constant for a period of 8 hours (15). Bansal and Mason (8) came to different

conclusions using the same cell line when they added cycloheximide, an eukaryotic protein

27

synthesis inhibitor, to the cultured cells. They could see a 50 % reduction in HA production after

less than 2 hours treatment. Other studies have also indicated a more rapid turnover of HAS protein

using cycloheximide (68).

An explanation to the contradictory results of HAS protein turnover could be that different

HAS isozymes are influenced by the cell type that they are expressed in, by the cell culture

conditions and/or that the HAS isoforms exhibit different turnover rates. Itano et al. (86) have

shown that the three enzymes deviated from each other in terms of their catalytical stability in vitro,

HAS1 having the narrowest window of activity (1 hour) while the HAS2 (4 hours) and HAS3

(8 hours) windows were broader.

4.2.3.2 Effects of growth factors

Platelet derived growth factor (PDGF) is a growth factor with broad specificity which stimulates

proliferation in connective-tissue cells (67). The transforming growth factor (TGF) is a family of

growth factors acting on a broad range of cells exhibiting stimulatory and inhibitory effects in a cell

specific manner (143). Heldin and co-workers (68) demonstrated that PDGF-BB stimulation of

human mesothelial cells led to increased production of HA and that antibodies towards PDGF-BB

partly inhibited this increase (3). The authors showed that mesothelioma cells, the transformed

counterpart of mesothelial cells, could generate stimulatory signals for HA production.

Mesothelioma cells do not produce HA by themselves; rather they secrete factors that stimulate HA

synthesis in fibroblasts and mesothelial cells (3). Furthermore, inhibition of the de novo protein

synthesis with cycloheximide diminished the effect of PDGF in mesothelial cells, suggesting that

PDGF controls the HA synthesis at the transcriptional level (68). In another study, PDGF-BB and

TGF-ß1 were shown to stimulate HA synthesis in human foreskin fibroblasts (186). These signals

were also dependent on protein synthesis since cycloheximide inhibited in part the stimulatory

effects.

Schor and collegues have been studying HA production in skin fibroblasts in relation to

growth substratum and growth factors. These studies revealed that fetal and adult fibroblasts

upregulate their HA synthesis when cultured on a collagen substratum compared to plastic dishes,

and that TGF-ß1 stimulated HA synthesis in confluent fetal cells growing on plastic substratum, but

inhibited HA synthesis when cells were growing on collagen (38). In 3-dimensional collagen gels

fetal fibroblast HA synthesis was unaffected by PDGF, epidermal growth factor (EGF) and

fibroblast growth factor (FGF), but was inhibited by TGF-ß1. Adult fibroblast HA synthesis was

stimulated by the same factors, but was unaffected by TGF-ß1 (37).

More recently, the promotor sequence of mouse HAS1 have been cloned and partly

analyzed (223). Using the Internet Transfac program several potential regulatory cis-elements have

28

been identified, such as AP-2, CREB, MyoD, SRY and Sox-5, which in future might help to predict

and determine the importance of various growth stimuli for HA production.

4.2.3.3 Involvement of protein kinases

Several reports have indicated that protein kinases are involved in the stimulatory signals leading to

increased HA production, including direct phosphorylation of the HAS protein (125)(146). After

the cloning and characterization of the HAS enzyme family (220) caution is advised about the

conclusions drawn in these former reports.

Studies by Heldin and co-workers revealed that the stimulatory effect of PDGF on HA

biosynthesis in mesothelial cells and fibroblasts was partly dependent on protein kinase C and

partly dependent on de novo protein synthesis. Interestingly, the phorbol ester PMA, an activator of

protein kinase C, stimulated HAS proteins through a pathway which was not dependent on de novo

synthesis (68)(186). Treatment of mesothelial cells with phosphotyrosine phosphatase inhibitors

have also been shown to stimulate HA production (68)(134).

Stimulation of mesothelial cells from rabbit pericardial cavity with EGF and insulin-like

growth factor (IGF) gave a cooperative enhancement of HA biosynthesis, which could be

suppressed by a tyrosine kinase inhibitor (78). This result suggested that the inhibitory effect was

on the tyrosine kinase activity of the cytoplasmic domains of the IGF-I and EGF receptors, rather

than a direct influence on the HAS protein. Similar results with IGF-I were also seen in peritubular

rat testis cells (193). The HA production in pericardial mesothelial cells could be stimulated by

prostaglandin E2 through a cAMP-mediated signal, which could also be seen when the intracellular

level of cAMP was raised by addition of synthetic analogues or stimulators of cAMP production to

the cell culture media (79). They could also show that the signal transduction pathway from cAMP

was to protein kinase A and not protein kinase C in these cells.

Salustri and co-workers have shown that the cumulus cells increase their HA synthesis

during the expansion of the cumulus oophorus in the mammalian preovulatory follicle (163) in

response to follicle stimulating hormone (FSH) and a soluble oocyte factor(s) (162). FSH has been

shown to increase the intracellular cAMP levels, which correlated with the net increase of HA

production in mouse cumulus cells, while tyrosine kinase inhibitors suppressed the effect of FSH

partially (196). The HA synthesis was abolished after treatment with RNA synthesis inhibitors,

indicating that the induction of HA production is primarily controlled at the transcriptional level.

4.2.4 Biosynthetic directionality

A question that still remains unclear is the biosynthetic directionality of the HAS enzyme, i.e.

whether monosaccharides are added to the reducing end or the non-reducing end of the growing

HA polymer. In nature the predominant biosynthetic directionality for carbohydrate polymers

29

seems to be at the non-reducing end, exemplified by the sulfated GAGs (96), but elongation in the

reducing end has also been reported (polymerization of dextran (155)). The existence of two classes

of HAS enzymes, probably with different evolutionary origin (see section 4.2.1), makes it possible

that the different classes exhibit different polymerization mechanisms. Several attempts using

different approaches have been made to clarify the polymerization directionality of class I Has but

conclusive results are still lacking. However, the polymerization directionality of class II Has

(Pasturella multocida) has been determined.

Stoolmiller and Dorfman in 1969 showed experimental data that HAS protein from

Streptococcus polymerized the HA chain at the non-reducing end (185). The study was based on an

in vitro pulse experiment with UDP-[14C]GlcA alone, with subsequent digestion of the labeled HA

chain with Streptococcal hyaluronidase. The hyaluronidase digests ß1-4 linkages in the HA chain

by an eliminase reaction rendering predominantly unsaturated disaccharides (∆di-HA), and a smaller

portion of saturated disaccharides (di-HA), which are formed from the non-reducing terminal

disaccharide of the HA chain. The majority (88 %) of the radioactivity was found in di-HA after the

pulse experiment, indicating chain growth in the non-reducing end.

Prehm obtained the opposite result in 1983 when assaying differentiated teratocarcinoma

cells by pulse-chase experiments (144). HA polymerization was done with a pulse of UDP-

[14C]GlcA and UDP-GlcNAc followed by a chase with unlabeled UDP-GlcA. The pulse-chase

labeled HA was then degraded from the non-reducing end with exo-enzymes (ß-glucuronidase and

ß-N-acetylglucosaminidase). Prehm concluded that HA synthesis occurs at the reducing end since

the [14C]GlcA residues were cleaved off the HA chains directly in the presence of the digestive

enzymes; if the HA chain had been polymerized from the non-reducing end there would have been a

time lag before a drop in radioactivity could have been seen.

DeAngelis has studied the unique Pasturella multocida HAS expressed in the yeast

S. cerevisiae in terms of its polymerization directionality (27). The system has the advantage of

being devoid of nascent HA chains, since yeast cells are not able to synthesize UDP-GlcA. This

means that all HA produced in the in vitro HA polymerization assay is synthesized de novo. Under

these experimental conditions, DeAngelis was able to show addition of a monosaccharide unit to a

HA tetrasaccharide which could be cleaved off with exoglucosidases from the non-reducing end.

This was also the first time that a HA primer, other than UDP-GlcA or UDP-GlcNAc, could be

used in HA biosynthesis. In summary, rigid evidence that the P. multocida HAS elongates HA from

the non-reducing end was provided.

4.3 Biosynthesis of sulfated glycosaminoglycansThe three groups of sulfated GAGs, HS/heparin, CS/DS and KS, are synthesized as proteoglycans

in the Golgi. A linkage region, connecting the protein core with the GAG chain, initializes the chain

30

polymerization before the chain is further polymerized and modified by an elaborate biosynthetic

machinery.

4.3.1 Linkage region

The linkage region, connecting HS/heparin and CS/DS chains to a serine residue on the protein

core, consists of the tetrasaccharide GlcAß1-3Galß1-3Galß1-4Xylß1-Serine (156) and is initiated

by a unique transferase adding Xyl from a UDP-Xyl donor. The addition of a α-GlcNAc and ß-

GalNAc residue after the terminal GlcA residue of the linkage region constitutes the bifurcation of

HS/heparin and CS/DS biosynthesis, respectively. The aa sequence in the vicinity of the serine

residue partly influences the nature of the GAG chain that is going to be synthesized; repetitive

serine-glycine residues together with an acidic aa cluster and a hydrophobic patch promotes HS

polymerization (232)(231).

The linkage region in KS is different from the tetrasaccharide described above. KS chains

can either be linked to an asparagine or a serine/threonine thus resembeling the linkage of N- and

O-linked glycans, respectively (57).

4.3.2 Polymerization and modification

The heparin synthesis is the best studied and understood and is referred to as the default

modification (161). It begins with alternating transfer of GlcA and GlcNAc from corresponding

UDP-sugars to the linkage region, followed by the action of the first modification enzyme N-

deacetylase/N-sulfotransferase (NDST) that replaces the N-bound acetyl group on the GlcNAc

residue with a sulfate group. The subsequent enzymatic modifications are partly coupled to the

action of the NDST enzyme and therefore cannot take place before N-sulfation has occured. The

C5-epimerase converts GlcA to IdoA which promotes the 2-O-sulfation of the IdoA residue,

although GlcA residues can be modified in the same way. The last step in the chain synthesis is the

6-O-sulfation and selective 3-O-sulfation on GlcNAc residues. The default modification results in a

highly sulfated polymer with a typical disaccharide structure -IdoA(2-OSO3)-GlcNSO3(6-OSO3)-;

this disaccharide constitutes about 80 % of the disaccharides in heparin.

HS is structurally more diverse than heparin because the HS chains escape the default

modification pathway for reasons that are not fully understood. Only about 10 % of the

disaccharides in HS are default modified (161). Four different NDST isozymes have now been

cloned and seem to contribute to the different N-sulfation patterns seen in HS and heparin.

Transfection of NDST-2, an isozyme predominantly expressed in heparin-producing mast cells, into

HS producing cells increased their N-sulfation (17), which was not the case when more widely

distributed NDST-1 was overexpressed (84). Several isoforms of the 2-O-, 3-O and 6-O-

31

sulfotransferases exhibiting different substrate specificities in vitro have been cloned, that could

contribute to generation of specific sulfation patterns in vivo (115).

The polymerization of CS/DS and KS is less well understood (96). The GalNAc residues

remain acetylated in CS/DS chains rendering the chains less structurally diverse than HS. KS can

undergo modifications through selective 6-O-sulfation on both Gal and GlcNAc residues which are

non-randomly distributed along the chain.

5 Degradation of HA5.1 Tissue turnoverHA is constantly being turned over in the body. In a series of experiments performed by Laurent

and Fraser (108)(49)(46) it has been estimated that about one third of the total HA content (11-17 g

for a 70 kg man) is degraded per day. The degradation takes place both locally in the tissue and in

central organs such as the liver, kidney, spleen and bone marrow. The local tissue turnover rate of

HA has been estimated to t1/2=12 hours in the skin studying injected [3H]HA (151). The HA that

escapes degradation locally is transported via the lymph to the lymph nodes (191) where 50-90 %

of the HA is degraded (47). The blood receives HA from lymph nodes (10-100 mg/day), where it

disappears within minutes (t1/2= 2-6 min). The liver takes up about 90 % of the HA that reaches the

blood by a receptor mediated process (41). The receptor(s) have been studied functionally and

biochemically by several groups and their cloning seems to be at hand (120)(236).

5.2 HyaluronidasesThe HA that is taken up by the cells is degraded inside the lysosomes by hyaluronidases and the

exoenzymes ß-glucuronidase and ß-N-acetylglucosaminidase (108). The hyaluronidases are a

diverse group of enzymes isolated from different origins such as vertebrates, leeches and bacteria,

and the enzymes exhibit different kinds of activities (102). The vertebrate enzymes are endo-ß-N-

acetyl-D-hexosaminidases that degrade HA (and CS to a lesser extent) yielding tetrasaccharides and

hexasaccharides as final products.

Up to now six paralogue human hyaluronidase genes have been identified (21)(184) which

are arranged in two tightly linked triplets on chromosomes 3p21.3 (genes HYAL1, HYAL2,

HYAL3) and 7q31.3 (genes HYAL4, SPAM1, HYALP1) and codes for the proteins Hyal-1, Hyal-2,

Hyal-3 and Hyal-4, PH-20 respectively; SPAM1 is a pseudogene with no protein product. HYAL1,

HYAL2 and HYALP1 are widely expressed, whereas HYAL3 and HYAL4 are differentially

expressed in bone marrow and testis, and placenta and skeletal muscle, respectively. The genes for

human hyaluronidases are mapped to loci which have been described as candidate tumor suppressor

loci (101).

32

It is not known if the functional implications of the hyaluronidases are restricted to HA degradation

in the lysosomes due to the fact that the enzymes are only active in acidic pH (52); the role of

hyaluronidases remains largely to be elucidated. An exception is PH-20, a 64 kDa GPI-anchored

hyaluronidase acting at neutral pH, found in the sperm membrane (102). The protein enables the

sperm to penetrate the cumulus cell layer surrounding the egg during the fertilization process (114).

Hyaluronidases have been implicated in tumorigenesis. Frost et al. (54) showed that head

and neck squamous cell carcinomas had inactivated their HYAL1 gene by aberrant splicing of pre-

mRNA, which rendered the cells devoid of hyaluronidase activity. Additional pieces of information

suggesting a broader function of the hyaluronidases in HA metabolism have been presented. The

presence of Hyal-1 in human plasma and urine (10)(53) and the novel specificity of Hyal-2,

hydrolysing HA to 20 kDa fragments (50-60 disaccharide units) instead of tetrasachharides (110),

opens up the field for further investigations.

33

PRESENT INVESTIGATION

6 Aims of the studyAt the start of this study the knowledge about the molecular mechanisms that regulate HA

biosynthesis was limited. Purification of proteins involved in HA polymerization as well as studying

the expression and regulation HAS proteins were general topics of interest. We were also interested

in investigating the effects of HA overproduction for cell function and cell behavior. The specific

aims of the thesis were to study:

• if HA is synthesized by a functional multi-protein complex. Are there proteins associated with

the HAS protein that might assist in the HA polymerization process or regulation of it? If yes, is

it possible to co-purify these accessory proteins from solubilized lysates of HA producing cells?

For these studies, we wanted to develop immunoprecipitating antibodies towards the HAS

protein(s).

• the polymerization directionality of HA synthesis. Is HA polymerized in the same way as the

sulfated GAGs, where monosaccharides are added to the non-reducing end of the polymer or

does it have a unique way of polymerization from the reducing end?

• the overexpression of HAS proteins. Do the HAS isozymes synthesize the same product and are

they equally efficient in HA polymerization? What are the consequences of HA overproduction

for cell function?

• if the HAS isoforms are differentially expressed and regulated. Which signals and growth

factors are involved in upregulation of the HAS genes? Are there signals that can down-regulate

HAS gene and protein expression and thereby decrease HA production?

• the role of UDPGDH in GAG biosynthesis. How is the GAG production affected in cells

overexpressing the UDPGDH and/or HAS enzymes? Is UDPGDH a regulatory or rate-limiting

enzyme in mammalian GAG biosynthesis?

34

7 HAS enzyme in glioma cells can be solubilized in an active form

(Paper I) The knowledge about the enzyme(s) involved in HA biosynthesis was limited when this project was

initiated; none of the HAS isozymes were purified to homogeneity or cloned due to the difficulties

to study membrane associated proteins. We were interested in purifying the HAS protein and

identifying possible accessory proteins involved in the HA polymerization process or regulation of

it. During the progress of the project, our research group was involved in the characterization of the

cloned HAS1 gene (172), which made it possible for us to develop peptide antibodies against the C-

terminal part of the protein.

7.1 Solubilization and purification The glioma cell line U-118 MG produces large amounts of HA and was therefore a potentially

useful source to try to purify and characterize HA polymerizing protein(s). It was important to have

maximal enzymatic activity at the start of the solubilization and purification process, since the

method available to monitor the HAS protein was a functional in vitro assay. We discovered that

when glioma cells were pretreated with the phorbol ester PMA in combination with testicular

hyaluronidase prior to cell harvesting, the in vitro enzymatic activity was increased 23-fold. A

battery of different detergents was tested for their ability to solubilize HAS activity. Only digitonin

extracted an appreciable amount of the enzymatic activity (26 % of non-solubilized protein),

compared to other detergents such as CHAPS, Triton X-100 and Thesit (2-6 %). We concluded that

the HAS protein, despite its intimate association with the plasma membrane, can be solubilized in

active form.

The enzymatically active HAS protein was submitted to gel chromatography on a Superdex-

200 column. The eluted fractions from the column were tested for protein content, HA synthase

activity and immunoblotting. The fractions eluted just after the the void volume (about 600 kDa)

contained the highest HA synthase activity and also exhibited the strongest bands in Western

blotting. These bands were around 66 kDa, under both reducing and non-reducing conditions. A

12-fold purification measured in specific enzymatic activity was achieved after the chromatography.

In summary, we could show that the active HAS enzyme was eluted in the high molecular

weight fractions from a Superdex-200 column, suggesting that the HAS protein(s) forms a

complex with other components. Gel electrophoreses and immunoblotting indicate that such a

complex, if it exists, is not brought together with covalent disulfide bonds, but probably weaker

ionic interactions. In the literature, several reports have indicated that HA is polymerized by a single

polypeptide chain with no additional components (199)(147)(229)(see section 4.2.2). An

explanation to our result from the chromatography, where the HAS protein is eluted in the 600 kDa

35

fraction, is that partly synthesized HA chains are still connected to the HAS proteins during the

purification process thereby increasing the apparent size of the eluted proteins. The cell preparations

used for solubilization have a high HA content (about 2 mg/mg protein) which supports this

explanation (Marianne Svarvare, unpublished observations).

7.2 Antibodies towards HAS proteins The cloning of HAS1 (172) and HAS2 (177) led us to develop peptide antibodies towards different

epitopes of HAS in our laboratory. The GVR-antibody was produced against the C-terminal portion

of HAS1 (peptide sequence GVRRLCRRRTGGYRVQV) and the TIY-antibody against the C-

terminal portion of HAS2 (peptide sequence TIYKESKKPFSESKQT). The specificity of the

antibodies was tested against protein preparations from recombinant expressed HAS (Paper I) or

from cell lines with specific HAS mRNA expression and high HA production (Paper III). This

approch was complicated by the finding that some cell lines have low HA production but a more

pronounced mRNA expression of HAS genes (Jonas Brinck, unpublished observations) and the

difficulty to find native HAS null cells. In paper I, the GVR-antibody was used in immunoblotting

of the HAS protein in glioma cells, a cell line which later showed to express mainly HAS2

(Paper III). The explanation for this could be two-fold. First, the specificity of polyclonal peptide

antibodies is not absolute and there is a similarity of about 30 % between the HAS1 and HAS2 and

HAS3 protein sequences at the epitope site of the GVR-antibody. Second, glioma cells may exhibit

a low HAS1 expression that we did not detect with Northern blots (Paper III).

To be able to identify and study a possible functional HAS protein complex we decided to

develop immunoprecipitating antibodies against HAS using a GST (glutathione-S-transferase)

fusion protein system. Four GST-HAS constructs were developed against conserved and unique

parts of the intracellular domain of various HAS isoforms, expressed in a low protease active

Escherichia coli strain (BL21) and the purified proteins were injected into rabbits for antisera

production. Initial attempts to use these antibodies for immunoprecipitations of HAS proteins have

been unsuccessful.

8 Is HA synthesized from the reducing end? (Paper I) The polymerization directionality of the HAS enzyme had been adressed before (185)(144)(145)

resulting in contradictory conclusions. Therefore we addressed this question by further developing

the pulse-chase experiments performed by Prehm (144).

HAS containing glioma cell membranes were incubated in two different pulse-chase

experiments (see Fig. 5). In the first approach (A), HA was polymerized during a pulse period in the

presence of UDP-[14C]GlcA and UDP-GlcNAc, followed by a chase period where the UDP-

36

[14C]GlcA had been competed out by high concentration of unlabeled UDP-GlcA. In the second

approach (B), HA was polymerized during a pulse period with UDP-GlcA and UDP-GlcNAc,

followed by a chase with addition of UDP-[14C]GlcA. The unincorporated radioactivity was washed

away before the samples were submitted to hydrolysis at the non-reducing end with ß-glucuronidase

and ß-N-acetylglucosaminidase in dialysis bags. Results showed that the HA chains polymerized in

approach A were immediately susceptible to hydrolysis, whereas the chains polymerized in

approach B had a time lag between the start of the enzyme digestion and the drop in radioactivity. In

conclusion, these results support the hypothesis that HA is polymerized at the reducing end.

synthesis

-R

degradation

-R

Approach Asynthesis

-R

degradation

-R

Approach B

100

80

60

40

20

30 60 90

No

n-di

alys

able

rad

ioac

tivity

(%)

Time (min)

Figure 5. Schematic picture of the pulse-chase experiments of HA polymerization and the result of

hydrolysis of pulse-chased chains in dialysis bags.

The -R indicates the reducing end of the chain and the black bars show the [14C]labeled parts under

the assumption that synthesis occurs at the reducing end. Symbols in the lower panel: approach A,

filled boxes; approach B, empty boxes; control with no exoenzymes, filled circles.

The pulse-chase experiments reported here and elsewhere (185)(144) have obvious limitations when

it comes to the experimental set-up. At the starting point of the pulse-chase incubation the HAS

membranes contain already initialized HA chains that will be elongated together with de novo

37

synthesised chains, which will create heterogenity in the synthesized HA species. The digestion of

the pulse-chase labeled HA chains by exoglucosidases cannot be achieved completely giving

substantial levels of non-degradable radioactivity. One explanation to these residual levels could be

that the part of the HA chain closest to the HAS protein is less accessible for enzymatic degradation.

In summary, rigid evidence to the question of polymerization directionality of HA is still lacking.

9 Expression of HAS affects cell behavior (Paper II) After the identification and cloning of HAS in 1996 (i.e. HAS1) the cloning of two other isoforms

(HAS2, HAS3) was done during 1996-1997 by separate groups, thus revealing a potentially new

possibility for the regulation of the HA biosynthesis. The aim of this project was to study the

functional expression of the three HAS isoforms and the effects of overproduction of HA on cell

behavior.

9.1 Synthesis and accumulation of HA The Chinese hamster ovary (CHO) cell line was chosen for HAS expression due to its relatively

high genomic stability and high expression levels. Cells were transfected with cDNAs from human

HAS1, mouse HAS2 and mouse HAS3, and stabile expressing clones were established. The

sequence similarities between the mouse and human HAS orthologues is high, about 98 % (180),

indicating a conserved mechanisms of action between the species; we concluded that we could

compare the expression of the different cDNAs in the same system. Screening the clones for HA in

the cell culture media displayed a difference in HA synthesizing capacity between the different HAS

isoforms. The HAS3 transfected clones exhibited the highest HA synthesizing capacity (at

maximum 1.2 µg/105 cells/24 hours) compared to 0.07 µg and 0.6 µg per 105 cells/24 hours for the

most efficiently transfected HAS1 and HAS2 clones, respectively. The increase seen in the HAS3

clone corresponds to a 1000-fold increase in HA production compared to wild-type CHO cells. The

HA content in the culture media of the transfected cells (up to 1.2 µg/105 cells/24 hours) is in the

same range as in media of human mesothelial cells and glioma cells (Paper III). This indicates that

the HA amounts produced are within the physiological range, although not for the CHO cell line.

We also wanted to investigate if overexpression of HAS enzymes could lead to increased

amounts of cell associated HA. A classical way of visualizing pericellular HA since the 1960’s has

been the in vitro particle exclusion assay (18), where the cell surface associated HA can be seen as a

halo between the cell membrane and excluded particles (e.g. erythrocytes). In our CHO cell

model, we could show that whereas wild-type cells did not have the capacity to form HA matrices,

38

their transfected counterparts attained this ability. The HA synthesizing capacity needed to generate

such matrices was about 1 µg of HA per 105 cells and 24 hours. This accumulation of HA around

transfected cells could be confirmed by comparing cell associated [3H]HA amounts from low and

high HA producing clones.

9.2 HA chain length The chain length of HA synthesized by HAS transfected CHO cells were studied both after in vitro

and in vivo synthesis. The in vitro assay is based on HA polymerization by membrane preparations

incubated with UDP-[14C]GlcA,UDP-GlcNAc and MgCl2, while HA synthesized in intact cells in

the presence of [3H]glucosamine mimics in vivo conditions. Results showed that only HAS2

produced high molecular weight HA (Mr >3.9 x 106) when assayed in vitro, compared to HAS1

(average Mr 0.12 x 106) and HAS3 (Mr 0.12-1 x 106). However, under in vivo conditions, all three

isozymes generated HA of high molecular weight (Mr >3.9 x 106). As a comparison, HA chains

synthesized by mesothelial cells (cells with native expression of HAS1, HAS2 and HAS3;

Paper III) were of high molecular weight both in vitro and in vivo.

We also studied the HA chain length after solubilization of the HAS protein from glioma

cells (Paper I) and after stimulation of mesothelial cells with various growth factors and other

compounds (Paper III). Glioma cells, which predominantly express the HAS2 isoform, produced

high molecular weight HA when assayed in vitro. Solubilization of the glioma membranes with

digitonin prior to the assay rendered a more polydisperse product, but still HA polymers were still

produced. Treatment of mesothelial cells with growth factors such as PDGF-BB and TGF-ß led to

HA synthesis of high molecular weight in vivo. Interestingly, phorbol ester PMA pretreatment

induced synthesis of HA chains of polydisperse character (Mr 0.25 x 106 to over 3.9 x 106).

The results show that all three HAS isoforms, whether recombinantly expressed in CHO

cells or native in human mesothelial cells, have the capacity to synthesize high molecular weight HA

in vivo. We also show that the HA chain length is dependent on the cell type where it is synthesized.

Taken together with the result that the phorbol ester PMA can influence the chain size in mesothelial

cells, interesting questions are raised about the molecular mechanims behind the regulation of HA

chain length, which needs further investigation. The matter is important since the function of HA

polymers is partly dependent on their size. Large HA polymers are important structural elements in

the extracellular matrix (106) whereas short HA fragments can induce angiogenesis (221)(148) and

inflammatory reponses (121)(80).

39

9.3 Cell migration The migration capacity of wild-type CHO cells and HAS3 transfectants exhibiting different HA

synthesizing capacities were quantified using a Boyden chamber assay. The results revealed that the

migration capacity was inversely correlated to the HA synthesizing capacity. When excessive

amounts of exogenous HA was added to the migration assay, the migration was inhibited for cells

producing low amounts of HA, but not for high producing cells. Another finding was that the

amounts of cell surface CD44 (as detected by flow cytometry) decreased in high HA producing

cells. For the HAS3 clone producing the highest amounts of HA, a 75 % decrease in migration rate

and a 50 % reduction in cell surface CD44 was observed. When an antibody towards CD44,

blocking the HA binding site, was included in the assays the migration was inhibited up to 50 %.

In conclusion, our results suggests that the overproduction and accumulation of pericellular

HA impairs normal cell function and reduces cell migration in HAS transfected cells. The

concomitant down-regulation of CD44 molecules on the surface of these cells explains partly the

reduction in their migratory ability. The total amount of CD44 does not seem to reduce significantly

when visualized in Western blotting of total cell lysates (Jonas Brinck, unpublished observations),

suggesting a receptor mediated endocytosis of the receptor; such a mechanism has been described

in the literature (22). Knudson and Knudson (97) have presented the hypothesis where the

establishment of a pericellular HA matrix would interfere with and disrupt normal cell migration.

This model has to some extent been experimentally supported by Perris et al. (141), where HA

binding proteoglycans that promote pericellular matrix formation was shown to inhibit migration in

neural crest cells.

10 HAS isoforms show specificity in gene regulation (Paper III) Under pathological conditions such as chronic inflammation and carcinogenesis an upregulation of

HA production is seen, which could be of importance for the pathological development of the

disease state (190)(203). To understand the regulation of the HAS genes we therefore investigated

the expression HAS genes in response to growth conditions and external stimuli.

10.1 Effects of culture conditions Subconfluent and confluent cultures of glioma cells, mesothelioma cells, mesothelial cells, skin

fibroblasts and lung fibroblasts were tested for their ability to express the three HAS isoforms. The

culture conditions affected the HAS expression levels by reducing all HAS genes when the cell

density increased, indicating a cell-cell contact as a negative factor for HAS expression. Northern

blots revealed that only mesothelial cells expressed all three HAS isoforms. The mRNA levels of

HAS in the various cell cultures were well correlated with their HA production. Mesothelial cells

40

displayed the highest HA synthesizing capacity (approx. 18 ng HA per 103 cells after 24 hours of

culture). Interestingly, mesothelioma cells had a different HAS mRNA expression pattern compared

to mesothelial cells; mesothelioma cells expressed HAS3 isoform and synthesized only minute

amounts of HA.

10.2 Effects of external stimuli Mesothelial cells were chosen for studying the differential expression of the HAS genes in response

to external stimuli since they have native expression of all three isoforms. We stimulated mesothelial

cells with the PDGF-BB and TGF-ß1, which are known to be important regulators of HA synthesis

in cells of mesenchymal origin (68)(186). Stimulation of the cells with PDGF-BB increased the

mRNA level of HAS2 to a peak at 6 hours whereas TGF-ß1 treatment attenuated it. The HAS2

mRNA increase observed after PDGF-BB stimulation was accompanied by increased protein levels

and production of high molecular weight HA. The effects of growth factors on HAS1 and HAS3

mRNA expression levels were less pronounced as for HAS2. Treatment of the mesothelial cells

with TGF-ß1 increased HAS1 isoform at both the mRNA and protein level whereas HAS3 isoform

was not markedly affected by either PDGF-BB or TGF-ß1.

Treatment of mesothelial cells with hydrocortisone (125 µg/ml) suppressed HAS2 mRNA

expression considerably (about 80 % suppression compared to normal levels), whereas HAS1 and

HAS3 mRNA levels only decreased marginally or not at all. The down-regulatory effect of

hydrocortisone at HAS2 mRNA level was also seen at the protein level and on the production of

HA. Approximately a 40 % decrease in HA production was observed after 6 hours of treatment with

hydrocortisone compared to control treated cells.

In conclusion, we have found that the three HAS genes are differentially regulated in

different cell types in response to growth conditions and external stimuli. It seems that the growth

factors mediate their major stimulatory effects through the HAS2 isoform, and therefore HAS2

could be the most important enzyme for HA production in the body. Such an hypothesis is in line

with the fact that only HAS2 knock-out mice are embryonically lethal compared to HAS1 and

HAS3 knock-outs (181)(16). In addition, the functional inhibition of HA production seen, after

down-regulation of HAS2 caused by hydrocortisone treatment, could constitute a strategy for

pharmacological treatment of disease in a future perspective.

11 The role of UDP-glucose dehydrogenase in mammalianglycosaminoglycan biosynthesis (Paper IV) The UDPGDH enzyme converts UDP-Glc to UDP-GlcA, which is an essential building block in

mammalian GAG biosynthesis. The enzyme has been postulated to play a key regulatory role in the

cell wall polysaccharide biosynthesis in plants (192)(154) and is indispensable for vertebrate

41

embryonic development (179)(181) by being the sole producer of UDP-GlcA. The aim of this

study was to investigate the role of UDPGDH in mammalian GAG biosynthesis.

We have created an experimental cell system where the rate of synthesis and consumption of

UDP-GlcA can be modulated in human embryonic kidney (293T) cells by transient overexpression

of UDPGDH and HAS3. The dehydrogenase and HA synthase activities were measured in vitro

and were found to be over 400-fold greater in favor of the dehydrogenase activity in 293T cells;

even after the cells were transfected with UDPGDH and/or HAS3 the overcapacity of the

dehydrogenase enzyme persisted. The mRNA expression of UDPGDH was also investigated in

CHO cells stably transfected with HAS3. The basal mRNA expression level of UDPGDH in wild-

type cells was sufficient for HA production in HAS3 transfected cells exhibiting high HA

production; the levels of mRNA in HAS3 expressing cells did not increase after transfection.

The amounts of synthesized GAGs were determined after UDPGDH and/or HAS3

transfection. Overexpression of UDPGDH increased the production of CS about 1.5-fold, but did

not affect the amount of synthesized HA and HS. HAS3 transfection on the other hand, led to a 20-

fold increase of HA production. The overproduction of HA was in part at the expense of HS

production which was reduced to 50 % compared to mock transfected cells (CS amounts were

normal). Co-expression of UDPGDH with HAS3 did not rescue HS production back to normal

levels or increase HA or CS production. The isolated HA, HS and CS chains from the UDPGDH

and/or HAS3 transfected cells did not display any major structural changes, in terms of their chain

lengths or their the degree of sulfation (HS and CS), compared to GAG chains from mock

transfected cells.

In conclusion, our findings indicate that the activity of UDPGDH is not rate limiting in

mammalian GAG biosynthesis. The diverse effects seen on the HS and CS production in response

to UDPGDH and HAS3 transfection could suggest that the biosynthetic machineries responsible

for HS and CS synthesis are located in separate Golgi compartments, or that they have different

affinities for the substrates (Km-values). Studies are currently underway to determine the

concentrations and the fluctuations in the UDP-sugar pools, which will assist in the interpretation of

the effects seen on HA, HS and CS production.

42

12 Conclusions• We have shown that the HAS protein(s) can be solubilized in active form. Results indicate that

the partially purified HAS isozyme(s) is most likely not asscociated covalently with any

accessory components that may be needed in the regulation of HA biosynthesis.

• We have results supporting the hypothesis that HA is polymerized at the reducing end of the

polymer thereby being unique in the polymerization directionality in the family of GAGs.

• We have shown that all three HAS isoforms are able to synthesize high molecular weight HA in

intact mammalian cells and that the regulation of the HA chain length involves cell specific

elements as well as external stimulatory factors.

• We have shown that overexpression of HAS in CHO cells leads to increased HA production and

accumulation of pericellular HA and that these cells exhibit a reduced migration capacity and

reduced amounts of cell surface CD44 molecules.

• We have shown that the HAS1, HAS2 and HAS3 genes are differentially expressed and

regulated in response to culture conditions and external stimuli. PDGF-BB and TGF-ß1 are

important regulators of HAS at both transcriptional and translational level.

• We have results indicating that UDPGDH is not a rate-limiting enzyme in mammalian GAG

biosynthesis.

43

13 Future perspectives The sequencing of the whole human genome is soon complete (175), revealing the genetic code of

man, whereas the work with discovering the function of many of these genes is still in its infancy.

Although the three genes involved in the biosynthesis of HA were sequenced and cloned during the

time period of this thesis study, the action and regulation of the HAS enzymes are still topics of

scientific interest and many questions remain to be answered.

The mechanism of how small HA fragments, found in extracellular matrix of various

mammalian tissues under physiological and pathological conditions, are formed is still not resolved.

Reports by us and others (Paper II, (180)(86)) indicating that HA can be synthesized as shorter

chains can be an in vitro finding with little or no physiological relevance. Two main stand-points can

be taken when trying to explain the origin of these fragments. The first conjecture is that HA is

always synthesized as full size polymers (Mr 5-10 x 106) in vivo. In such a model, the production of

low molecular weight HA could instead be due to hyaluronidase activity. The fact that the human

hyaluronidase gene family now comprise six paralogous genes corresponding to five different

hyaluronidase proteins strengthens their role in HA metabolism (21)(184). Specific growth factors

could thereby regulate hyaluronidase expression and promote or decrease the formation of HA

oligosaccharides. However, before these enzymes (PH-20 excluded) can be ascribed an essential

role in HA chain regulation rigid evidence for hyaluronidase activity in the extracellular matrix at

physiological pH must first be provided.

The second conjecture is that the regulation of the HA chain length is controled by the HAS

molecule itself, or an accessory component, during the process of synthesis. Preliminary results by

Spicer and co-workers show that a positive aa cluster in the C-terminal part of the HAS protein

displays chain length regulation capacity by binding to the nascent intracellular part of the HA chain

(178). More suggestive evidence, based on the fact that HA polymerization is regulated in a cell

specific manner (Paper II)(180)(86), indicates that an additional component might be involved. The

nature of this component, if it exists, remains to be elucidated.

In several pathological conditions such as carcinogenesis and chronical inflammation

elevated levels of HA can be detected. Is this increased production of HA a primary or secondary

effect of the pathological condition and how are the other GAGs affected by the shift in HA

synthesis? In paper IV we investigated the potentially regulatory role of UDP-sugars and

UDPGDH for GAG biosynthesis. Our results indicate that the UDPGDH enzyme does not exert a

rate-limiting effect on mammalian GAG biosynthesis. However, the HS production was markedly

affected in cells overproducing HA, suggesting that shifts in UDP-sugar fluxes can influence their

GAG production quantitatively. It will be of importance to investigate the susceptibility of GAG

biosynthetic machineries to changes in UDP-sugar concentrations, since it could constitute a

mechanism of regulation. The functional diversity of the three HAS isoforms in terms of their HA

44

coat forming capacity (86) could be an example of this kind of regulation. Also, the observed

vulnerability of the HS production in HA overproducing cells could lead to an impairment in the

affected cells’ growth factor signaling, which is known to be mediated through HS containing

proteoglycans (159).

45

ACKNOWLEDGEMENTS

I wish to express my gratitude to:

Evi Heldin, my supervisor, for sharing her enthusiasm for science with me and teaching me to be a

critical scientist.

Professors Ulf Lindahl and Torvard Laurent for guidance and encouragement in the field of science

in general and glycosaminoglycans in particular.

Marion Kusche-Gullberg and Dorothe Spillman for generous supervision in the field of complex

polysaccharides. All my co-authors and collaborators for their work and advice in the four papers

presented in this thesis.

Marianne Svarvare for always being at hand with excellent technical advice and for teaching me all

the practical things I thought I already knew but apparently did not.

Örjan Zetterqvist, my mentor in everything that has concerned teaching of the graduate students, for

inspiring discussion and for entrusting me with part of the graduate student education. It has been

very stimulating and challenging!

Annica, Johan and Mehdi for creating that day-to-day (year-to-year!) atmosphere that makes it

possible to get through four years of pipetting and cell culture work. Don’t give up – your day will

come soon! Special thanks to Maria for letting me borrow her brother’s “slagborrmaskin” and

many, many other things. Also thanks to all the people in Ulf Lindahl’s group, and especially

Elisabet, for partially adopting me (and “my” polysaccharide) into your group.

All the persons, past and present, at the department for creating a pleasant working atmosphere.

The people in Lena Claesson-Welsh’s group for letting me in on a few secrets in field of cell

biology and for teaching me how to use your (expensive) equipment.

Anna for being at my side.

46

REFERENCES1. Aruffo, A., Stamenkovic, I., Melnick, M., Underhill, C.B. and Seed, B. (1990) CD44 is the

principal cell surface receptor for hyaluronate. Cell. 61:1303-1313.

2. Asplund, T. and Heldin, P. (1994) Hyaluronan receptors are expressed on human malignant

mesothelioma cells but not on normal mesothelial cells. Cancer Res. 54:4516-4523.

3. Asplund, T., Versnel, M.A., Laurent, T.C. and Heldin, P. (1993) Human mesothelioma cells

produce factors that stimulate the production of hyaluronan by mesothelial cells and

fibroblasts. Cancer Res. 53:388-392.

4. Bajorath, J., Greenfield, B., Munro, S.B., Day, A.J. and Aruffo, A. (1998) Identification of

CD44 residues important for hyaluronan binding and delineation of the binding site. J Biol

Chem. 273:338-343.

5. Balduini, C., Brovelli, A. and Castellani, A.A. (1970) Biosynthesis of glycosaminoglycans in

bovine cornea. The effect of uridine diphosphate xylose. Biochem J. 120:719-723.

6. Banerji, S., Day, A.J., Kahmann, J.D. and Jackson, D.G. (1998) Characterization of a

functional hyaluronan-binding domain from the human CD44 molecule expressed in

Escherichia coli. Protein Expr Purif. 14:371-381.

7. Banerji, S., Ni, J., Wang, S.X., Clasper, S., Su, J., Tammi, R., Jones, M. and Jackson, D.G.

(1999) LYVE-1, a new homologue of the CD44 glycoprotein, is a lymph-specific receptor for

hyaluronan. J Cell Biol. 144:789-801.

8. Bansal, M.K. and Mason, R.M. (1986) Evidence for rapid metabolic turnover of hyaluronate

synthetase in Swarm rat chondrosarcoma chondrocytes. Biochem J. 236:515-519.

9. Bdolah, A. and Feingold, D.S. (1968) Uridine diphosphate-D-glucose dehydrogenase of hen

oviduct. Biochim Biophys Acta. 159:176-178.

10. Bonner, W.M., Jr. and Cantey, E.Y. (1966) Colorimetric method for determination of serum

hyaluronidase activity. Clin Chim Acta. 13:746-752.

11. Bourguignon, L.Y., Kalomiris, E.L. and Lokeshwar, V.B. (1991) Acylation of the lymphoma

transmembrane glycoprotein, GP85, may be required for GP85-ankyrin interaction. J Biol

Chem. 266:11761-11765.

12. Bourin, M.C. and Lindahl, U. (1993) Glycosaminoglycans and the regulation of blood

coagulation. Biochem J. 289:313-330.

13. Brandli, A.W., Hansson, G.C., Rodriguez-Boulan, E. and Simons, K. (1988) A polarized

epithelial cell mutant deficient in translocation of UDP-galactose into the Golgi complex. J

Biol Chem. 263:16283-16290.

47

14. Brown, T.A., Bouchard, T., St. John, T., Wayner, E. and Carter, W.G. (1991) Human

keratinocytes express a new CD44 core protein (CD44E) as a heparan-sulfate intrinsic

membrane proteoglycan with additional exons. J Cell Biol. 113:207-221.

15. Calabro, A. and Hascall, V.C. (1994) Differential effects of brefeldin A on chondroitin sulfate

and hyaluronan synthesis in rat chondrosarcoma cells. J Biol Chem. 269:22764-22770.

16. Camenisch, T.D., Spicer, A.P., Brehm-Gibson, T. and McDonald, J.A. (1999) Identification

of a hyaluronan synthase2 as a critical source of hyaluronan during mammalian

embryogenesis. Conference abstract in "International Conference on Molecular Interactions

of Proteoglycans", Kanagawa, Japan.

17. Cheung, W.F., Eriksson, I., Kusche-Gullberg, M., Lindhal, U. and Kjellen, L. (1996)

Expression of the mouse mastocytoma glucosaminyl N-deacetylase/ N-sulfotransferase in

human kidney 293 cells results in increased N-sulfation of heparan sulfate. Biochemistry.

35:5250-5256.

18. Clarris, B.J. and Fraser, J.R. (1968) On the pericellular zone of some mammalian cells in

vitro. Exp Cell Res. 49:181-193.

19. Crater, D.L., Dougherty, B.A. and van de Rijn, I. (1995) Molecular characterization of hasC

from an operon required for hyaluronic acid synthesis in group A streptococci.

Demonstration of UDP- glucose pyrophosphorylase activity. J Biol Chem. 270:28676-

28680.

20. Crater, D.L. and van de Rijn, I. (1995) Hyaluronic acid synthesis operon (has) expression in

group A streptococci. J Biol Chem. 270:18452-18458.

21. Csoka, A.B., Scherer, S.W. and Stern, R. (1999) Expression analysis of six paralogous

human hyaluronidase genes clustered on chromosomes 3p21 and 7q31. Genomics. 60:356-

361.

22. Culty, M., Nguyen, H.A. and Underhill, C.B. (1992) The hyaluronan receptor (CD44)

participates in the uptake and degradation of hyaluronan. J Cell Biol. 116:1055-1062.

23. Dalchau, R., Kirkley, J. and Fabre, J.W. (1980) Monoclonal antibody to a human brain-

granulocyte-T lymphocyte antigen probably homologous to the W 3/13 antigen of the rat. Eur

J Immunol. 10:745-749.

24. Dasgupta, A., Takahashi, K., Cutler, M. and Tanabe, K.K. (1996) O-linked glycosylation

modifies CD44 adhesion to hyaluronate in colon carcinoma cells. Biochem Biophys Res

Commun. 227:110-117.

25. Day, A.J. and Parkar, A.A. (1998) The structure of the Link module: a hyaluronan binding

domain. In "The Chemistry, Biology and Medical Applications of Hyaluronan and its

Derivatives". Vol. 72. Laurent, T.C., editor. Portland Press, London. 141-147.

48

26. DeAngelis, P.L. (1999) Hyaluronan synthases: fascinating glycosyltransferases from

vertebrates, bacterial pathogens, and algal viruses. Cell. Mol. Life Sci. 56:670-682.

27. DeAngelis, P.L. (1999) Molecular directionality of polysaccharide polymerization by the

Pasteurella multocida hyaluronan synthase. J Biol Chem. 274:26557-26562.

28. DeAngelis, P.L. and Achyuthan, A.M. (1996) Yeast-derived recombinant DG42 protein of

Xenopus can synthesize hyaluronan in vitro. J Biol Chem. 271:23657-23660.

29. DeAngelis, P.L., Jing, W., Drake, R.R. and Achyuthan, A.M. (1998) Identification and

molecular cloning of a unique hyaluronan synthase from Pasturella multicoda. J Biol Chem.

273:8454-8458.

30. DeAngelis, P.L., Jing, W., Graves, M.V., Burbank, D.E. and Van Etten, J.L. (1997)

Hyaluronan synthase of chlorella virus PBCV-1. Science. 278:1800-1803.

31. DeAngelis, P.L., Papaconstantinou, J. and Weigel, P.H. (1993) Molecular cloning,

identification, and sequence of the hyaluronan synthase gene from group A Streptococcus

pyogenes. J Biol Chem. 268:19181-19184.

32. DeGrendele, H.C., Estess, P. and Siegelman, M.H. (1997) Requirement for CD44 in activated

T cell extravasation into an inflammatory site. Science. 278:672-675.

33. Dougherty, B.A. and van de Rijn, I. (1992) Molecular characterization of a locus required for

hyaluronic acid capsule production in group A streptococci. J Exp Med. 175:1291-1299.

34. Dougherty, B.A. and van de Rijn, I. (1993) Molecular characterization of hasB from an

operon required for hyaluronic acid synthesis in group A streptococci. Demonstration of

UDP- glucose dehydrogenase activity. J Biol Chem. 268:7118-7124.

35. Dougherty, B.A. and van de Rijn, I. (1994) Molecular characterization of hasA from an

operon required for hyaluronic acid synthesis in group A streptococci. J Biol Chem. 269:169-

175.

36. Eggli, P.S. and Graber, W. (1995) Association of hyaluronan with rat vascular endothelial and

smooth muscle cells. J Histochem Cytochem. 43:689-697.

37. Ellis, I., Banyard, J. and Schor, S.L. (1997) Differential response of fetal and adult fibroblasts

to cytokines: cell migration and hyaluronan synthesis. Development. 124:1593-1600.

38. Ellis, I.R. and Schor, S.L. (1996) Differential effects of TGF-beta1 on hyaluronan synthesis

by fetal and adult skin fibroblasts: implications for cell migration and wound healing. Exp

Cell Res. 228:326-333.

39. Entwistle, J., Hall, C.L. and Turley, E.A. (1996) HA receptors: regulators of signalling to the

cytoskeleton. J Cell Biochem. 61:569-577.

49

40. Entwistle, J., Zhang, S., Yang, B., Wong, C., Li, Q., Hall, C.L., A, J., Mowat, M., Greenberg,

A.H. and Turley, E.A. (1995) Characterization of the murine gene encoding the hyaluronan

receptor RHAMM. Gene. 163:233-238.

41. Eriksson, S., Fraser, J.R., Laurent, T.C., Pertoft, H. and Smedsrod, B. (1983) Endothelial cells

are a site of uptake and degradation of hyaluronic acid in the liver. Exp Cell Res. 144:223-

228.

42. Faassen, A.E., Schrager, J.A., Klein, D.J., Oegema, T.R., Couchman, J.R. and McCarthy, J.B.

(1992) A cell surface chondroitin sulfate proteoglycan, immunologically related to CD44, is

involved in type I collagen-mediated melanoma cell motility and invasion. J Cell Biol.

116:521-531.

43. Feinberg, R.N. and Beebe, D.C. (1983) Hyaluronate in vasculogenesis. Science. 220:1177-

1179.

44. Fessler, J.H. and Fessler, L.I. (1966) Electron microscopic visualization of the polysaccharide

hyaluronic acid. Proc Natl Acad Sci U S A. 56:141-147.

45. Fieber, C., Plug, R., Sleeman, J., Dall, P., Ponta, H. and Hofmann, M. (1999) Characterisation

of the murine gene encoding the intracellular hyaluronan receptor IHABP (RHAMM). Gene.

226:41-50.

46. Fraser, J.R., Appelgren, L.E. and Laurent, T.C. (1983) Tissue uptake of circulating hyaluronic

acid. A whole body autoradiographic study. Cell Tissue Res. 233:285-293.

47. Fraser, J.R., Kimpton, W.G., Laurent, T.C., Cahill, R.N. and Vakakis, N. (1988) Uptake and

degradation of hyaluronan in lymphatic tissue. Biochem J. 256:153-158.

48. Fraser, J.R., Laurent, T.C. and Laurent, U.B. (1997) Hyaluronan: its nature, distribution,

functions and turnover. J Intern Med. 242:27-33.

49. Fraser, J.R., Laurent, T.C., Pertoft, H. and Baxter, E. (1981) Plasma clearance, tissue

distribution and metabolism of hyaluronic acid injected intravenously in the rabbit. Biochem

J. 200:415-424.

50. Freeze, H. (1999) Monosaccharide Metabolism. In "Essentials of Glycobiology". Varki, A.,

Cummings, R., Esko, J., Freeze, H., Hart, G. and Marth, J., editors. Coldspring Harbor

Laboratory Press, New York. 69-84.

51. Friedlander, D.R., Milev, P., Karthikeyan, L., Margolis, R.K., Margolis, R.U. and Grumet, M.

(1994) The neuronal chondroitin sulfate proteoglycan neurocan binds to the neural cell

adhesion molecules Ng-CAM/L1/NILE and N-CAM, and inhibits neuronal adhesion and

neurite outgrowth. J Cell Biol. 125:669-680.

52. Frost, G.I., Csoka, T. and Stern, R. (1996) The hyaluronidases: A Chemical, Biological and

Clinical Overview. Trends in Glycoscience and Glycotechnology. 8:419-434.

50

53. Frost, G.I., Csoka, T.B., Wong, T. and Stern, R. (1997) Purification, cloning, and expression

of human plasma hyaluronidase. Biochem Biophys Res Commun. 236:10-15.

54. Frost, G.I., Mohapatra, G., Wong, T.M., Csoka, A.B., Gray, J.W. and Stern, R. (2000)

HYAL1LUCA-1, a candidate tumor suppressor gene on chromosome 3p21.3, is inactivated in

head and neck squamous cell carcinomas by aberrant splicing of pre-mRNA. Oncogene.

19:870-877.

55. Gainey, P.A. and Phelps, C.F. (1972) Uridine diphosphate glucuronic acid production and

utilization in various tissues actively synthesizing glycosaminoglycans. Biochem J. 128:215-

227.

56. Glaser, L. (1959) The Biosynthesis of N-Acetylgalactoseamine. J Biol Chem. 234:2801-

2805.

57. Greiling, H. (1994) Structure and biological functions of keratan sulfate proteoglycans. Exs.

70:101-122.

58. Grumet, M., Flaccus, A. and Margolis, R.U. (1993) Functional characterization of chondroitin

sulfate proteoglycans of brain: interactions with neurons and neural cell adhesion molecules. J

Cell Biol. 120:815-824.

59. Gunthert, U., Hofmann, M., Rudy, W., Reber, S., Zoller, M., Haussmann, I., Matzku, S.,

Wenzel, A., Ponta, H. and Herrlich, P. (1991) A new variant of glycoprotein CD44 confers

metastatic potential to rat carcinoma cells. Cell. 65:13-24.

60. Hall, C.L., Yang, B., Yang, X., Zhang, S., Turley, M., Samuel, S., Lange, L.A., Wang, C.,

Curpen, G.D., Savani, R.C., Greenberg, A.H. and Turley, E.A. (1995) Overexpression of the

hyaluronan receptor RHAMM is transforming and is also required for H-ras transformation.

Cell. 82:19-28.

61. Handley, C.J. and Phelps, C.F. (1972) The biosynthesis in vitro of keratan sulphate in bovine

cornea. Biochem J. 128:205-213.

62. Handley, C.J. and Phelps, C.F. (1972) The concentrations of sugar nucleotides in bovine

corneal epithelium and endothelium. Biochem J. 127:911-912.

63. Hardwick, C., Hoare, K., Owens, R., Hohn, H.P., Hook, M., Moore, D., Cripps, V., Austen, L.,

Nance, D.M. and Turley, E.A. (1992) Molecular cloning of a novel hyaluronan receptor that

mediates tumor cell motility. J Cell Biol. 117:1343-1350.

64. Hathcock, K.S., Hirano, H., Murakami, S. and Hodes, R.J. (1993) CD44 expression on

activated B cells. Differential capacity for CD44-dependent binding to hyaluronic acid. J

Immunol. 151:6712-6722.

65. Hayen, W., Goebeler, M., Kumar, S., Riessen, R. and Nehls, V. (1999) Hyaluronan stimulates

tumor cell migration by modulating the fibrin fiber architecture. J Cell Sci. 112:2241-2251.

51

66. Heinegard, D. and Oldberg, A. (1989) Structure and biology of cartilage and bone matrix

noncollagenous macromolecules. Faseb J. 3:2042-2051.

67. Heldin, C.H. and Westermark, B. (1999) Mechanism of action and in vivo role of platelet-

derived growth factor. Physiol Rev. 79:1283-1316.

68. Heldin, P., Asplund, T., Ytterberg, D., Thelin, S. and Laurent, T.C. (1992) Characterization of

the molecular mechanism involved in the activation of hyaluronan synthetase by platelet-

derived growth factor in human mesothelial cells. Biochem J. 283:165-170.

69. Heldin, P. and Pertoft, H. (1993) Synthesis and assembly of the hyaluronan-containing coats

around normal human mesothelial cells. Exp Cell Res. 208:422-429.

70. Heldin, P., Suzuki, M., Teder, P. and Pertoft, H. (1995) Chondroitin sulfate proteoglycan

modulates the permeability of hyaluronan-containing coats around normal human mesothelial

cells. J Cell Physiol. 165:54-61.

71. Heldin, P., Torres, M., Ytterberg, D. and Bergh, J. (1996) Differential synthesis and binding

of hyaluronan by human breast cancer cell lines: Relationship to hormone receptor status.

Oncology reports. 3:1011-1016.

72. Henke, C.A., Roongta, U., Mickelson, D.J., Knutson, J.R. and McCarthy, J.B. (1996) CD44-

related chondroitin sulfate proteoglycan, a cell surface receptor implicated with tumor cell

invasion, mediates endothelial cell migration on fibrinogen and invasion into a fibrin matrix. J

Clin Invest. 97:2541-2552.

73. Hirschberg, C.B., Robbins, P.W. and Abeijon, C. (1998) Transporters of nucleotide sugars,

ATP, and nucleotide sulfate in the endoplasmic reticulum and Golgi apparatus. Annu Rev

Biochem. 67:49-69.

74. Hirschberg, C.B. and Snider, M.D. (1987) Topography of glycosylation in the rough

endoplasmic reticulum and Golgi apparatus. Annu Rev Biochem. 56:63-87.

75. Hofmann, M., Assmann, V., Fieber, C., Sleeman, J.P., Moll, J., Ponta, H., Hart, I.R. and

Herrlich, P. (1998) Problems with RHAMM: a new link between surface adhesion and

oncogenesis? Cell. 95:591-593.

76. Hofmann, M., Fieber, C., Assmann, V., Gottlicher, M., Sleeman, J., Plug, R., Howells, N., von

Stein, O., Ponta, H. and Herrlich, P. (1998) Identification of IHABP, a 95 kDa intracellular

hyaluronate binding protein. J Cell Sci. 111:1673-1684.

77. Holmer, E., Lindahl, U., Backstrom, G., Thunberg, L., Sandberg, H., Soderstrom, G. and

Anderson, L.O. (1980) Anticoagulant activities and effects on platelets of a heparin fragment

with high affinity for antithrombin. Thromb Res. 18:861-869.

78. Honda, A., Noguchi, N., Takehara, H., Ohashi, Y., Asuwa, N. and Mori, Y. (1991)

Cooperative enhancement of hyaluronic acid synthesis by combined use of IGF-I and EGF,

52

and inhibition by tyrosine kinase inhibitor genistein, in cultured mesothelial cells from rabbit

pericardial cavity. J Cell Sci. 98:91-98.

79. Honda, A., Sekiguchi, Y. and Mori, Y. (1993) Prostaglandin E2 stimulates cyclic AMP-

mediated hyaluronan synthesis in rabbit pericardial mesothelial cells. Biochem J. 292:497-

502.

80. Horton, M.R., McKee, C.M., Bao, C., Liao, F., Farber, J.M., Hodge-DuFour, J., Pure, E.,

Oliver, B.L., Wright, T.M. and Noble, P.W. (1998) Hyaluronan fragments synergize with

interferon-gamma to induce the C-X-C chemokines mig and interferon-inducible protein-10

in mouse macrophages. J Biol Chem. 273:35088-35094.

81. Hua, Q., Knudson, C.B. and Knudson, W. (1993) Internalization of hyaluronan by

chondrocytes occurs via receptor-mediated endocytosis. J Cell Sci. 106:365-375.

82. Iozzo, R.V. (1998) Matrix proteoglycans: from molecular design to cellular function. Annu

Rev Biochem. 67:609-652.

83. Iozzo, R.V. and Murdoch, A.D. (1996) Proteoglycans of the extracellular environment: clues

from the gene and protein side offer novel perspectives in molecular diversity and function.

Faseb J. 10:598-614.

84. Ishihara, M., Guo, Y., Wei, Z., Yang, Z., Swiedler, S.J., Orellana, A. and Hirschberg, C.B.

(1993) Regulation of biosynthesis of the basic fibroblast growth factor binding domains of

heparan sulfate by heparan sulfate-N-deacetylase/N-sulfotransferase expression. J Biol

Chem. 268:20091-20095.

85. Itano, N. and Kimata, K. (1996) Expression cloning and molecular characterization of HAS

protein, a eukaryotic hyaluronan synthase. J Biol Chem. 271:9875-9878.

86. Itano, N., Sawai, T., Yoshida, M., Lenas, P., Yamada, Y., Imagawa, M., Shinomura, T.,

Hamaguchi, M., Yoshida, Y., Ohnuki, Y., Miyauchi, S., Spicer, A.P., McDonald, J.A. and

Kimata, K. (1999) Three isoforms of mammalian hyaluronan synthases have distinct

enzymatic properties. J Biol Chem. 274:25085-25092.

87. Jackson, D.G., Buckley, J. and Bell, J.I. (1992) Multiple variants of the human lymphocyte

homing receptor CD44 generated by insertions at a single site in the extracellular domain. J

Biol Chem. 267:4732-4739.

88. Jakoby, W.B. (1994) Detoxication: Conjugation and Hydrolysis. In "The Liver: Biology and

Pathobiology". Arias, I.M., Boyer, J.L., Fausto, N., Jakoby, W.B., Schachter, D.A. and

Shafritz, D.A., editors. Raven Press Ltd, New York. 429-440.

89. Jalkanen, S. and Jalkanen, M. (1992) Lymphocyte CD44 binds the COOH-terminal heparin-

binding domain of fibronectin. J Cell Biol. 116:817-825.

53

90. Jenniskens, G.J., Oosterhof, A., Brandwijk, R., Veerkamp, J.H. and van Kuppevelt, T.H.

(2000) Heparan sulfate heterogeneity in skeletal muscle basal lamina: demonstration by phage

display-derived antibodies. J Neurosci. 20:4099-4111.

91. Kalomiris, E.L. and Bourguignon, L.Y. (1988) Mouse T lymphoma cells contain a

transmembrane glycoprotein (GP85) that binds ankyrin. J Cell Biol. 106:319-327.

92. Katoh, S., Zheng, Z., Oritani, K., Shimozato, T. and Kincade, P.W. (1995) Glycosylation of

CD44 negatively regulates its recognition of hyaluronan. J Exp Med. 182:419-429.

93. Kearns, A.E., Vertel, B.M. and Schwartz, N.B. (1993) Topography of glycosylation and

UDP-xylose production. J Biol Chem. 268:11097-11104.

94. Kimura, J.H., Hardingham, T.E., Hascall, V.C. and Solursh, M. (1979) Biosynthesis of

proteoglycans and their assembly into aggregates in cultures of chondrocytes from the Swarm

rat chondrosarcoma. J Biol Chem. 254:2600-2609.

95. Kincade, P.W., Zheng, Z., Katoh, S. and Hanson, L. (1997) The importance of cellular

environment to function of the CD44 matrix receptor. Curr Opin Cell Biol. 9:635-642.

96. Kjellen, L. and Lindahl, U. (1991) Proteoglycans: structures and interactions. Annu Rev

Biochem. 60:443-475.

97. Knudson, C.B. and Knudson, W. (1993) Hyaluronan-binding proteins in development, tissue

homeostasis, and disease. Faseb J. 7:1233-1241.

98. Knudson, W., Bartnik, E. and Knudson, C.B. (1993) Assembly of pericellular matrices by

COS-7 cells transfected with CD44 lymphocyte-homing receptor genes. Proc Natl Acad Sci

U S A. 90:4003-4007.

99. Knudson, W. and Knudson, C.B. (1991) Assembly of a chondrocyte-like pericellular matrix

on non-chondrogenic cells. Role of the cell surface hyaluronan receptors in the assembly of a

pericellular matrix. J Cell Sci. 99:227-235.

100. Kohda, D., Morton, C.J., Parkar, A.A., Hatanaka, H., Inagaki, F.M., Campbell, I.D. and Day,

A.J. (1996) Solution structure of the link module: a hyaluronan-binding domain involved in

extracellular matrix stability and cell migration. Cell. 86:767-775.

101. Kok, K., Osinga, J., Carritt, B., Davis, M.B., van der Hout, A.H., van der Veen, A.Y.,

Landsvater, R.M., de Leij, L.F., Berendsen, H.H., Postmus, P.E. and et al. (1987) Deletion of

a DNA sequence at the chromosomal region 3p21 in all major types of lung cancer. Nature.

330:578-581.

102. Kreil, G. (1995) Hyaluronidases--a group of neglected enzymes. Protein Sci. 4:1666-1669.

103. Landolt, R.M., Vaughan, L., Winterhalter, K.H. and Zimmermann, D.R. (1995) Versican is

selectively expressed in embryonic tissues that act as barriers to neural crest cell migration

and axon outgrowth. Development. 121:2303-2312.

54

104. Landstein, D., Graves, M.V., Burbank, D.E., DeAngelis, P. and Van Etten, J.L. (1998)

Chlorella virus PBCV-1 encodes functional glutamine: fructose-6-phosphate

amidotransferase and UDP-glucose dehydrogenase enzymes. Virology. 250:388-396.

105. Lauffenburger, D.A. and Horwitz, A.F. (1996) Cell migration: a physically integrated

molecular process. Cell. 84:359-369.

106. Laurent, T.C. and Fraser, J.R. (1992) Hyaluronan. Faseb J. 6:2397-2404.

107. Laurent, T.C., Fraser, J.R., Pertoft, H. and Smedsrod, B. (1986) Binding of hyaluronate and

chondroitin sulphate to liver endothelial cells. Biochem J. 234:653-658.

108. Laurent, T.C. and Fraser, J.R.E. (1991) Catabolism of Hyaluronan. In "Degradation of

Bioactive Substances: Physiology and Pathology". Henriksen, J.H., editor. CRC Press, Boca

Raton. 249-265.

109. LeBaron, R.G., Zimmermann, D.R. and Ruoslahti, E. (1992) Hyaluronate binding properties

of versican. J Biol Chem. 267:10003-10010.

110. Lepperdinger, G., Strobl, B. and Kreil, G. (1998) HYAL2, a human gene expressed in many

cells, encodes a lysosomal hyaluronidase with a novel type of specificity. J Biol Chem.

273:22466-22470.

111. Lesley, J., He, Q., Miyake, K., Hamann, A., Hyman, R. and Kincade, P.W. (1992)

Requirements for hyaluronic acid binding by CD44: a role for the cytoplasmic domain and

activation by antibody. J Exp Med. 175:257-266.

112. Lesley, J. and Hyman, R. (1998) CD44 structure and function [In Process Citation]. Front

Biosci. 3:D616-630.

113. Lesley, J., Hyman, R. and Kincade, P.W. (1993) CD44 and its interaction with extracellular

matrix. Adv Immunol. 54:271-335.

114. Lin, Y., Mahan, K., Lathrop, W.F., Myles, D.G. and Primakoff, P. (1994) A hyaluronidase

activity of the sperm plasma membrane protein PH-20 enables sperm to penetrate the cumulus

cell layer surrounding the egg. J Cell Biol. 125:1157-1163.

115. Lindahl, U., Kusche-Gullberg, M. and Kjellen, L. (1998) Regulated diversity of heparan

sulfate. J Biol Chem. 273:24979-24982.

116. Liu, D. and Sy, M.S. (1997) Phorbol myristate acetate stimulates the dimerization of CD44

involving a cysteine in the transmembrane domain. J Immunol. 159:2702-2711.

117. Lloyd, J.B. (1996) Metabolite efflux and influx across the lysosome membrane. Subcell

Biochem. 27:361-386.

118. Maiti, A., Maki, G. and Johnson, P. (1998) TNF-alpha induction of CD44-mediated

leukocyte adhesion by sulfation. Science. 282:941-943.

55

119. Markovitz, A., Cifonelli, J.A. and Dorfman, A. (1959) The biosynthesis of HA by group A

Streptococcus. VI. Biosynthesis from uridine nucleotides in cell-free extracts. J Biol Chem.

234:2343-2350.

120. McCourt, P.A., Smedsrod, B.H., Melkko, J. and Johansson, S. (1999) Characterization of a

hyaluronan receptor on rat sinusoidal liver endothelial cells and its functional relationship to

scavenger receptors. Hepatology. 30:1276-1286.

121. McKee, C.M., Lowenstein, C.J., Horton, M.R., Wu, J., Bao, C., Chin, B.Y., Choi, A.M. and

Noble, P.W. (1997) Hyaluronan fragments induce nitric-oxide synthase in murine

macrophages through a nuclear factor kappaB-dependent mechanism. J Biol Chem.

272:8013-8018.

122. Meyer, F.A. (1983) Macromolecular basis of globular protein exclusion and of swelling

pressure in loose connective tissue (umbilical cord). Biochim Biophys Acta. 755:388-399.

123. Meyer, K. and Palmer, J.W. (1934) The polysaccharide of the vitreous humor. J Biol Chem.

107:629-634.

124. Meyer, M.F. and Kreil, G. (1996) Cells expressing the DG42 gene from early Xenopus

embryos synthesize hyaluronan. Proc Natl Acad Sci U S A. 93:4543-4547.

125. Mian, N. (1986) Characterization of a high-Mr plasma-membrane-bound protein and

assessment of its role as a constituent of hyaluronate synthase complex. Biochem J. 237:343-

357.

126. Milev, P., Maurel, P., Haring, M., Margolis, R.K. and Margolis, R.U. (1996) TAG-1/axonin-1

is a high-affinity ligand of neurocan, phosphacan/protein-tyrosine phosphatase-zeta/beta, and

N-CAM. J Biol Chem. 271:15716-15723.

127. Milla, M.E., Clairmont, C.A. and Hirschberg, C.B. (1992) Reconstitution into

proteoliposomes and partial purification of the Golgi apparatus membrane UDP-galactose,

UDP-xylose, and UDP-glucuronic acid transport activities. J Biol Chem. 267:103-107.

128. Milla, M.E. and Hirschberg, C.B. (1989) Reconstitution of Golgi vesicle CMP-sialic acid and

adenosine 3'-phosphate 5'-phosphosulfate transport into proteoliposomes. Proc Natl Acad Sci

U S A. 86:1786-1790.

129. Moll, J., Khaldoyanidi, S., Sleeman, J.P., Achtnich, M., Preuss, I., Ponta, H. and Herrlich, P.

(1998) Two different functions for CD44 proteins in human myelopoiesis. J Clin Invest.

102:1024-1034.

130. Neame, S.J. and Isacke, C.M. (1992) Phosphorylation of CD44 in vivo requires both Ser323

and Ser325, but does not regulate membrane localization or cytoskeletal interaction in

epithelial cells. Embo J. 11:4733-4738.

56

131. Neame, S.J. and Isacke, C.M. (1993) The cytoplasmic tail of CD44 is required for basolateral

localization in epithelial MDCK cells but does not mediate association with the detergent-

insoluble cytoskeleton of fibroblasts. J Cell Biol. 121:1299-1310.

132. Neufeld, E. and Hall, C. (1965) Inhibition of UDP-D-glucos dehydrogenase by UDP-D-

xylose. Biochem Biophys Res Commun. 19:456-461.

133. Ng, K.F. and Schwartz, N.B. (1989) Solubilization and partial purification of hyaluronate

synthetase from oligodendroglioma cells. J Biol Chem. 264:11776-11783.

134. Ohashi, Y., Honda, A., Iwai, T. and Mori, Y. (1988) Stimulatory effect of vanadate on

hyaluronic acid synthesis in mesothelial cells from rabbit pericardium. Biochem Int. 16:293-

302.

135. Orkin, R.W. and Toole, B.P. (1978) Hyaluronidase activity and hyaluronate content of the

developing chick embryo heart. Dev Biol. 66:308-320.

136. Parkar, A.A. and Day, A.J. (1997) Overlapping sites on the Link module of human TSG-6

mediate binding to hyaluronan and chrondroitin-4-sulphate. FEBS Lett. 410:413-417.

137. Peach, R.J., Hollenbaugh, D., Stamenkovic, I. and Aruffo, A. (1993) Identification of

hyaluronic acid binding sites in the extracellular domain of CD44. J Cell Biol. 122:257-264.

138. Peck, D. and Isacke, C.M. (1996) CD44 phosphorylation regulates melanoma cell and

fibroblast migration on, but not attachment to, a hyaluronan substratum. Curr Biol. 6:884-890.

139. Perozich, J., Leksana, A. and Hempel, J. (1995) UDP-glucose dehydrogenase. Structural

characteristics. Adv Exp Med Biol. 372:79-84.

140. Perris, R. and Johansson, S. (1990) Inhibition of neural crest cell migration by aggregating

chondroitin sulfate proteoglycans is mediated by their hyaluronan-binding region. Dev Biol.

137:1-12.

141. Perris, R., Perissinotto, D., Pettway, Z., Bronner-Fraser, M., Morgelin, M. and Kimata, K.

(1996) Inhibitory effects of PG-H/aggrecan and PG-M/versican on avian neural crest cell

migration. Faseb J. 10:293-301.

142. Philipson, L.H. and Schwartz, N.B. (1984) Subcellular localization of hyaluronate synthetase

in oligodendroglioma cells. J Biol Chem. 259:5017-5023.

143. Piek, E., Heldin, C.H. and Ten Dijke, P. (1999) Specificity, diversity, and regulation in TGF-

beta superfamily signaling. Faseb J. 13:2105-2124.

144. Prehm, P. (1983) Synthesis of hyaluronate in differentiated teratocarcinoma cells.

Characterization of the synthase. Biochem J. 211:181-189.

145. Prehm, P. (1983) Synthesis of hyaluronate in differentiated teratocarcinoma cells. Mechanism

of chain growth. Biochem J. 211:191-198.

57

146. Prehm, P. (1989) Identification and regulation of the eukaryotic hyaluronate synthase. In

"The Biology of Hyaluronan". Vol. 143. Evered, D. and Whelan, J., editors. Wiley,

Chichester. 21-31.

147. Pummill, P.E., Achyuthan, A.M. and DeAngelis, P.L. (1998) Enzymological characterization

of recombinant xenopus DG42, a vertebrate hyaluronan synthase. J Biol Chem. 273:4976-

4981.

148. Rahmanian, M., Pertoft, H., Kanda, S., Christofferson, R., Claesson-Welsh, L. and Heldin, P.

(1997) Hyaluronan oligosaccharides induce tube formation of a brain endothelial cell line in

vitro. Exp Cell Res. 237:223-230.

149. Rapraeger, A.C., Krufka, A. and Olwin, B.B. (1991) Requirement of heparan sulfate for

bFGF-mediated fibroblast growth and myoblast differentiation. Science. 252:1705-1708.

150. Rauch, U., Karthikeyan, L., Maurel, P., Margolis, R.U. and Margolis, R.K. (1992) Cloning

and primary structure of neurocan, a developmentally regulated, aggregating chondroitin

sulfate proteoglycan of brain. J Biol Chem. 267:19536-19547.

151. Reed, R.K., Laurent, U.B., Fraser, J.R. and Laurent, T.C. (1990) Removal rate of

[3H]hyaluronan injected subcutaneously in rabbits. Am J Physiol. 259:H532-535.

152. Reed, R.K., Lilja, K. and Laurent, T.C. (1988) Hyaluronan in the rat with special reference to

the skin. Acta Physiol Scand. 134:405-411.

153. Ripellino, J.A., Bailo, M., Margolis, R.U. and Margolis, R.K. (1988) Light and electron

microscopic studies on the localization of hyaluronic acid in developing rat cerebellum. J Cell

Biol. 106:845-855.

154. Robertson, D., Beech, I. and Bolwell, G.P. (1995) Regulation of the Enzymes of UDP-sugar

Metabolism during Differentiation of French Bean. Phytochemistry. 39:21-28.

155. Robyt, J.F., Kimble, B.K. and Walseth, T.F. (1974) The mechanism of dextransucrase action.

Direction of dextran biosynthesis. Arch Biochem Biophys. 165:634-640.

156. Roden, L., Koerner, T., Olson, C. and Schwartz, N.B. (1985) Mechanisms of chain initiation

in the biosynthesis of connective tissue polysaccharides. Fed Proc. 44:373-380.

157. Rome, L.H. and Hill, D.F. (1986) Lysosomal degradation of glycoproteins and

glycosaminoglycans. Efflux and recycling of sulphate and N-acetylhexosamines. Biochem J.

235:707-713.

158. Rooney, P., Kumar, S., Ponting, J. and Wang, M. (1995) The role of hyaluronan in tumour

neovascularization (review). Int J Cancer. 60:632-636.

159. Ruoslahti, E. and Yamaguchi, Y. (1991) Proteoglycans as modulators of growth factor

activities. Cell. 64:867-869.

160. Sainio, M., Zhao, F., Heiska, L., Turunen, O., den Bakker, M., Zwarthoff, E., Lutchman,

M., Rouleau, G.A., Jaaskelainen, J., Vaheri, A. and Carpen, O. (1997) Neurofibromatosis

58

2 tumor suppressor protein colocalizes with ezrin and CD44 and associates with actin-

containing cytoskeleton. J Cell Sci. 110:2249-2260.

161. Salmivirta, M., Lidholt, K. and Lindahl, U. (1996) Heparan sulfate: a piece of information.

Faseb J. 10:1270-1279.

162. Salustri, A., Yanagishita, M. and Hascall, V.C. (1990) Mouse oocytes regulate hyaluronic acid

synthesis and mucification by FSH-stimulated cumulus cells. Dev Biol. 138:26-32.

163. Salustri, A., Yanagishita, M., Underhill, C.B., Laurent, T.C. and Hascall, V.C. (1992)

Localization and synthesis of hyaluronic acid in the cumulus cells and mural granulosa cells

of the preovulatory follicle. Dev Biol. 151:541-551.

164. Sargent, T.D. and Dawid, I.B. (1983) Differential gene expression in the gastrula of Xenopus

laevis. Science. 222:135-139.

165. Scott, J.E. (1989) Secondary structures in hyaluronan solutions: chemical and biological

implications. In "The Biology of Hyaluronan". Vol. 143. Evered, D. and Whelan, J., editors.

Wiley, Chichester. 6-20.

166. Scott, J.E., Cummings, C., Brass, A. and Chen, Y. (1991) Secondary and tertiary structures of

hyaluronan in aqueous solution, investigated by rotary shadowing-electron microscopy and

computer simulation. Hyaluronan is a very efficient network-forming polymer. Biochem J.

274:699-705.

167. Scott, J.E. and Heatley, F. (1999) Hyaluronan forms specific stable tertiary structures in

aqueous solution: a 13C NMR study. Proc Natl Acad Sci U S A. 96:4850-4855.

168. Seidenbecher, C.I., Richter, K., Rauch, U., Fassler, R., Garner, C.C. and Gundelfinger, E.D.

(1995) Brevican, a chondroitin sulfate proteoglycan of rat brain, occurs as secreted and cell

surface glycosylphosphatidylinositol-anchored isoforms. J Biol Chem. 270:27206-27212.

169. Semino, C.E., Specht, C.A., Raimondi, A. and Robbins, P.W. (1996) Homologs of the

Xenopus developmental gene DG42 are present in zebrafish and mouse and are involved in

the synthesis of Nod-like chitin oligosaccharides during early embryogenesis. Proc Natl Acad

Sci U S A. 93:4548-4553.

170. Sheetz, M.P., Felsenfeld, D.P. and Galbraith, C.G. (1998) Cell migration: regulation of force

on extracellular-matrix-integrin complexes. Trends Cell Biol. 8:51-54.

171. Sherman, L., Sleeman, J., Herrlich, P. and Ponta, H. (1994) Hyaluronate receptors: key

players in growth, differentiation, migration and tumor progression. Curr. Opin. Cell Biol.

6:726-733.

172. Shyjan, A.M., Heldin, P., Butcher, E.C., Yoshino, T. and Briskin, M.J. (1996) Functional

cloning of the cDNA for a human hyaluronan synthase. J Biol Chem. 271:23395-23399.

59

173. Sleeman, J., Rudy, W., Hofmann, M., Moll, J., Herrlich, P. and Ponta, H. (1996) Regulated

clustering of variant CD44 proteins increases their hyaluronate binding capacity. J Cell Biol.

135:1139-1150.

174. Sleeman, J.P., Kondo, K., Moll, J., Ponta, H. and Herrlich, P. (1997) Variant exons v6 and v7

together expand the repertoire of glycosaminoglycans bound by CD44. J Biol Chem.

272:31837-31844.

175. Smaglik, P. and Butler, D. (2000) Celera turns to public genome data to speed up endgame.

Nature. 403:119.

176. Sommarin, Y. and Heinegard, D. (1983) Specific interaction between cartilage proteoglycans

and hyaluronic acid at the chondrocyte cell surface. Biochem J. 214:777-784.

177. Spicer, A.P., Augustine, M.L. and McDonald, J.A. (1996) Molecular cloning and

characterization of a putative mouse hyaluronan synthase. J Biol Chem. 271:23400-23406.

178. Spicer, A.P., Camenisch, T., Nguyen, T., Haudenschild, D. and McDonald, J. (1999)

Hyaluronan biosynthesis and function in vertebrates. Conference abstract in "Glyco XV",

Tokyo, Japan.

179. Spicer, A.P., Kaback, L.A., Smith, T.J. and Seldin, M.F. (1998) Molecular cloning and

characterization of the human and mouse UDP- glucose dehydrogenase genes. J Biol Chem.

273:25117-25124.

180. Spicer, A.P. and McDonald, J.A. (1998) Characterization and molecular evolution of a

vertebrate hyaluronan synthase gene family. J Biol Chem. 273:1923-1932.

181. Spicer, A.P. and Nguyen, T.K. (1999) Mammalian hyaluronan synthases: investigation of

functional relationships in vivo. Biochem Soc Trans. 27:109-115.

182. Spicer, A.P., Olson, J.S. and McDonald, J.A. (1997) Molecular cloning and characterization

of a cDNA encoding the third putative mammalian hyaluronan synthase. J Biol Chem.

272:8957-8961.

183. Spicer, A.P., Seldin, M.F., Olsen, A.S., Brown, N., Wells, D.E., Doggett, N.A., Itano, N.,

Kimata, K., Inazawa, J. and McDonald, J.A. (1997) Chromosomal localization of the human

and mouse hyaluronan synthase genes. Genomics. 41:493-497.

184. Stern, R. and Csoka, A. (2000) Mammalian Hyaluronidases. Glycoforum Website "Science

of Hyaluronan Today" <http://www.glycoforum.gr.jp>.

185. Stoolmiller, A.C. and Dorfman, A. (1969) The biosynthesis of hyaluronic acid by

Streptococcus. J Biol Chem. 244:236-246.

186. Suzuki, M., Asplund, T., Yamashita, H., Heldin, C.-H. and Heldin, P. (1995) Stimulation of

hyaluronan biosynthesis by platelet-derived growth factor-BB and transforming growth

factor-β1 involves activation of protein kinase C. Biochem J. 307:817-821.

60

187. Sweeney, C., Mackintosh, D. and Mason, R.M. (1993) UDP-sugar metabolism in Swarm rat

chondrosarcoma chondrocytes. Biochem J. 290:563-570.

188. Takahashi, K., Stamenkovic, I., Cutler, M., Dasgupta, A. and Tanabe, K.K. (1996) Keratan

sulfate modification of CD44 modulates adhesion to hyaluronate. J Biol Chem. 271:9490-

9496.

189. Tammi, R., MacCallum, D., Hascall, V.C., Pienimaki, J.P., Hyttinen, M. and Tammi, M.

(1998) Hyaluronan bound to CD44 on keratinocytes is displaced by hyaluronan

decasaccharides and not hexasaccharides. J Biol Chem. 273:28878-28888.

190. Teder, P., Nettelbladt, O. and Heldin, P. (1995) Characterization of the mechanism involved in

bleomycin-induced increased hyaluronan production in rat lung. Am J Respir Cell Mol Biol.

12:181-189.

191. Tengblad, A., Laurent, U.B., Lilja, K., Cahill, R.N., Engstrom-Laurent, A., Fraser, J.R.,

Hansson, H.E. and Laurent, T.C. (1986) Concentration and relative molecular mass of

hyaluronate in lymph and blood. Biochem J. 236:521-525.

192. Tenhaken, K. and Thulke, O. (1996) Cloning of an enzyme that synthesize a key nucleotide-

sugar precursor of hemicellulose biosynthesis from soybean: UDP-glucose dehydrogenase.

Plant Phys. 112:1127-1134.

193. Thiebot, B., Bichoualne, L., Langris, M., Bonnamy, P.J., Barbey, P., Carreau, S. and Bocquet,

J. (1997) IGF-1 stimulates synthesis of undersulfated proteoglycans and of hyaluronic acid

by peritubular cells from immature rat testis. Biochim Biophys Acta. 1358:127-141.

194. Thomas, L., Byers, H.R., Vink, J. and Stamenkovic, I. (1992) CD44H regulates tumor cell

migration on hyaluronate-coated substrate. J Cell Biol. 118:971-977.

195. Tiedemann, K., Malmstrom, A. and Westergren-Thorsson, G. (1997) Cytokine regulation of

proteoglycan production in fibroblasts: separate and synergistic effects. Matrix Biol. 15:469-

478.

196. Tirone, E., D'Alessandris, C., Hascall, V.C., Siracusa, G. and Salustri, A. (1997) Hyaluronan

synthesis by mouse cumulus cells is regulated by interactions between follicle-stimulating

hormone (or epidermal growth factor) and a soluble oocyte factor (or transforming growth

factor beta1). J Biol Chem. 272:4787-4794.

197. Tlapak-Simmons, V., Baggenstoss, B.A., Clyne, T. and Weigel, P. (1999) Purification and

lipid dependence of the recombinant hyaluronan synthases from Streptococcus pyogenes and

streptococcus equisimilis. J Biol Chem. 274:4239-4245.

198. Tlapak-Simmons, V.L., Baggenstoss, B.A., Kumari, K., Heldermon, C. and Weigel, P.H.

(1999) Kinetic characterization of the recombinant hyaluronan synthases from Streptococcus

pyogenes and Streptococcus equisimilis. J Biol Chem. 274:4246-4253.

61

199. Tlapak-Simmons, V.L., Kempner, E.S., Baggenstoss, B.A. and Weigel, P.H. (1998) The

active streptococcal hyaluronan synthases (HASs) contain a single HAS monomer and

multiple cardiolipin molecules. J Biol Chem. 273:26100-26109.

200. Tolg, C., Hofmann, M., Herrlich, P. and Ponta, H. (1993) Splicing choice from ten variant

exons establishes CD44 variability. Nucleic Acids Res. 21:1225-1229.

201. Toma, L., Pinhal, M.A., Dietrich, C.P., Nader, H.B. and Hirschberg, C.B. (1996) Transport of

UDP-galactose into the Golgi lumen regulates the biosynthesis of proteoglycans. J Biol

Chem. 271:3897-3901.

202. Toole, B.P. (1982) Developmental role of hyaluronate. Connect Tissue Res. 10:93-100.

203. Toole, B.P., Biswas, C. and Gross, J. (1979) Hyaluronate and invasiveness of the rabbit V2

carcinoma. Proc Natl Acad Sci U S A. 76:6299-6303.

204. Toole, B.P. and Trelstad, R.L. (1971) Hyaluronate production and removal during corneal

development in the chick. Dev Biol. 26:28-35.

205. Tsifrina, E., Ananyeva, N.M., Hastings, G. and Liau, G. (1999) Identification and

characterization of three cDNAs that encode putative novel hyaluronan-binding proteins,

including an endothelial cell-specific hyaluronan receptor. Am J Pathol. 155:1625-1633.

206. Tsukita, S., Oishi, K., Sato, N., Sagara, J. and Kawai, A. (1994) ERM family members as

molecular linkers between the cell surface glycoprotein CD44 and actin-based cytoskeletons.

J Cell Biol. 126:391-401.

207. Turley, E.A., Austen, L., Vandeligt, K. and Clary, C. (1991) Hyaluronan and a cell-associated

hyaluronan binding protein regulate the locomotion of ras-transformed cells. J Cell Biol.

112:1041-1047.

208. Uff, C.R., Neame, S.J. and Isacke, C.M. (1995) Hyaluronan binding by CD44 is regulated by

a phosphorylation-independent mechanism. Eur J Immunol. 25:1883-1887.

209. Ujita, M., Shinomura, T., Ito, K., Kitagawa, Y. and Kimata, K. (1994) Expression and binding

activity of the carboxyl-terminal portion of the core protein of PG-M, a large chondroitin

sulfate proteoglycan. J Biol Chem. 269:27603-27609.

210. Underhill, C.B., Chi-Rosso, G. and Toole, B.P. (1983) Effects of detergent solubilization on

the hyaluronate-binding protein from membranes of simian virus 40-transformed 3T3 cells. J

Biol Chem. 258:8086-8091.

211. van de Rijn, I. and Drake, R.R. (1992) Analysis of the streptococcal hyaluronic acid synthase

complex using the photoaffinity probe 5-azido-UDP-glucuronic acid. J Biol Chem.

267:24302-24306.

212. van Kuppevelt, T.H., Dennissen, M.A., van Venrooij, W.J., Hoet, R.M. and Veerkamp, J.H.

(1998) Generation and application of type-specific anti-heparan sulfate antibodies using

62

phage display technology. Further evidence for heparan sulfate heterogeneity in the kidney. J

Biol Chem. 273:12960-12966.

213. Varki, A. (1996) Does DG42 synthesize hyaluronan or chitin?: A controversy about

oligosaccharides in vertebrate development. Proc Natl Acad Sci U S A. 93:4523-4525.

214. Varki, A. (1998) Factors controlling the glycosylation potential of the Golgi apparatus.

Trends Cell Biol. 8:34-40.

215. Vertel, B.M., Walters, L.M., Flay, N., Kearns, A.E. and Schwartz, N.B. (1993) Xylosylation is

an endoplasmic reticulum to Golgi event. J Biol Chem. 268:11105-11112.

216. von Andrian, U.H., Hasslen, S.R., Nelson, R.D., Erlandsen, S.L. and Butcher, E.C. (1995) A

central role for microvillous receptor presentation in leukocyte adhesion under flow. Cell.

82:989-999.

217. Wang, C., Entwistle, J., Hou, G., Li, Q. and Turley, E.A. (1996) The characterization of a

human RHAMM cDNA: conservation of the hyaluronan-binding domains. Gene. 174:299-

306.

218. Watanabe, E., Aono, S., Matsui, F., Yamada, Y., Naruse, I. and Oohira, A. (1995) Distribution

of a brain-specific proteoglycan, neurocan, and the corresponding mRNA during the

formation of barrels in the rat somatosensory cortex. Eur J Neurosci. 7:547-554.

219. Watanabe, K. and Yamaguchi, Y. (1996) Molecular identification of a putative human

hyaluronan synthase. J Biol Chem. 271:22945-22948.

220. Weigel, P.H., Hascall, V.C. and Tammi, M. (1997) Hyaluronan synthases. J Biol Chem.

272:13997-14000.

221. West, D.C., Hampson, I.N., Arnold, F. and Kumar, S. (1985) Angiogenesis induced by

degradation products of hyaluronic acid. Science. 228:1324-1326.

222. Yamada, H., Fredette, B., Shitara, K., Hagihara, K., Miura, R., Ranscht, B., Stallcup, W.B. and

Yamaguchi, Y. (1997) The brain chondroitin sulfate proteoglycan brevican associates with

astrocytes ensheathing cerebellar glomeruli and inhibits neurite outgrowth from granule

neurons. J Neurosci. 17:7784-7795.

223. Yamada, Y., Itano, N., Zako, M., Yoshida, M., Lenas, P., Niimi, A., Ueda, M. and Kimata, K.

(1998) The gene structure and promoter sequence of mouse hyaluronan synthase 1. Biochem

J. 330:1223-1227.

224. Yamaguchi, Y. (1996) Brevican: a major proteoglycan in adult brain. Perspect Dev Neurobiol.

3:307-317.

225. Yang, B., Yang, B.L., Savani, R.C. and Turley, E.A. (1994) Identification of a common

hyaluronan binding motif in the hyaluronan binding proteins RHAMM, CD44 and link

protein. Embo J. 13:286-296.

63

226. Yang, B., Zhang, L. and Turley, E.A. (1993) Identification of two hyaluronan-binding

domains in the hyaluronan receptor RHAMM. J Biol Chem. 268:8617-8623.

227. Yannariello-Brown, J., Frost, S.J. and Weigel, P.H. (1992) Identification of the Ca(2+)-

independent endocytic hyaluronan receptor in rat liver sinusoidal endothelial cells using a

photoaffinity cross-linking reagent. J Biol Chem. 267:20451-20456.

228. Yayon, A., Klagsbrun, M., Esko, J.D., Leder, P. and Ornitz, D.M. (1991) Cell surface,

heparin-like molecules are required for binding of basic fibroblast growth factor to its high

affinity receptor. Cell. 64:841-848.

229. Yoshida, M., Itano, N., Yamada, Y. and Kimata, K. (2000) In vitro synthesis of hyaluronan by

a single protein derived from mouse HAS1 gene and characterization of amino acid residues

essential for the activity. J Biol Chem. 275:497-506.

230. Ysart, G.E. and Mason, R.M. (1994) Serum factors, growth factors and UDP-sugar

metabolism in bovine articular cartilage chondrocytes. Biochem J. 303:713-721.

231. Zhang, L., David, G. and Esko, J.D. (1995) Repetitive Ser-Gly sequences enhance heparan

sulfate assembly in proteoglycans. J Biol Chem. 270:27127-27135.

232. Zhang, L. and Esko, J.D. (1994) Amino acid determinants that drive heparan sulfate assembly

in a proteoglycan. J Biol Chem. 269:19295-19299.

233. Zhang, S., Chang, M.C., Zylka, D., Turley, S., Harrison, R. and Turley, E.A. (1998) The

hyaluronan receptor RHAMM regulates extracellular-regulated kinase. J Biol Chem.

273:11342-11348.

234. Zhao, M., Yoneda, M., Ohashi, Y., Kurono, S., Iwata, H., Ohnuki, Y. and Kimata, K. (1995)

Evidence for the covalent binding of SHAP, heavy chains of inter-alpha-trypsin inhibitor, to

hyaluronan. J Biol Chem. 270:26657-26663.

235. Zheng, Z., Cummings, R.D., Pummill, P.E. and Kincade, P.W. (1997) Growth as a solid

tumor or reduced glucose concentrations in culture reversibly induce CD44-mediated

hyaluronan recognition by Chinese hamster ovary cells. J Clin Invest. 100:1217-1229.

236. Zhou, B., Oka, J.A., Singh, A. and Weigel, P.H. (1999) Purification and subunit

characterization of the rat liver endocytic hyaluronan receptor. J Biol Chem. 274:33831-

33834.

237. Zoller, M. (1995) CD44: physiological expression of distinct isoforms as evidence for organ-

specific metastasis formation. J Mol Med. 73:425-438.