This article was downloaded by: [Tulane University] On: 23 January 2015, At: 07:28 Publisher: Taylor & Francis Informa Ltd Registered in England and Wales Registered Number: 1072954 Registered office: Mortimer House, 37-41 Mortimer Street, London W1T 3JH, UK

Journal of Biomolecular Structure and Dynamics Publication details, including instructions for authors and subscription information: http://www.tandfonline.com/loi/tbsd20

Cell communication using intrinsically disordered proteins: what can syndecans say? ab

Elena I. Leonova

a

& Oxana V. Galzitskaya

a

Institute of Protein Research, Russian Academy of Sciences, Moscow Region, Pushchino 142290, Russia b

Institute of Biochemistry and Physiology of Microorganisms, Russian Academy of Sciences, Moscow Region, Pushchino 142290, Russia Published online: 23 Jun 2014.

Click for updates To cite this article: Elena I. Leonova & Oxana V. Galzitskaya (2014): Cell communication using intrinsically disordered proteins: what can syndecans say?, Journal of Biomolecular Structure and Dynamics, DOI: 10.1080/07391102.2014.926256 To link to this article: http://dx.doi.org/10.1080/07391102.2014.926256

PLEASE SCROLL DOWN FOR ARTICLE Taylor & Francis makes every effort to ensure the accuracy of all the information (the “Content”) contained in the publications on our platform. However, Taylor & Francis, our agents, and our licensors make no representations or warranties whatsoever as to the accuracy, completeness, or suitability for any purpose of the Content. Any opinions and views expressed in this publication are the opinions and views of the authors, and are not the views of or endorsed by Taylor & Francis. The accuracy of the Content should not be relied upon and should be independently verified with primary sources of information. Taylor and Francis shall not be liable for any losses, actions, claims, proceedings, demands, costs, expenses, damages, and other liabilities whatsoever or howsoever caused arising directly or indirectly in connection with, in relation to or arising out of the use of the Content. This article may be used for research, teaching, and private study purposes. Any substantial or systematic reproduction, redistribution, reselling, loan, sub-licensing, systematic supply, or distribution in any form to anyone is expressly forbidden. Terms & Conditions of access and use can be found at http:// www.tandfonline.com/page/terms-and-conditions

Journal of Biomolecular Structure and Dynamics, 2014 http://dx.doi.org/10.1080/07391102.2014.926256

Cell communication using intrinsically disordered proteins: what can syndecans say? Elena I. Leonovaa,b and Oxana V. Galzitskayaa* a Institute of Protein Research, Russian Academy of Sciences, Moscow Region, Pushchino 142290, Russia; bInstitute of Biochemistry and Physiology of Microorganisms, Russian Academy of Sciences, Moscow Region, Pushchino 142290, Russia

Communicated by Ramaswamy H. Sarma.

Downloaded by [Tulane University] at 07:28 23 January 2015

(Received 14 February 2014; accepted 16 May 2014) Because intrinsically disordered proteins are incapable of forming unique tertiary structures in isolation, their interaction with partner structures enables them to play important roles in many different biological functions. Therefore, such proteins are usually multifunctional, and their ability to perform their major function, as well as accessory functions, depends on the characteristics of a given interaction. The present paper demonstrates, using predictions from two programs, that the transmembrane proteoglycans syndecans are natively disordered because of their diverse functions and large number of interaction partners. Syndecans perform multiple functions during development, damage repair, tumor growth, angiogenesis, and neurogenesis. By mediating the binding of a large number of extracellular ligands to their receptors, these proteoglycans trigger a cascade of reactions that subsequently regulate various cell processes: cytoskeleton formation, proliferation, differentiation, adhesion, and migration. The occurrences of 20 amino acids in syndecans 1–4 from 25 animals were compared with those in 17 animal proteomes. Gly + Ala, Thr, Glu, and Pro were observed to predominate in the syndecans, contributing to the lack of an ordered structure. In contrast, there were many fewer amino acids in syndecans that promote an ordered structure, such as Cys, Trp, Asn, and His. In addition, a region rich in Asp has been identified between two heparan sulfate-binding sites in the ectodomains, and a region rich in Lys has been identified in the conserved C1 site of the cytoplasmic domain. These particular regions play an essential role in the various functions of syndecans due to their lack of structure. Keywords: transmembrane proteoglycans; intrinsically disordered proteins; disordered regions; library of disordered patterns; adhesion; cell matrix

1. Introduction The traditional protein–structure paradigm describes that a protein folds into a single functional, three-dimensional structure, and that its amino acid sequence determines its unique native conformation and unique function (Anfinsen, 1973). The category of intrinsically disordered proteins (IDPs) is a notable exception to this because these proteins are unable to achieve a fixed, unique tertiary structure on their own, yet they can form such structures after binding to various partners. Therefore, the structures and functions of such proteins depend on the characteristics of the macromolecules to which they bind. Many IDPs that are known to be essential for diverse cellular functions have been discovered individually (Dyson & Wright, 2005; Tompa, 2012; Uversky & Dunker, 2010). Interestingly, many genes in many different organisms encode IDPs. For example, a large fraction of RNA-binding proteins consist of domains that lack a three-dimensional structure, and they are predicted to have prion-like domains that play a crucial role in RNP assembly but can also cause pathological aggregation (Alberti, Halfmann, King, Kapila, & *Corresponding author. Email: [email protected] © 2014 Taylor & Francis

Lindquist, 2009; Gitler & Shorter, 2011; Halfmann et al., 2011). For certain proteins, the presence of natively unstructured regions is suggested to be of functional importance. Syndecans, which are transmembrane proteoglycans, can be natively disordered due to their diverse functions and large number of interaction partners. Syndecans (meaning “to link” in Greek) were first discovered in 1989 by the group of M. Burnfield in California (Stanford University School of Medicine) during the study of the proteins that are associated with the cell surface and contain heparan sulfate and chondroitin sulfate chains (Saunders, Jalkanen, O’Farrell, & Bernfield, 1989). It has been noted that syndecans are synthesized by most of the animal cells. They consist of a core protein with covalently linked glycosaminoglycan chains (Kokenyesi & Bernfield, 1994; Weinbaum, Tarbell, & Damiano, 2007). The synthesis of the polypeptide chain of the core protein begins on membrane-bound ribosomes and proceeds in the lumen of the endoplasmatic reticulum. Binding of the glycosaminoglycan chains proceeds in the Golgi apparatus, and then the proteoglycans are delivered by exocytosis to the cell surface.

Downloaded by [Tulane University] at 07:28 23 January 2015

2

E.I. Leonova and O.V. Galzitskaya

Additionally, the polysaccharide chains in the Golgi apparatus are modified by epimerization and the addition of sulfate groups. There are four types of syndecans in vertebrates, all of which are composed of a core protein and covalently linked glycosaminoglycan chains (Alberts et al., 2008). Glycosaminoglycan chains may carry 50–200 negatively charged disaccharide units due to the attached sulfate groups. Therefore, they can hold a large number of positively charged molecules. Moreover, because of the accumulation of charge, glycosaminoglycan chains are repelled from each other and extend into space to increase the area of their interaction (Manon-Jensen, Itoh, & Couchman, 2010). The sulfation of glycosaminoglycans is not homogeneous; there are sites with a small number of sulfate groups that are distributed among other sites that have a large content of sulfate groups. Frequently, slightly sulfated sites are concentrated closer to the core protein of a syndecan (Manon-Jensen et al., 2010). The core protein, in turn, contains three types of domains: extracellular, transmembrane, and cytoplasmic (see Figure 1).

The expression of genes that encodes syndecans depends on the type of tissues: the gene of syndecan-1 is expressed predominantly in epithelial tissues, the gene of syndecan-2 in mesenchymal tissues, the gene of syndecan-3 in nervous tissues, and the gene of syndecan-4 in almost in all tissues (Kim, Goldberger, Gallo, & Bernfield, 1994). As a rule, cells express limited types of syndecans, and the expression varies, depending on the developmental stage. The intensive synthesis of syndecan-2 in the mesenchymal tissues proceeds in the late stages of morphogenesis, whereas it is almost entirely unobserved in the early stages. In contrast, the synthesis of syndecan-1 occurs consistently in the early stages and is almost absent in the later stages of morphogenesis (David et al., 1993). The expression of the gene of syndecan-1 is increased in the later stages of differentiation of immune system cells both in plasmatic cells (Kim et al., 1994) and tissue macrophages (Wang, Haller, Wen, Wang, & Chaikof, 2008). The various syndecans of vertebrates originated from only one type of syndecan of invertebrates and, after

Figure 1. Schematic illustration of vertebrate syndecans. Notes: Syndecans 1 and 3 are longer than syndecans 2 and 4. The glycosaminoglycan chains are covalently linked to a core protein: the serine is linked to xylose, which is successively attached to two galactose residues and one glucuronic acid residue. Heparan sulfate consists of N-acetylglucosamine with glucuronic acid or N-acetylglucosamine with iduronic acid. The chondroitin sulfate consists of N-acetylgalactosamine with glucuronic acid. The cytoplasmic domain has two conserved С1 and С2 sites and one variable V site that is unique for each type of syndecan. ТМ (transmembrane domains) have GxxxG motifs that are conserved for each type of syndecan. For syndecan-1, it is GGLVG; for syndecan-2, it is GGVIG; for syndecan-3, it is GGVVG; and for syndecan-4, it is GGxVG.

Downloaded by [Tulane University] at 07:28 23 January 2015

Cell communication gene duplication, formed two groups of syndecans; one of them was an ancestor of syndecans 1 and 3, and the other was an ancestor of syndecans 2 and 4 (Chakravarti & Adams, 2006; Chen, Couchman, Smith, & Woods, 2002). This explains the high homology within each group. Syndecans 2 and 4 are shorter than the others and carry only heparin sulfate chains, whereas syndecans 1 and 3 are longer and carry both heparin sulfate and chondroitin sulfate chains (Figure 1). It is known that many receptors and co-receptors, including syndecans, undergo dimerization (Heldin, 1995). Usually, this occurs in lipid rafts, the parts of a cell membrane rich in cholesterol and sphingolipids (Simons & Ikonen, 1997). Lipid rafts are crucial for receptor binding and signal transduction from the cell surface into the cell. Often, lipid rafts are floating on the cell surface and are stabilized by the cooperation of specific receptor proteins in the cell membrane. The transmembrane domain of all syndecans includes a conserved motif GxxxG that is necessary for the dimerization of syndecans and for retaining cholesterol in the membrane (Barrett et al., 2012). It is known that decreasing the cholesterol level in the lipid rafts leads to their degradation and to the damage of the signaling cascade (Li, Park, Ye, Kim, & Kim, 2006). It should be noted that syndecans vary in their tendency to form homodimers: syndecan-1 dimerizes only weakly, syndecans 3 and 4 dimerize strongly and syndecan-2 dimerizes very strongly (Dews & MacKenzie, 2007). Moreover, certain types of syndecans can form heterodimers and even heterotrimers. The exceptions are syndecans 1 and 4, which cannot form heterodimers with each other. Ligands that bind to syndecans are able to promote their dimerization and, thus, may drive few signaling cascades (Dews & MacKenzie, 2007). In organisms, syndecans can be found in two forms: membrane-incorporated and “soluble” (Bernfield et al., 1999; Wang, Götte, Bernfield, & Reizes, 2005). The “soluble” form is the ectodomain that has been shed from the cell surface by a process governed by extracellular, zinc-dependent endopeptidases and metalloproteinases (MMPs) (Figure 2). This shedding is regulated by a large number of extracellular stimulating agents: chemokines, growth factors, trypsin, heparanase, virulent components of bacteria, insulin, and cellular stress. In general, the ectodomain shedding is multiplied in response to inflammation and other destructive processes that proceed in an organism (Li, Park, Wilson, & Parks, 2002; Manon-Jensen et al., 2010). The “soluble” syndecans 1 and 4 are concentrated in large amounts in the fluid that accumulates around wounds. The shedding of the extracellular domain is increased in response to the thrombin and epidermal growth factors that are especially active during wound healing (Subramanian, Fitzgerald, & Bernfield, 1997). Certain signaling transducers, such as protein

3

kinase C and nuclear transcription factor NF-кB, also influence the shedding of ectodomains. In addition, the phosphorylation of tyrosines in the conserved sites of the cytoplasmic domain initiates the shedding of ectodomains (Manon-Jensen et al., 2010). Interestingly, tyrosine residues are absent in the conserved C1 and C2 sites in frog syndecan-4. At the present time, the detailed mechanism of the activation of ectodomain shedding by various stimulants remains obscure. Syndecan-1 plays a significant role in lipid metabolism, the regulation of leukocyte migration, wound healing, and the remodeling of the cytoskeleton of monolayer endothelial cells in response to shear stress, which induces the protection of vessel walls and prevents the development of atherosclerosis (Savery, Jiang, Park, & Damiano, 2013; Stanford et al., 2009; Vanhoutte et al., 2007). However, the overexpression of syndecan-1 in macrophages leads to the formation of atherosclerotic plaques (Asplund, Ostergren-Lunden, Camejo, Stillemark-Billton, & Bondjers, 2009; Wang, Götte, Bernfield, & Reizes, 2005). A higher level of “soluble” syndecan-1 in the blood serum may illustrate the development of multiple myeloma (Caraux et al., 2010; Kim, Lee, Cho, & Ihm, 2010). Syndecan-2 is crucial in the development of the nervous system during embryogenesis and angiogenesis (Chen, Hermanson, & Ekker, 2004). Syndecan-2 also plays an important role in the formation of amyloid plaques, which leads to the development of neurodegenerative diseases that accompany aging (Manich et al., 2011; Zhao, Liu, & Lilly, 2012). In addition, syndecan-2 increases intestinal cancer cell migration (Choi et al., 2012; Lee et al., 2011; Lin, Lei, Hong, & Hsueh, 2007). Syndecan-3 is very important for the normal functioning of the brain, appetite regulation, memory strengthening, and skeletal musculature development (Kaksonen et al., 2002; Karlsson-Lindahl et al., 2012; Pisconti, Cornelison, Olguin, Antwine, & Olwin, 2010). After knockout of the encoding gene of syndecan-3, a mouse phenotype became thin and inactive (Zheng et al., 2010). The understanding of the mechanism of regulation by “soluble” syndecan-3 that underlies the functioning of the hippocampus and the development of the brain cortex requires additional investigation. Syndecan-4 synthesized by all types of cells plays essential roles in angiogenesis, focal adhesion, decreased cell migration, and normalization of blood pressure (Baciu et al., 2000; Denhez et al., 2002; Fitzgerald, Wang, Park, Murphy, & Bernfield, 2000; Oh, Woods, Lim, Theibert, & Couchman, 1998; Wilcox-Adelman, Denhez, & Goetinck, 2002). In this paper, we demonstrate that the extracellular and cytoplasmic domains of syndecans are intrinsically disordered regions, and we present a sequence analysis of the syndecans from 25 animal organisms.

Downloaded by [Tulane University] at 07:28 23 January 2015

4

E.I. Leonova and O.V. Galzitskaya

Figure 2. Schematic representation of the structure of syndecans 2 and 4. Notes: The external ectodomains are shed from the membrane by MMPS to form a “soluble” form. The variable (V) site in the cytoplasmic domain is unique for each syndecan. The C2 site of the cytoplasmic domain contains a PDZ-binding motif that interacts with intracellular adapter proteins. The C1 site interacts with intracellular actin-binding proteins. A GxxxG motif in the transmembrane domain is necessary for the dimerization of syndecans in the membrane.

2. Materials and methods 2.1. Generation of a protein database For each type of syndecan, we can collect a different number of syndecans with known sequences for various animals. As a result, there are 32 sequences for syndecan1 (Leonova & Galzitskaya, 2013a). We considered four types of syndecan sequences for 25 animal organisms. The sequences were obtained from the GenBank (http:// www.ncbi.nlm.nih.gov/nuccore/) and UniProt (ftp://ftp. ebi.ac.uk/pub/databases/SPproteomes/uniprot/proteomes/). See http://bioinfo.protres.ru/misc/syndecan_25.xlsx. 2.2. Search for disordered residues Disordered residues were predicted using the IsUnstruct program, which is based on the Ising model (Lobanov & Galzitskaya, 2011). The parameters of the program were determined and optimized on the basis of protein structure statistics. Testing demonstrated that the program yields reliable predictions. The program was previously developed and is available at http://bioinfo.protres.ru/ IsUnstruct (Lobanov, Sokolovskiy, & Galzitskaya, 2013). It should be noted that the Ising model has been used successfully to describe the two states related to the helix–coil transition for homopolypeptide chains. The

PONDR-FIT method was used to check the reliability of the predictions; a metaserver yields a consensus prediction for 10 programs (Xue, Dunbrack, Williams, Dunker, & Uversky, 2010). 3. Results and discussion 3.1. Syndecan structure and comparative characteristics of the amino acid sequences The diversity of the functions of syndecans is dependent on the syndecan structures. As mentioned above, there is a common (typical of all syndecans) pattern of attachment of glycosaminoglycans to a core protein. Both the heparan sulfate and chondroitin sulfate chains are linked covalently to serine residues. It is common to all syndecans that glycine links covalently next in the sequence to the serine residue. The binding is driven by a “special tetrasaccharide” that consists of xylose successively attached to two galactose residues and one glucuronic acid residue. This structural unit acts as an “original” primer that triggers the growth of the polysaccharide chain. Heparan sulfate is composed of two repeating disaccharide units N-acetylglucosamine–glucuronic acid and N-acetylglucosamine–iduronic acid, whereas the chondroitin sulfate consists of a repeating disaccharide

Cell communication

Comparison of the ectodomains for the syndecan 1–4 proteins of 25 animals. Identical regions are shadowed by gray color.

Downloaded by [Tulane University] at 07:28 23 January 2015

Table 1.

5

(Continued)

6

E.I. Leonova and O.V. Galzitskaya

Downloaded by [Tulane University] at 07:28 23 January 2015

Table 1. (Continued )

(Continued)

Cell communication

7

Downloaded by [Tulane University] at 07:28 23 January 2015

Table 1. (Continued )

unit that contains N-acetylgalactosamine–glucuronic acid (see Figure 1). Sequence analyses of all types of syndecans from 25 organisms have shown that there is a region rich in aspartic acid (D) that is located between two heparan sulfate-binding sites in the ectodomains and promotes the lack of structure (Table 1). In spite of a significant difference in the amino acid sequence of the ectodomains, these regions remain conserved and are typical of all types of syndecans. The absence of these regions in syndecan-3 of certain organisms, such as Pan troglodytes (common chimpanzee), Saimiri boliviensis (squirrel monkey), Ovis aries (sheep), Felis catus (domestic cat), and Mustela putorius furo (ferret), indicates alternative splicing (Table 1). It has been proposed that the genes encoding natively disordered proteins undergo an alternative splicing more frequently than the genes encoding ordered proteins (Romero et al., 2006). According to the Consensus CDS protein set CCDS database (http://www.ncbi.nlm.nih. gov/CCDS/CcdsBrowse.cgi), the genes that encode syndecans have five exons: the first exon encodes a signal peptide; the second exon encodes the attachment sites for heparan sulfate; the third and fourth exons encode the sites of chondroitin sulfate binding to syndecans 1 and 3 only; and the fifth exon encodes transmembrane and cytoplasmic domains. The second and fourth exons are assumed to undergo alternative splicing. As a result, during mRNA “maturation,” they may be spliced from pre-mRNA along with introns in various combinations and to create four forms of mRNA. Each type of syndecans may be found in four isoforms (Leonova & Galzitskaya, 2013b). However, there is no experimental confirmation of the existence of syndecan isoforms; in some sequences of syndecan-3, we have found the absence of a region referring to the second exon. Therefore, taking into account the fact that the second alternative exon encodes three sites for the attachment of heparan sulfate in syndecans 1, 2, and 4 and four sites in syndecan-3, it is possible to suggest the existence of isoforms of syndecans that lack heparan sulfate. In addition, the fourth

alternative exon (in syndecans 1 and 3) contains a site for the attachment of chondroitin sulfate chains (Figure 3), and the deletion of these sites can lead to a lack of chondroitin sulfate chains in particular isoforms. The presence of such isoforms may seriously affect their functions and lead to various diseases, including atherosclerosis, hypertriglyceridemia, or, as in anorexia, reduction in the neuromediator secretion by nervous cells. Additional investigation is required to confirm the existence of isoforms of syndecans. The cytoplasmic domain of syndecans has two conserved C1 and C2 sites that flank the variable region V. The conserved site C2 (EFYA) of the cytoplasmic domain binds to PDZ-binding proteins, such as synbindin, synectin, CASK/LIN-2, and syntenin, that play an important role in vesicular transportation, synaptic signaling, neuronal migration, and cancer metastasizing (Beauvais & Rapraeger, 2004). The conserved C1 site interacts with the actin-binding proteins ezrin, radixin, and moesin that regulate the organization of the actin cytoskeleton. It should be noted that the presence of Lys in region C1 can contribute to the lack of a three-dimensional structure. The exception is syndecan-4 in frogs (Table 2) because of the absence of lysine in region C1. The variable region V determines the specificity and uniqueness of each type of syndecan. The conserved sites C1 and C2 are typical not only of vertebrates but also of invertebrates (Brown, 2011). The transmembrane domain of all syndecans contains the conserved motif GxxxG (Table 2) that is necessary for the dimerization of syndecans and for the retention of cholesterol in the membrane (Barrett et al., 2012). 3.2. Comparison of the amino acid compositions of the syndecans of various animals The occurrences of 20 amino acid residues in 17 animal proteomes are summarized in Table 3 according to our recent estimates (Lobanov & Galzitskaya, 2012). The occurrences established for animal syndecan proteins are summarized in Table 3. In spite of the wide variation

Downloaded by [Tulane University] at 07:28 23 January 2015

8

E.I. Leonova and O.V. Galzitskaya

observed for the ectodomain, certain regularities persisted, i.e. Cys residues were almost absent, whereas Pro and Thr were abundant (except for in syndecan-4). Based on the occurrences of 20 amino acid residues in the syndecan proteins of various organisms, it can be observed that syndecan-1 contains higher levels of proline, glutamic acid, alanine, glycine, and threonine and lower levels of cysteine, phenylalanine, isoleucine, asparagine, tyrosine, arginine, and lysine than the average levels of these amino acids in the 17 proteomes. Syndecan-2 contains higher levels of alanine, threonine, serine, lysine, and asparaginic and glutamic acids, and lower levels of cysteine, asparagine, tryptophan, glutamine, arginine, and histidine. Syndecan-3 typically has higher levels of alanine, valine, threonine, proline, and glutamic acid and lower levels of cysteine, methionine, phenylalanine, isoleucine, tryptophan, tyrosine, glutamine, asparagine, histidine, and arginine. Syndecan-4 contains higher levels of isoleucine, leucine, valine, glycine, proline, and aspartic and glutamic acids, and lower levels of cysteine, tryptophan, threonine, asparagine, glutamine, histidine, and arginine. It should be noted that glycine, threonine, glutamine, glutamic acid, and proline

are members of residues that facilitate the disordered arrangement of proteins (Campen et al., 2008; Dunker et al., 2001; Garbuzynskiy, Lobanov, & Galzitskaya, 2004; Lobanov, Garbuzynskiy & Galzitskaya, 2010; Romero et al., 2001; Williams et al., 2001). In addition, the ectodomains of certain animals include disordered motifs that we previously observed in the protein structure database (Chen et al., 2004; Lee et al., 2011), such as polyhistidine and polythreonine repeats in frog syndecan-1; EPTNSS in Gallus gallus syndecan-1; APEDPE in Bos taurus syndecan-1; PSPPP in syndecan-3 of certain animals, such as Callithrix jacchus, Pongo abelii, Saimiri boliviensis; and others. The motif SGDLDD is present in most animal syndecan-4 structures and a polyD motif is present in frogs. Additionally, the poly-D motif was observed in all of the syndecan proteins from our database. This motif is located between two heparan sulfate-binding sites in ectodomains (Table 1). For the reliable prediction of the location of disordered regions, it is advantageous to use several programs employing different principles (Ferron, Longhi, Canard, & Karlin, 2006). Here, we utilize the program IsUnstruct (Lobanov & Galzitskaya, 2011) based on the Ising model

Figure 3. Schematic representation of alternative splicing variants of mouse syndecan-3 (Homo sapiens) as an example. (A) full amino acid sequence of a core protein (442 amino acid residues). Heparan sulfate is attached to serine residues at positions 80, 82, 84, and 91, and chondroitin sulfate is attached to serine residues at positions 314 and 367. (B) product of syndecan-3 mRNA without the 2nd exon. (C) product of syndecan-3 mRNA without the 4th exon. (D) product of syndecan-3 of mRNA without the 2nd and 4th exons.

Cell communication Comparison of the cytoplasmic domains for the syndecan 1–4 proteins of 25 animals.

Downloaded by [Tulane University] at 07:28 23 January 2015

Table 2.

(Continued)

9

10

E.I. Leonova and O.V. Galzitskaya

Downloaded by [Tulane University] at 07:28 23 January 2015

Table 2. (Continued ).

Note: Identical regions of C1 and C2 are shadowed by gray and black, respectively; mismatching amino acid residues of the variable region (V) are shadowed by dark gray.

Cell communication Table 3.

Comparative occurrences (%) of 20 amino acid residues in 17 animal proteomes and in each type of syndecan.

Occurrences (%) of 20 amino acid residues In 17 animal proteomes In syndecan-1 in Syndecan-2 In Syndecan-3 In syndecan-4

Downloaded by [Tulane University] at 07:28 23 January 2015

11

C

M

F

I

2.2 2.0 3.9 5.0 .8 .4 .2 .2

1.6 1.9 1.2 2.1

2.2 3.0 2.4 4.0

L

V

9.4 6

1.9 8.0 4.8 8.5 2.5 8.1 6.1 10.4

5.7 5.6 7.6 7.6

W

Y

1.1 3.0 .8 .5 .4 .1

A

G

6.9 6.1

T

S

Q

N

5.6 8.2 5.0 4.3

1.4 10.8 9.6 9.1 7.5 4.0 9.4 5.3 7.2 9.8 1.4 11.7 6.6 13.2 7.1 2.7 6.0 8.5 3.8 7.7

5.1 2.0 3.1 2.0

E

D

H

R

K

6.6 5.1 3.0 6.0 5.8

1.9 8.2 2.2 9.2 1.0 8.3 3.0 10.1

P 5.0

5.9 3.0 3.8 3.4 9.0 7.8 .9 4.6 7.2 5.8 4.4 .4 4.4 2.8 12.9 8.5 .9 3.5 5.6 7.2

Figure 4. IsUnstruct and PONDR-FIT analyses of human syndecans 1–4. Notes: Both the IsUnstruct (black circles) and PONDR-FIT (white circles) scores run from 0 to 1 with values below .5 generally indicating ordered/folded regions and above .5 indicating intrinsically unfolded/disordered regions.

and the metaserver PONDR-FIT (Xue et al., 2010), which combines the results of multiple disorder predictors for the analysis of the syndecan sequences. The profiles of the disordered regions in human syndecans are shown in Figure 3. The two prediction programs report very similar order/disorder trends over the entire sequence (see Figure 4 and the Supplementary material). The two programs indicate that the large variation for syndecan-2 is localized at positions 70–90, residues that are unfolded in the case of IsUnstruct and folded in the case of PONDR-FIT; and for syndecan-3, at positions 200–220, residues that are unfolded in the case of PONDR-FIT and folded in the case of IsUnstruct. In

general, (see http://bioinfo.protres.ru/misc/syndecan_25. xlsx) the syndecan 1–4 proteins are characterized by their disordered structure, with the exception of the transmembrane domain and the signal peptide. Because of their diverse functions and numerous interaction partners, syndecans can be expected to be natively disordered. Natively disordered proteins lack a unique tertiary structure in isolation and acquire it only when interacting with a partner (Dyson & Wright, 2005; Tompa, 2012; Uversky & Dunker, 2010). The conformation of such a protein in a complex is determined to a certain extent by the interaction partner, rather than depending solely on the amino acid sequence, as is

Downloaded by [Tulane University] at 07:28 23 January 2015

12

E.I. Leonova and O.V. Galzitskaya

characteristic of structured proteins. Natively disordered proteins are usually multifunctional, performing accessory functions in addition to the major function. In general, the size of the disordered region may vary from several amino acid residues to the total protein length. IDPs such as syndecans are resistant to proteases because they exist in a complex with their partner (or ligand). The cell surface may act as a partner as well (Iakoucheva, Brown, Lawson, Obradović, & Dunker, 2002). As soon as the partner is removed from the complex, the protein may degrade rapidly (Serdyuk, 2007; Wright & Dyson, 1999). This degradation is typical of a cell that must respond quickly and efficiently; an example is the regulation of the protein content during transcription and translation. In conclusion, syndecans as transmembrane proteoglycans are able to perform a wide range of functions that are determined by their structure, which is expected to be natively disordered. Natively disordered proteins lack a unique tertiary structure in isolation and acquire it only when interacting with a partner. The conformation of such a protein in a complex is determined to a certain extent by the interaction partner, rather than depending solely on the amino acid sequence, as is characteristic of structured proteins. It has been noted that the occurrences of 20 amino acids in syndecans 1–4 from 25 animals are different from those in 17 animal proteomes. The number of occurrences of Gly + Ala, Thr (except in syndecan-4), Glu, and Pro were higher in all syndecans. These amino acid residues are known to induce the lack of an ordered structure. In contrast, lower quantities were observed of amino acids that promote an ordered structure, such as Cys, Trp, Asn, and His (except in syndecan-1). In addition, all types of syndecans contain the poly-D motif in a region between two heparan sulfatebinding sites in the ectodomains and a region rich in Lys in the conserved C1 site of the cytoplasmic domain. The amino acid compositions of syndecans play an essential role in the formation of the disordered regions and govern their functions. Supplementary material The supplementary material for this paper is available online at http://dx.doi.10.1080/07391102.2014.926256. Acknowledgment We are grateful to D. Reifsnyder Hickey for assistance in preparation of the paper.

Funding The present work was performed under the financial support of the Russian Science Foundation [grant number No14-14-00536].

References Alberti, S., Halfmann, R., King, O., Kapila, A., & Lindquist, S. (2009). A systematic survey identifies prions and illuminates sequence features of prionogenic proteins. Cell, 137, 146–158. doi:10.1016/j.cell.2009.02.044 Alberts, B., Johnson, A., Lewis, J., Raff, M., Roberts, K., & Walter, P. (2008). MolecularBiology of the cell (5th ed.). New York, NY: Garland Science. Anfinsen, C. B. (1973). Principles that govern the folding of protein chains. Science, 181, 223–230. Asplund, A., Ostergren-Lunden, G., Camejo, G., Stillemark-Billton, P., & Bondjers, G. (2009). Hypoxia increases macrophage motility, possibly by decreasing the heparan sulfate proteoglycan biosynthesis. Journal of Leukocyte Biology, 86, 381–388. Baciu, P. C., Saoncella, S., Lee, S. H., Denhez, F., Leuthardt, D., & Goetinck, P. F. (2000). Syndesmos, a protein that interacts with the cytoplasmic domain of syndecan-4, mediates cell spreading and actin cytoskeletal organization. Journal of Cell Science, 113, 315–324. Barrett, P. J., Song, Y., Van Horn, W. D., Hustedt, E. J., Schafer, J. M., Hadziselimovic, A., … Sanders, C. R. (2012). The amyloid precursor protein has a flexible transmembrane domain and binds cholesterol. Science, 336, 1168–1171. doi:10.1126/science.1219988 Beauvais, D. M., & Rapraeger, A. C. (2004). Syndecans in tumor cell adhesion and signaling. Reproductive Biology and Endocrinology, 2, 3. Review doi:10.1186/1477-7827-2-3 Bernfield, M., Götte, M., Park, P. W., Reizes, O., Fitzgerald, M. L., Lincecum, J., & Zako, M. (1999). Functions of cell surface heparan sulfate proteoglycans. Annual Review of Biochemistry, 68, 729–777. Review. Brown, N. H. (2011). Extracellular matrix in development: Insights from mechanisms conserved between invertebrates and vertebrates. Cold Spring Harbor Perspectives in Biology, 3, a005082. doi:10.1101/cshperspect.a005082 Campen, A., Williams, R. M., Brown, C. J., Meng, J., Uversky, V. N., & Dunker, A. K. (2008). TOP-IDP-scale: A new amino acid scale measuring propensity for intrinsic disorder. Protein & Peptide Letters, 15, 956–963. PubMed PMID: 18991772. Caraux, A., Klein, B., Paiva, B., Bret, C., Schmitz, A., Fuhler, G. M., … Perez-Andres, M. (2010). Circulating human B and plasma cells. Age-associated changes in counts and detailed characterization of circulating normal CD138- and CD138+ plasma cells. Haematologica, 95, 1016–1020. doi:10.3324/haematol.2009.018689 Chakravarti, R., & Adams, J. C. (2006). Comparative genomics of the syndecans defines an ancestral genomic context associated with matrilins in vertebrates. BMC Genomics, 7, 83. Chen, E., Hermanson, S., & Ekker, S. C. (2004). Syndecan-2 is essential for angiogenic sprouting during zebrafish development. Blood, 103, 1710–1719. Chen, L., Couchman, J. R., Smith, J., & Woods, A. (2002). Molecular characterization of chicken syndecan-2 proteoglycan. Biochemical Journal, 366, 481–490. Choi, S., Kim, J. Y., Park, J. H., Lee, S. T., Han, I. O., & Oh, E. S. (2012). The matrix metalloproteinase-7 regulates the extracellular shedding of syndecan-2 from colon cancer cells. Biochemical and Biophysical Research Communications, 417, 1260–1264. doi:10.1016/j.bbrc.2011.12.120 David, G., Bai, X. M., Van der Schueren, B., Marynen, P., Cassiman, J. J., & Van den Berghe, H. (1993). Spatial and temporal changes in the expression of fibroglycan (syndecan-2) during mouse embryonic development. Development, 119, 841–854.

Downloaded by [Tulane University] at 07:28 23 January 2015

Cell communication Denhez, F., Wilcox-Adelman, S. A., Baciu, P. C., Saoncella, S., Lee, S., French, B., … Goetinck, P. F. (2002). Syndesmos, a syndecan-4 cytoplasmic domain interactor, binds to the focal adhesion adaptor proteins paxillin and Hic-5. Journal of Biological Chemistry, 277, 12270–12274. Dews, I. C., & MacKenzie, K. R. (2007). Transmembrane domains of the syndecan family of growth factor coreceptors display a hierarchy of homotypic and heterotypic interactions. Proceedings of the National academy of Sciences of the United States of America, 104, 20782–20787. Dunker, A. K., Lawson, J. D., Brown, C. J., Williams, R. M., Romero, P., Oh, J. S., … Obradovic, Z. (2001). Intrinsically disordered protein. Journal of Molecular Graphics and Modelling, 19, 26–59. Review. Dyson, H. J., & Wright, P. E. (2005). Intrinsically unstructured proteins and their functions. Nature Reviews Molecular Cell Biology, 6, 197–208. Review. Ferron, F., Longhi, S., Canard, B., & Karlin, D. (2006). A practical overview of protein disorder prediction methods. Proteins: Structure, Function, and Bioinformatics, 65, 1–14. Fitzgerald, M. L., Wang, Z., Park, P. W., Murphy, G., & Bernfield, M. (2000). Shedding of syndecan-1 and -4 ectodomains is regulated by multiple signaling pathways and mediated by a TIMP-3-sensitive metalloproteinase. The Journal of Cell Biology, 148, 811–824. Garbuzynskiy, S. O., Lobanov, M. Y., & Galzitskaya, O. V. (2004). To be folded or to be unfolded? Protein Science, 13, 2871–2877. PubMed PMID: 15498936. Gitler, A. D., & Shorter, J. (2011). RNA-binding proteins with prion-like domains in ALS and FTLD-U. Prion, 5, 179– 187. doi:10.4161/pri.5.3.17230. Review. Halfmann, R., Alberti, S., Krishnan, R., Lyle, N., O’Donnel, C. W., King, O.D., … Lindquist, S. (2011). Opposing effects of glutamine and asparagine govern prion formation by intrinsically disordered proteins. Molecular Cell, 43, 72–84. doi:10.1016/j.molcel.2011.05.013 Heldin, C. H. (1995). Dimerization of cell surface receptors in signal transduction. Cell, 80, 213–223. Review. Iakoucheva, L. M., Brown, C. J., Lawson, J. D., Obradović, Z., & Dunker, A. K. (2002). Intrinsic disorder in cell-signaling and cancer-associated proteins. Journal of Molecular Biology, 323, 573–584. Kaksonen, M., Pavlov, I., Voikar, V., Lauri, S. E., Hienola, A., Riekki, R., … Rauvala, H. (2002). Syndecan-3-deficient mice exhibit enhanced LTP and impaired hippocampus-dependent memory. Molecular and Cellular Neuroscience, 21, 158–172. Karlsson-Lindahl, L., Schmidt, L., Haage, D., Hansson, C., Taube, M., Egeciouglu, E., … Dickson, S. L. (2012). Heparanase affects food intake and regulates energy balance in mice. PLoS One, 7, e34313. doi:10.1371/journal.pone. 0034313 Kim, C. W., Goldberger, O. A., Gallo, R. L., & Bernfield, M. (1994). Members of the syndecan family of heparan sulfate proteoglycans are expressed in distinct cell-, tissue-, and development-specific patterns. Molecular Biology of the Cell, 5, 797–805. Kim, J. M., Lee, J. A., Cho, I. S., & Ihm, C. H. (2010). Soluble syndecan-1 at diagnosis and during follow up of multiple myeloma: A single institution study. The Korean Journal of Hematology, 45, 115–119. doi:10.5045/ kjh.2010.45.2.115 Kokenyesi, R., & Bernfield, M. (1994). Core protein structure and sequence determine the site and presence of heparan sulfate and chondroitin sulfate on syndecan-1. Journal of Biological Chemistry, 269, 12304–12309.

13

Lee, H., Kim, Y., Choi, Y., Choi, S., Hong, E., & Oh, E. S. (2011). Syndecan-2 cytoplasmic domain regulates colon cancer cell migration via interaction with syntenin-1. Biochemical and Biophysical Research Communications, 409, 148–153. doi:10.1016/j.bbrc.2011.04.135 Leonova, E. I., & Galzitskaya, O. V. (2013a). Comparative characteristics of the structure and function for syndecan-1 from animal organisms. Molecular Biology (Mosk), 47, 505–512. Leonova, E. I., & Galzitskaya, O. V. (2013b). Structure and functions of syndecans in vertebrates. Biochemistry (Moscow), 78, 1071–1085. doi:10.1134/S0006297913100015 Li, Q., Park, P. W., Wilson, C. L., & Parks, W. C. (2002). Matrilysin shedding of syndecan-1 regulates chemokine mobilization and transepithelial efflux of neutrophils in acute lung injury. Cell, 111, 635–646. PMID: 12464176 Li, Y. C., Park, M. J., Ye, S. K., Kim, C. W., & Kim, Y. N. (2006). Elevated levels of cholesterol-rich lipid rafts in cancer cells are correlated with apoptosis sensitivity induced by cholesterol-depleting agents. The American Journal of Pathology, 168, 1107–1118; quiz 1404-5. Lin, Y. L., Lei, Y. T., Hong, C. J., & Hsueh, Y. P. (2007). Syndecan-2 induces filopodia and dendritic spine formation via the neurofibromin-PKA-Ena/VASP pathway. The Journal of Cell Biology, 177, 829–841. Lobanov, M. Yu., & Galzitskaya, O. V. (2011). The Ising model for prediction of disordered residues from protein sequence alone. Physical Biology, 8, 035004. doi:10.1088/ 1478-3975/8/3/035004 Lobanov, M. Yu., & Galzitskaya, O. V. (2012). Occurrence of disordered patterns and homorepeats in eukaryotic and bacterial proteomes. Molecular BioSystems, 8, 327–337. doi:10.1039/c1mb05318c Lobanov, M. Yu., Garbuzynskiy, S. O., & Galzitskaya, O. V. (2010). Statistical analysis of unstructured amino acid residues in protein structures. Biochemistry (Moscow), 75, 192–200. Lobanov, M. Yu., Sokolovskiy, I. V., & Galzitskaya, O. V. (2013). IsUnstruct: Prediction of the residue status to be ordered or disordered in the protein chain by a method based on the Ising model. Journal of Biomolecular Structure and Dynamics, 31, 1034–1043. doi:10.1080/ 07391102.2012.718529 Manich, G., Mercader, C., del Valle, J., Duran-Vilaregut, J., Camins, A., Pallas, M., … Pelegri, C. (2011). Characterization of amyloid-β granules in the hippocampus of SAMP8 mice. Journal of Alzheimers Disease, 25, 535–546. doi:10.3233/JAD-2011-101713 Manon-Jensen, T., Itoh, Y., & Couchman, J. R. (2010). Proteoglycans in health and disease: The multiple roles of syndecan shedding. FEBS Journal, 277, 3876–3889. doi:10.1111/ j.1742-4658.2010.07798.x Oh, E. S., Woods, A., Lim, S. T., Theibert, A. W., & Couchman, J. R. (1998). Syndecan-4 proteoglycan cytoplasmic domain and phosphatidylinositol 4,5-bisphosphate coordinately regulate protein kinase C activity. Journal of Biological Chemistry, 273, 10624–10629. Pisconti, A., Cornelison, D. D., Olguin, H. C., Antwine, T. L., & Olwin, B. B. (2010). Syndecan-3 and Notch cooperate in regulating adult myogenesis. The Journal of Cell Biology, 190, 427–441. doi:10.1083/jcb.201003081 Romero, P., Obradovic, Z., Li, X., Garner, E. C., Brown, C. J., & Dunker, A. K. (2001). Sequence complexity of disordered protein. Proteins: Structure, Function, and Genetics, 42, 38–48. PubMed PMID: 11093259.

Downloaded by [Tulane University] at 07:28 23 January 2015

14

E.I. Leonova and O.V. Galzitskaya

Romero, P. R., Zaidi, S., Fang, Y. Y., Uversky, V. N., Radivojac, P., Oldfield, C. J., … Dunker, A. K. (2006). Alternative splicing in concert with protein intrinsic disorder enables increased functional diversity in multicellular organisms. Proceedings of the National Academy of Sciences of the United States of America, 103, 8390–8395. Saunders, S., Jalkanen, M., O’Farrell, S., & Bernfield, M. (1989). Molecular cloning of syndecan, an integral membrane proteoglycan. The Journal of Cell Biology, 108, 1547–1556. Savery, M. D., Jiang, J. X., Park, P. W., & Damiano, E. R. (2013). The endothelial glycocalyx in syndecan-1 deficient mice. Microvascular Research, 87, 83–91. doi:10.1016/ j.mvr.2013.02.001 Serdyuk, I. N. (2007). Structured proteins and proteins with intrinsic disorder. Molecular Biology, 41, 262–277. Simons, K., & Ikonen, E. (1997). Functional rafts in cell membranes. Nature, 387, 569–572. Review. Stanford, K. I., Bishop, J. R., Foley, E. M., Gonzales, J. C., Niesman, I. R., Witztum, J. L., & Esko, J. D. (2009). Syndecan-1 is the primary heparan sulfate proteoglycan mediating hepatic clearance of triglyceride-rich lipoproteins in mice. Journal of Clinical Investigation, 119, 3236–3245. doi:10.1172/JCI38251 Subramanian, S. V., Fitzgerald, M. L., & Bernfield, M. (1997). Regulated shedding of syndecan-1 and -4 ectodomains by thrombin and growth factor receptor activation. Journal of Biological Chemistry, 272, 14713–14720. Tompa, P. (2012). Intrinsically disordered proteins: A 10-year recap. Trends in Biochemical Sciences, 37, 509–516. doi:10.1016/j.tibs.2012.08.004. Review. Uversky, V. N., & Dunker, A. K. (2010). Understanding protein non-folding. Biochimica et Biophysica Acta (BBA) – Proteins and Proteomics, 1804, 1231–1264. doi:10.1016/ j.bbapap.2010.01.017. Review. Vanhoutte, D., Schellings, M. W., Gotte, M., Swinnen, M., Herias, V., Wild, M. K., … Heymans, S. (2007). Increased expression of syndecan-1 protects against cardiac dilatation

and dysfunction after myocardial infarction. Circulation, 115, 475–482. Wang, W., Haller, C. A., Wen, J., Wang, P., & Chaikof, E. L. (2008). Decoupled syndecan 1 mRNA and protein expression is differentially regulated by angiotensin II in macrophages. Journal of Cellular Physiology, 214, 750–756. Wang, Z., Götte, M., Bernfield, M., & Reizes, O. (2005). Constitutive and accelerated shedding of murine syndecan-1 is mediated by cleavage of its core protein at a specific juxtamembrane site. Biochemistry, 44, 12355–12361. Weinbaum, S., Tarbell, J. M., & Damiano, E. R. (2007). The structure and function of the endothelial glycocalyx layer. Annual Review of Biomedical Engineering, 9, 121–167. Review. Wilcox-Adelman, S. A., Denhez, F., & Goetinck, P. F. (2002). Syndecan-4 modulates focal adhesion kinase phosphorylation. Journal of Biological Chemistry, 277, 32970–32977. Williams, R. M., Obradovi, Z., Mathura, V., Braun, W., Garner, E. C., Young, J., … Dunker, A. K. (2001). The protein non-folding problem: Amino acid determinants of intrinsic order and disorder. Pacific Symposium on Biocomputing, 89–100. PubMed PMID: 11262981. Wright, P. E., & Dyson, H. J. (1999). Intrinsically unstructured proteins: Re-assessing the protein structure-function paradigm. Journal of Molecular Biology, 293, 321–331. Xue, B., Dunbrack, R. L., Williams, R. W., Dunker, A. K., & Uversky, V. N. (2010). PONDR-FIT: A meta-predictor of intrinsically disordered amino acids. Biochimica et Biophysica Acta (BBA) – Proteins and Proteomics, 1804, 996– 1010. doi:10.1016/j.bbapap.2010.01.011 Zhao, N., Liu, H., & Lilly, B. (2012). Reciprocal regulation of syndecan-2 and notch signaling in vascular smooth muscle cells. Journal of Biological Chemistry, 287, 16111–16120. doi:10.1074/jbc.M111.322107 Zheng, Q., Zhu, J., Shanabrough, M., Borok, E., Benoit, S. C., Horvath, T. L., … Reizes, O. (2010). Enhanced anorexigenic signaling in lean obesity resistant syndecan-3 null mice. Neuroscience, 171, 1032–1040. doi:10.1016/ j.neuroscience.2010.09.060

Cell communication using intrinsically disordered proteins: what can syndecans say?

Because intrinsically disordered proteins are incapable of forming unique tertiary structures in isolation, their interaction with partner structures ...
6MB Sizes 0 Downloads 3 Views