28.02.2014 Views

Membranes and mammalian glycolipid transferring ... - Åbo Akademi

Membranes and mammalian glycolipid transferring ... - Åbo Akademi

Membranes and mammalian glycolipid transferring ... - Åbo Akademi

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

Accepted Manuscript<br />

Title: <strong>Membranes</strong> <strong>and</strong> <strong>mammalian</strong> <strong>glycolipid</strong> <strong>transferring</strong><br />

proteins<br />

Author: Jessica Tuuf Peter Mattjus<br />

PII:<br />

S0009-3084(13)00151-5<br />

DOI:<br />

http://dx.doi.org/doi:10.1016/j.chemphyslip.2013.10.013<br />

Reference: CPL 4241<br />

To appear in:<br />

Chemistry <strong>and</strong> Physics of Lipids<br />

Received date: 28-8-2013<br />

Revised date: 29-10-2013<br />

Accepted date: 30-10-2013<br />

Please cite this article as: Tuuf, J., Mattjus, P.,<strong>Membranes</strong> <strong>and</strong> <strong>mammalian</strong><br />

<strong>glycolipid</strong> <strong>transferring</strong> proteins, Chemistry <strong>and</strong> Physics of Lipids (2013),<br />

http://dx.doi.org/10.1016/j.chemphyslip.2013.10.013<br />

This is a PDF file of an unedited manuscript that has been accepted for publication.<br />

As a service to our customers we are providing this early version of the manuscript.<br />

The manuscript will undergo copyediting, typesetting, <strong>and</strong> review of the resulting proof<br />

before it is published in its final form. Please note that during the production process<br />

errors may be discovered which could affect the content, <strong>and</strong> all legal disclaimers that<br />

apply to the journal pertain.


<strong>Membranes</strong> <strong>and</strong> <strong>mammalian</strong> <strong>glycolipid</strong> <strong>transferring</strong> proteins<br />

Jessica Tuuf <strong>and</strong> Peter Mattjus<br />

Biochemistry, Department of Biosciences, <strong>Åbo</strong> <strong>Akademi</strong> University, Turku, Finl<strong>and</strong>.<br />

Keywords: GLTP, FAPP2, glycosphingolipid, membrane binding<br />

Abbreviations:<br />

ARF, ADP-ribosylation factor; DPPC, 1,2-dipalmitoyl-sn-glycero-3-phosphocholine; ER,<br />

endoplasmic reticulum; FAPP1/2, phosphoinositol 4-phosphate adaptor protein-1/2; FFAT,<br />

two phenylalanines (FF) in an acidic tract; GalCer, galactosylceramide; GCS,<br />

glucosylceramide synthase; GlcCer, glucosylceramide; GLTP, <strong>glycolipid</strong> transfer protein;<br />

GSL, glycosphingolipid; LacCer, lactosylceramide; OSBP, oxysterol-binding protein; PC,<br />

Accepted Manuscript<br />

phosphatidylcholine; PE, phosphatidylethanolamine; PH, pleckstrin homology;<br />

phosphatidylinositol; PI3P, phosphatidylinositol-3-phosphate; PI4P, phosphatidylinositol-4-<br />

phosphate; PM, plasma membrane; POPC, 1-palmitoyl-2-oleyl-sn-glycero-3-phosphocholine;<br />

PS, phosphatidylserine; SM, sphingomyelin; START, steroidogenic acute regulatory protein<br />

(StAR)-related lipid transfer; TGN, trans-Golgi-network; VAP, vesicle-associated membrane<br />

protein-associated protein<br />

PI,<br />

1<br />

Page 1 of 34


Abstract<br />

Glycolipids are synthesized in <strong>and</strong> on various organelles throughout the cell. Their<br />

trafficking inside the cell is complex <strong>and</strong> involves both vesicular <strong>and</strong> protein-mediated<br />

machineries. Most important for the bulk lipid transport is the vesicular system, however,<br />

lipids moved by transfer proteins is also becoming more characterized. Here we review<br />

the latest advances in the <strong>glycolipid</strong> transfer protein (GLTP) <strong>and</strong> the phosphoinositol 4-<br />

phosphate adaptor protein-2 (FAPP2) field, from a membrane point of view.<br />

Accepted Manuscript<br />

Page 2 of 34


1. Introduction<br />

The hydrophobic nature of the lipids requires different transport <strong>and</strong> trafficking mechanisms<br />

compared to water soluble biomolecules. Lipids are synthesized on <strong>and</strong> in different<br />

organelles, <strong>and</strong> cells constantly need to adjust <strong>and</strong> respond to changes in the lipid<br />

requirements. To do this efficiently they need different transport machineries, connected to<br />

the synthesis <strong>and</strong> degradation pathways. Several mechanisms in the cells are used for lipid<br />

distribution. Most important for the bulk lipid transport is the vesicular system, however, lipid<br />

movement mediated by transfer proteins is also widely characterized. In addition, lipid<br />

transfer proteins have also often been given the role as sensors, responsible for coordinating<br />

the transport routes within the synthesis <strong>and</strong> degradation machineries. There are several<br />

<strong>glycolipid</strong>-binding proteins in mammals, proteins that specifically recognize <strong>glycolipid</strong>s. The<br />

sphingolipid activator proteins (saposins) bind to different glycosphingolipids (GSL) <strong>and</strong> help<br />

in the degradation process of GSLs, with short oligosaccharide chains, in the lysosomes<br />

(Schulze et al., 2009). Mutations in the saposins can cause severe lysosomal storage disorders,<br />

like Gaucher disease (Schnabel et al., 1991). There is also evidence that non-specific lipid<br />

transfer proteins, isolated from beef liver, can recognize GSLs <strong>and</strong> transfer them between<br />

membranes (Bloj <strong>and</strong> Zilversmit, 1981). Another emerging group are the <strong>glycolipid</strong> transfer<br />

proteins. In this review we will focus mostly on <strong>glycolipid</strong> transfer protein (GLTP), but also<br />

on phosphoinositol 4-phosphate adaptor protein-2 (FAPP2), two proteins that belong to the<br />

GLTP superfamily.<br />

2. Glycolipid transfer protein (GLTP)<br />

GLTP is a small soluble, 24 kDa, protein that has been identified in many organisms. GLTP<br />

Accepted Manuscript<br />

was first discovered in the cytosolic fraction from bovine spleen (Metz <strong>and</strong> Radin, 1980;Metz<br />

<strong>and</strong> Radin, 1982) <strong>and</strong> has since then been found in e.g. liver <strong>and</strong> brain from various<br />

<strong>mammalian</strong> sources (Abe et al., 1982;Yamada <strong>and</strong> Sasaki, 1982a;Yamada <strong>and</strong> Sasaki,<br />

1982b;Wong et al., 1984). In these studies GLTP was always purified from the cytosolic<br />

fraction <strong>and</strong>, consequently, the protein was postulated to be cytosolic. It has later been<br />

confirmed that GLTP indeed remains in the cytoplasm <strong>and</strong> not within other major cellular<br />

organelles (Tuuf <strong>and</strong> Mattjus, 2007). Homologs to <strong>mammalian</strong> GLTPs have furthermore been<br />

discovered in both plants (West et al., 2008) <strong>and</strong> yeast (Saupe et al., 1994). The expression<br />

2<br />

Page 3 of 34


levels of GLTP in bovine tissue have been analyzed (Lin et al., 2000). The results show that<br />

the cerebrum <strong>and</strong> kidney display the highest GLTP mRNA levels, which coincides with<br />

normally quite high amount of glycosphingolipids. GLTP can accelerate the transfer of<br />

<strong>glycolipid</strong>s, both diacylglycerol- <strong>and</strong> sphingoid-based, between two lipid membranes.<br />

Substrates that can be used by GLTPs include glucosylceramide (GlcCer), lactosylceramide<br />

(LacCer), galactosylceramide (GalCer), sulfatide, GM 1 , <strong>and</strong> GM 3 (Yamada et al., 1985;Brown<br />

et al., 1985). Furthermore, the initial sugar residue, linked to the ceramide or diacylglycerol<br />

backbone, needs to be in β-configuration for GLTP to recognize it as a substrate (Yamada et<br />

al., 1986). On the other h<strong>and</strong>, GLTP cannot transfer phosphatidylcholine (PC),<br />

phosphatidylethanolamine (PE), sphingomyelin (SM), phosphatidylinositol (PI), cholesterol<br />

<strong>and</strong> cholesterol oleate (Yamada et al., 1985). The activity of GLTP is sensitive to changes in<br />

the surrounding pH (West et al., 2006). Results show that GLTP has the highest transfer<br />

activity at pH 7.0, 66% transfer activity at pH 10 <strong>and</strong> only 11% activity at pH 4.0. This could<br />

give an important clue on GLTP subcellular localization <strong>and</strong> cellular activity, since GLTP<br />

will encounter different pH conditions within the cell.<br />

The amino acid sequence of GLTPs from different mammals is highly conserved. The protein<br />

sequence of GLTP from pig brain was the first to be determined by Edman degradation (Abe,<br />

1990). Later, it was demonstrated that GLTP contains 209 amino acids <strong>and</strong> that the sequence<br />

between porcine <strong>and</strong> bovine is 100% identical (Lin et al., 2000). Human GLTP has almost as<br />

high identity <strong>and</strong> shows 98% identity compared to porcine <strong>and</strong> bovine amino acid sequences<br />

(Li et al., 2004). Furthermore, the human GLTP amino acid sequence is 100% identical to the<br />

amino acid sequences of chimpanzee <strong>and</strong> macaque, 99% to the nonprimate mammal Canis<br />

familaris <strong>and</strong> 98% identical to the Bos taurus sequence (Zou et al., 2008). Two single-copy<br />

GLTP genes have been found in human cells, on chromosome 11 <strong>and</strong> 12. The<br />

transcriptionally active GLTP is on chromosome 12 <strong>and</strong> contains five exons <strong>and</strong> four introns,<br />

Accepted Manuscript<br />

while the inactive pseudogene is located on chromosome 11 (Zou et al., 2008). The<br />

expression of GLTP mRNA is regulated by two transcription factors, Sp1 <strong>and</strong> Sp3. These<br />

factors bind to multiple sites in the GC-boxes localized to the promoter of GLTP (Zou et al.,<br />

2011). In the same study, the ability of different sphingolipids to enhance the GLTP<br />

transcription of GLTP was analyzed. Interestingly, only ceramide (but not for example<br />

GlcCer, GM 1 or sulfatide) could enhance the transcription levels of GLTP via the Sp1 <strong>and</strong><br />

Sp3 transcription factors. The structure of both bovine <strong>and</strong> human GLTP has been<br />

determined, both in its apo-form <strong>and</strong> in complex with different <strong>glycolipid</strong>s (Malinina et al.,<br />

3<br />

Page 4 of 34


2004;West et al., 2004;Airenne et al., 2006;Malinina et al., 2006). The GLTP structure<br />

contains eight α-helices, that are arranged in a two-layer fold, a structure that is completely<br />

novel for lipid <strong>transferring</strong> proteins. The carbohydrate group will bind to a recognition center<br />

on the surface of GLTP, through hydrophobic contacts <strong>and</strong> hydrogen bonding, <strong>and</strong> the<br />

hydrocarbon acyl chains will be accommodated into a hydrophobic pocket within the protein<br />

(Malinina et al., 2004). The binding of <strong>glycolipid</strong>s to GLTP will be discussed extensively in<br />

chapter 4.<br />

The formation of GLTP dimers has been discussed in early publications. There are three<br />

cysteine residues in the GLTP sequence, two of them reside inside the protein <strong>and</strong> the third is<br />

on the surface of GLTP. Abe <strong>and</strong> coworker suggested that the two internal cysteines might<br />

form a intramolecular disulfide bond, while the cysteine on the outside could form a disulfide<br />

bond with another GLTP molecule (Abe <strong>and</strong> Sasaki, 1989a;Abe <strong>and</strong> Sasaki, 1989b). The<br />

disulfide bridge forming ability of GLTP was suggested to function as a means for regulating<br />

GLTP transfer activity, since it was demonstrated that the dimer had almost no transfer<br />

activity in contrast to GLTP with an intermolecular disulfide bridge, which had a high transfer<br />

activity. However, results from crystallization studies show no intra- or intermolecular<br />

disulfide bonds in the GLTP structure (Malinina et al., 2004;Airenne et al., 2006), but based<br />

on structural requirements it was suggested that it might be possible that an internal disulfide<br />

bridge would help in regulating GLTP activity (Airenne et al., 2006). It is still unknown<br />

whether disulfide bond-formation actually occurs in vivo as a regulatory mechanism for<br />

GLTP function. Recently a new transfer protein was discovered that moves ceramide-1-<br />

phosphate, <strong>and</strong> was named CPTP, ceramide-1-phosphate transfer protein (Simanshu et al.,<br />

2013). Structurally, CPTP has a very similar fold compared to the GLTP fold, however, CPTP<br />

associates with the trans-Golgi network (TGN), nucleus <strong>and</strong> the plasma membrane (PM). A<br />

CPTP-mediated ceramide-1-phosphate decrease in plasma membranes <strong>and</strong> increase in the<br />

Accepted Manuscript<br />

Golgi complex stimulates cPLA 2 α release of arachidonic acid, triggering pro-inflammatory<br />

eicosanoid generation.<br />

3. Phosphoinositol 4–phosphate adaptor protein-2 (FAPP2)<br />

There are two closely related homologs of four-phosphate adaptor proteins, type 1 <strong>and</strong> 2<br />

(FAPP1 <strong>and</strong> FAPP2). FAPP1 was originally discovered in a study, when trying to identify<br />

4<br />

Page 5 of 34


novel proteins that could interact with various phosphatidylinositol-phosphates (Dowler et al.,<br />

2000). The FAPP1 gene is located on chromosome 2 <strong>and</strong> encodes a protein with 300 amino<br />

acids. FAPP1 has in its amino-terminal a pleckstrin homology (PH) domain <strong>and</strong> in the C-<br />

terminus resides a proline rich-motif (Godi et al., 2004). FAPP2, on the other h<strong>and</strong>, is a<br />

protein with 519 amino acids <strong>and</strong> is encoded by a single-copy gene on chromosome 7. In<br />

addition to a PH domain in the N-terminus, FAPP2 has a GLTP domain in the carboxyterminal,<br />

which makes it a member of the GLTP superfamily (Godi et al., 2004;Kamlekar et<br />

al., 2013) (Fig. 1A). The PH domain of FAPP1 <strong>and</strong> FAPP2 is 80% identical (Godi et al.,<br />

2004). FAPP2 is able to transfer GlcCer, both fluorescently labeled <strong>and</strong> natural, but not SM,<br />

PC or ceramide (D'Angelo et al., 2007). Recently it was demonstrated that in addition to<br />

GlcCer, the GLTP domain of FAPP2 can specifically transfer GalCer <strong>and</strong> LacCer between<br />

membranes, but not the ganglioside GM 1 or the negatively charged sulfatide (Kamlekar et al.,<br />

2013). This substrate specificity makes the GLTP domain of FAPP2 more similar to HET-C2,<br />

a fungal homolog of GLTP, than to human GLTP itself.<br />

The FAPPs appear to be ubiquitously expressed <strong>and</strong> have a cytoplasmic distribution, though,<br />

with a strong preference towards the TGN via their targeting signals (Godi et al., 2004). For<br />

example, in mouse tissue FAPP2 is highly expressed in kidney, but is also found in testis, the<br />

small intestine, spleen <strong>and</strong> cerebrum (D'Angelo et al., 2013). Homologs of FAPP2 have been<br />

found in vertebrates <strong>and</strong> echinoderms (D'Angelo et al., 2008) <strong>and</strong> sequence analysis of<br />

various FAPP2 homologs reveals that both the PH domain as well as the GLTP domain are<br />

highly conserved (D'Angelo et al., 2012).<br />

By analyzing FAPP2 by small-angle x-ray scattering <strong>and</strong> analytical ultracentrifugation, it<br />

appears that FAPP2 is dimeric in solution (Cao et al., 2009). Whether this is the case in<br />

biological systems remains to be demonstrated. To date, it has been proven to be difficult to<br />

Accepted Manuscript<br />

crystallize FAPP2, probably because of its large size <strong>and</strong> flexibility. However, a lowresolution<br />

structure of dimeric FAPP2 by small-angle x-ray scattering has been obtained,<br />

showing a molecular shape that is well extended <strong>and</strong> curved, with a length of 30 nm (Cao et<br />

al., 2009). The GLTP domain has been analyzed using homology modeling. These models<br />

show a conserved folding structure against human GLTP, with the key residues, important for<br />

the <strong>glycolipid</strong> <strong>transferring</strong> properties retained (D'Angelo et al., 2007). A truncated FAPP2<br />

form, with a 212 amino acids C-terminus (the GLTP domain), was used to study the<br />

resemblance to GLTP (Kamlekar et al., 2013). This protein showed a structure with high<br />

5<br />

Page 6 of 34


content of alpha-helices <strong>and</strong> results from experiments using fluorescence emission of the<br />

intrinsic tryptophan residues suggested that the GLTP domain actually adopts a GLTP fold.<br />

However, there are some differences in FAPP2 compared to GLTP, e.g. FAPP2 has a more<br />

selective substrate specificity which is due to a restricted <strong>glycolipid</strong> head group recognition<br />

center (Kamlekar et al., 2013). The FAPP1 PH domain has been analyzed using nuclear<br />

magnetic resonance-based solution structures in a free state, together with micelles or<br />

phosphatidylinositol-4-phosphate (PI4P) (Lenoir et al., 2010). Shortly after, the crystal<br />

structure of FAPP1 PH domain was determined (He et al., 2011). The PH domain was<br />

reported to consist of a β-barrel capped by an α-helix at one edge. Three additional loops<br />

connect the str<strong>and</strong>s; a long β1- β2 loop, a short β6- β7 loop <strong>and</strong> a β3-β4 loop that is partially<br />

hidden in the structure. Between the β4 <strong>and</strong> β5 str<strong>and</strong> one can also find a one-turn α-helix.<br />

Hydrogen-deuterium exchange mass spectrometry has been used on FAPP2 to look at<br />

conformational changes upon GlcCer binding (D'Angelo et al., 2013). The GLTP <strong>and</strong> the PH<br />

domain appear to be separated by a very flexible linker region <strong>and</strong> results show that when<br />

FAPP2 binds GlcCer a conformational change will occur in the whole structure, not only in<br />

the <strong>glycolipid</strong>-binding domain, but also the linker region <strong>and</strong> the PH domain will be affected.<br />

4. Glycolipid binding<br />

The broad specificity of GLTP for various glycosphingolipids, for example <strong>glycolipid</strong>s with<br />

large variations in their head group structure, is due to the hydrogen bonding adaptable head<br />

group recognition site localized at the entrance to the hydrophobic tunnel of the protein. This<br />

enables GLTP to selectively bind the GSL head <strong>and</strong> the proximal ceramide amide group with<br />

multiple hydrogen bonds.<br />

Accepted Manuscript<br />

GLTP from different species all have the characteristic feature of an adaptive hydrophobic<br />

cavity. The physical function of GLTP is to protect the lipid substrate hydrocarbon chains<br />

from the aqueous environment once the <strong>glycolipid</strong> is extracted from the membrane. Due to the<br />

large variation in the chain lengths <strong>and</strong> saturation of natural GSLs, an adaptable cavity seems<br />

to be the evolutionary structural optimized form. This is in contrast to the precursor synthases<br />

responsible for ceramide syntheses, that seem to have developed into six different ceramide<br />

synthases for various chain lengths of ceramides (Tidhar <strong>and</strong> Futerman, 2013). Part of the<br />

hydrophobic tunnel of GLTP is partly collapsed in the apo-state. The 3D structure of GLTP,<br />

6<br />

Page 7 of 34


as determined by crystallization (West et al., 2004), reveals a s<strong>and</strong>wiched all-alpha-helical<br />

conformation (Malinina et al., 2004;Airenne et al., 2006) (Fig. 2A). Several amino acid<br />

residues have been identified based on the crystal structure studies, that are required for<br />

<strong>glycolipid</strong> binding (Malinina et al., 2004;Malakhova et al., 2005;Airenne et al., 2006;Malinina<br />

et al., 2006). Tryptophan at position 96 (helix 4) is in a key role in the forming of a platform<br />

of hydrogen bonds between the hydroxyl group of the first glucose or galactose. For FAPP2<br />

the corresponding tryptophan required for <strong>glycolipid</strong> transfer activity is at position 407.<br />

Further hydrogen bonds are formed between His-140 <strong>and</strong> the ceramide amide region of the<br />

<strong>glycolipid</strong> backbone (Fig. 2B). By comparison of the apo-GLTP <strong>and</strong> GLTP with a lipid<br />

bound, it is evident that there is a shifting of the alpha-2 <strong>and</strong> alpha-6 helices <strong>and</strong> by changing<br />

the side chain conformations of residues Phe-42, Ile-45, Phe-148 <strong>and</strong> Leu-152 when a lipid is<br />

bound. Presumable, when the <strong>glycolipid</strong> acyl chains come in contact with these hydrophobic<br />

amino acid side chains inside the hydrophobic tunnel, they change their conformation, <strong>and</strong><br />

consequently, a shift in helix-2 <strong>and</strong> helix-6 occurs. We believe that this acts as the process by<br />

which GLTP desorbs from the membrane into the surrounding aqueous solution. Neither<br />

GLTP nor FAPP2 contain the “KRTIQK“ amino acid motif described by Fantini <strong>and</strong><br />

coworkers as a <strong>glycolipid</strong>-binding algorithm (Fantini et al., 2006). They identified this domain<br />

in several amino acid sequences from Helicobacter pylori bacterial adhesion protein adhesin<br />

A (HpaA). HpsA is known to interact with membranes containing <strong>glycolipid</strong>s. HpaA, to our<br />

knowledge, appears not to be able to transfer <strong>glycolipid</strong>s.<br />

4.1 Sphingosine in versus sphingosine out<br />

The cavity that accommodates the hydrophobic parts of the GSL is adaptable <strong>and</strong> aligned with<br />

nonpolar side chains effectively excluding water (Malinina et al., 2004;Airenne et al., 2006).<br />

Thorough examination of different crystal structures of GLTP with different bound GSLs<br />

Accepted Manuscript<br />

with varying acyl chain composition reveals that either one or both acyl chains are<br />

accommodated inside the hydrophobic tunnel. In the sphingosine-in form, both chains are<br />

inside, whereas in the sphingosine-out conformation, the amide-linked chain occupies more of<br />

the hydrophobic cavity preventing the sphingosine chain to enter (Fig. 3A). Most of the<br />

crystallized forms of GLTP are in the sphingosine-out conformation. However, it is still not<br />

clear if GLTP has its substrate bound in the sphingosine-out mode in biological systems. It is<br />

neither known how different diacylglycerol-based <strong>glycolipid</strong>s are bound by GLTP, if they<br />

also adopt both a sphingosine-in <strong>and</strong> sphingosine-out binding mode. Encapsulation of both<br />

7<br />

Page 8 of 34


chains has only been observed for wildtype GLTP monomers <strong>and</strong> with either GlcCer or<br />

LacCer containing an amide-linked oleoyl (18:1 Δ 9 ) acyl chain (Malinina et al.,<br />

2004;Samygina et al., 2011). A proposed mechanism for the accommodation of the<br />

sphingosine chain inside GLTP involves His-140 hydrogen bonding to the ceramide amide<br />

group <strong>and</strong> movement of two phenylalanines (Phe-42 & Phe-148), creating additional space for<br />

chain encapsulation. A mutation of the residue His-140 leads to a complete inactivation of<br />

GLTP (Malinina et al., 2004). Another mutant GLTP (D48V) crystallizes with the 24:1 Δ 9<br />

sulfatide in a sphingosine-in form. The mutant has a structural resemblance very close to that<br />

of the wildtype GLTP with sphingosine-in with bound 18:1 GlcCer. The D48V mutation as<br />

well as another double mutant A47D/D48D are interesting proteins in that they do not bind<br />

neutral GSL, but can be engineered to become transfer proteins for 24:1 Δ 9 sulfatide<br />

(Samygina et al., 2011).<br />

What determines the sphingosine-in conformation, <strong>and</strong> is this simply a crystallization<br />

phenomenon? There are very small conformational differences between the sphingosine-in<br />

<strong>and</strong> sphingosine-out GLTP-lipid complexes. At a first glance, it appears that increasing the N-<br />

linked acyl chain length, <strong>and</strong> due to the adaptable hydrophobic cavity of GLTP, sphingosine<br />

appears to be simply forced out by a long chain. However, co-crystallization with short chains<br />

N-linked chains, such as C8 & C10, did not allow for a sphingosine-in alignment. Most likely<br />

the shorter acyl chain never reaches all the way to the bottom of the hydrophobic tunnel <strong>and</strong>,<br />

consequently, the tunnel remains narrow preventing the sphingosine from entering. Another<br />

peculiarity that has been complicating the GLTP crystal structure studies has been the<br />

presence of short extraneous hydrocarbon chains not originating from the bound GSL, found<br />

inside the hydrophobic cavity (Malinina et al., 2004;West et al., 2004;Airenne et al., 2006).<br />

These most likely originate from the bacterial expression of the protein used for<br />

crystallization. Finding hydrocarbons <strong>and</strong> detergents inside the crystal structure of<br />

Accepted Manuscript<br />

hydrophobic proteins is seen frequently, <strong>and</strong> is sometimes crucial for obtaining protein<br />

crystals (Haapalainen et al., 2001;Guan et al., 2001;Wright et al., 2003). The extraneous<br />

hydrocarbon seen in the apo-GLTP crystal structures is displaced from the hydrophobic<br />

tunnel when GLTP forms a complex with the <strong>glycolipid</strong> (Malinina et al., 2004;Airenne et al.,<br />

2006). It could be possible that in vivo, GLTP could also carry a hydrophobic molecule when<br />

no <strong>glycolipid</strong> is bound. It is not clear if the molecule enters from the <strong>glycolipid</strong> binding site or<br />

enters into the hydrophobic cavity from the opposite end, through the openings at the end of<br />

the tunnel, formed by helix-1, helix-7 <strong>and</strong> helix-8. The opening is structurally wide enough to<br />

8<br />

Page 9 of 34


let a hydrocarbon chain pass. The far end opening of the hydrophobic cavity can be seen in<br />

Fig. 3A, black arrow.<br />

4.2 Crystal-Related Dimerization in GSL-GLTP Complexes<br />

In a recent paper, the dimerization states of GLTP were analyzed (Samygina et al., 2013).<br />

Previously, homodimers of GLTP has been found in various crystal structures when GLTP<br />

has been complexed with different glycosphingolipids. It appears that when GLTP binds the<br />

<strong>glycolipid</strong> in the sphingosine–out mode, a dimer will form (Malinina et al., 2006). This event<br />

is suggested to be due to interaction of two different GLTP molecules as well as the adjacent<br />

sphingosine hydrocarbon chains (Samygina et al., 2013). The authors proposed that the<br />

function of GLTP might be regulated by forming reversible homodimers. This is further<br />

supported by the finding that when wildtype GLTP <strong>and</strong> a mutant GLTP (D48V) bind 24:1<br />

sulfatide in the absence of lipid membranes, a homodimer of GLTP will form in solution<br />

(Samygina et al., 2011). In contrast to these results it was shown, based on analytical<br />

ultracentrifugation experiments, size-exclusion chromatography <strong>and</strong> affinity-tag<br />

immunoadsorbtion experiments, that both apo-GLTP <strong>and</strong> GLTP complexed with GalCer<br />

(24:1) are monomeric in solution (Malinina et al., 2006). However, the authors still proposed,<br />

based on optimal docking area (ODA) energy contact calculations, that upon binding of<br />

GalCer in the sphingosine-out mode, the sphingosine chain would interact with a sphingosine<br />

molecule on a partner GLTP molecule, thereby forming crystal-related cross dimers.<br />

The crystal-related dimerization of GLTP allows the two sphingosine chains to interact with<br />

each other presumably in a more energetically favorable conformation than being inside the<br />

tunnel. Is the sphingosine-out form, <strong>and</strong> the subsequent dimerization of GLTP the biologically<br />

active carrier of GSLs in vivo? More than a dozen crystal structures of GLTP have the bound<br />

Accepted Manuscript<br />

<strong>glycolipid</strong> in the sphingosine-out conformation. Could this simple be a form generated during<br />

high protein concentrations <strong>and</strong> under limited hydration? The sphingosine chain might be<br />

jutted outwards into a low hydration environment, hydrophobic amino acid residues in close<br />

proximity to the tunnel entrance (Fig. 3B, red patch indicated with a black arrow) <strong>and</strong><br />

extraneous hydrocarbons occupying the far end of the tunnel might drive a sphingosine-out<br />

conformation. It is of vital importance to keep in mind that all dimers form when no<br />

membrane surfaces are present. Analysis of GLTP in solution over a wide range of<br />

concentrations using light scattering <strong>and</strong> gel-filtration all indicate that in the presence of<br />

9<br />

Page 10 of 34


membranes GLTP is a monomer (Zhai et al., 2009).<br />

4.3 GLTP-like proteins<br />

The Arabidopsis thaliana GLTP-like protein (Brodersen et al., 2002), ACD11 has also<br />

conserved residues, the GLTP amino acids Asp-48 <strong>and</strong> His-140 participate in the hydrogen<br />

binding to the sphingosine base <strong>and</strong> are conserved in ACD11 as Asp-60 <strong>and</strong> His-143<br />

respectively (Petersen et al., 2008). ACD11 appears to be able to adopt a GLTP-like structure<br />

(Petersen et al., 2008). ACD11 appears not to be capable of <strong>transferring</strong> <strong>and</strong> (binding?)<br />

glycosphingolipids in vitro, only the single chain sphingosine (Brodersen et al., 2002).<br />

Replacing the sphingosine hydrogen bond forming amino acids in ACD11 (Asp-60 & His-<br />

143) causes a complete loss of transfer activity. However, if tryptophan is introduced in<br />

position 103, to make ACD11 resemble GLTP, it does not improve its <strong>glycolipid</strong> transfer<br />

activity.<br />

5. Membrane composition consequences on protein activity<br />

The composition of the membrane has a great impact on protein function <strong>and</strong> activity.<br />

Processes like binding to membranes <strong>and</strong> <strong>glycolipid</strong> extraction by the GLTPs have been<br />

studied quite intensively. We will here give the latest insights in such events.<br />

5.1 Membrane binding<br />

Accepted Manuscript<br />

The rate-limiting step in the <strong>glycolipid</strong> transfer process is thought to be the GLTP-<strong>glycolipid</strong><br />

complex formation, or the dissociation of the protein-lipid complex from the membrane into<br />

the surrounding aqueous environment (Rao et al., 2005). The initial GLTP binding to the<br />

membrane, however, is not the limiting process <strong>and</strong> appears to have a minimal contribution to<br />

the overall changes to the enthalpy <strong>and</strong> entropy for the overall transfer process. Once docked<br />

to the membrane surface, GLTP is searching for the <strong>glycolipid</strong> substrate in the membrane.<br />

This process is probably a combination of both a lateral flow of both GLTP <strong>and</strong> lipids. GLTP<br />

transfer activity is different depending on the membrane composition <strong>and</strong> the lateral<br />

10<br />

Page 11 of 34


miscibility of the lipids in the membrane. If the <strong>glycolipid</strong> is in a membrane in a fluid phase<br />

the <strong>glycolipid</strong> removal is favored. GLTP binds more strongly to sphingomyelin containing<br />

membranes than to phosphatidylcholine membranes. This is true regardless of acyl chain<br />

composition, i.e. saturated or unsaturated (Ohvo-Rekilä <strong>and</strong> Mattjus, 2011).<br />

5.2 Lipid extraction<br />

Membrane composition. It has been shown that the rate of transfer by GLTP is sensitive to the<br />

membrane composition surrounding the <strong>glycolipid</strong>. The positively charged areas of GLTP<br />

affect how well GSLs are moved from donor to acceptor membranes. GalCer in negatively<br />

charged vesicles (5 or 10 mol% negative lipids) is transferred significantly slower than from<br />

the donor vesicles composed of neutral membranes (Mattjus et al., 2000). The same amount<br />

of negative charge in acceptor vesicles does not impede the transfer rate as effectively.<br />

Positively charged lipids in the donor vesicle membrane neither affect the GSL transfer<br />

(Mattjus et al., 2000). GLTP is positively charged at neutral pH (pI=9.0) <strong>and</strong> will be attracted<br />

to the negatively charged donor membrane through electrostatic interactions between the<br />

protein <strong>and</strong> the membrane surface, resulting in a slow transfer. It is speculated that the GLTP<br />

off-rate from the donor surface becomes slow <strong>and</strong> consequently the transfer rate is slow.<br />

GLTP cannot transfer GalCer <strong>and</strong> GlcCer from vesicles made of saturated SMs, like<br />

palmitoyl-SM, regardless of curvature of the donor vesicle membranes (Mattjus et al.,<br />

2002;Nylund <strong>and</strong> Mattjus, 2005). Addition of cholesterol has only minimal effect on the<br />

transfer rate. If donor vesicles are made of different SM analogues, that more resemble a PC,<br />

transfer is possible. If the 3-hydroxy group of SM is replaced by a hydrogen, <strong>and</strong> the trans-4,5<br />

double bond of sphingosine is reduced, transfer will occur (Nylund et al., 2006). GLTP is also<br />

able to transfer <strong>glycolipid</strong>s from membranes composed of unsaturated SMs, however much<br />

Accepted Manuscript<br />

slower than from a chain-matched PC (Mattjus et al., 2002;Nylund <strong>and</strong> Mattjus, 2005).<br />

Membrane curvature. The transfer rates for GLTP are more than 5 times greater for small<br />

highly curved fluid 1-palmitoyl-2-oleyl-sn-glycero-3-phosphocholine (POPC) membranes<br />

compared to that for large membranes, that are comparable to a planar bilayer (Rao et al.,<br />

2005;Nylund et al., 2007). The transfer rate of fluorescently labeled GalCer from fluid POPC<br />

vesicles decreases in a linear fashion as the vesicles become larger, to an almost undetectable<br />

rate for vesicles above 800 nm in diameter. From planar monolayers no <strong>glycolipid</strong> transfer<br />

11<br />

Page 12 of 34


occurs (Nylund et al., 2007). If the membrane is composed of 1,2-dipalmitoyl-sn-glycero-3-<br />

phosphocholine (DPPC) there is no difference in the anthrylvinyl-GalCer transfer rates<br />

regardless of the donor vesicle size except for the smallest vesicles (50 nm in diameter) that<br />

are significantly higher. It is noteworthy that the transfer is very small from gel-like<br />

membranes in general. It is speculated that GLTP can bind clustered <strong>glycolipid</strong>s in a fluid<br />

environment such as POPC, but with limited accessibility to <strong>glycolipid</strong>s well dispersed into a<br />

more tightly packed environment such as that in DPPC <strong>and</strong> palmitoyl-SM (Nylund <strong>and</strong><br />

Mattjus, 2005;Nylund et al., 2006;Brown <strong>and</strong> Mattjus, 2007;Nylund et al., 2007;Mattjus,<br />

2009;Ohvo-Rekilä <strong>and</strong> Mattjus, 2011). Recently Brown <strong>and</strong> colleagues found that GLTP was<br />

able to extract BODIPY-GlcCer from a planar monolayer composed of POPC <strong>and</strong> low<br />

amounts (physiological) of either PE or phospatidic acid (PA) (15 <strong>and</strong> 5 mol% respectively)<br />

at bilayer comparable surface pressures (30-35 mN/m) (Zhai et al., 2013). It is unclear what is<br />

happening with phosphatidylserine (POPS) in these experiments, it appears not to stimulate<br />

GlcCer transfer to the same extent as PE <strong>and</strong> PA (Zhai et al., 2013), but rather penetrate the<br />

monolayer. Previously Nylund <strong>and</strong> co-workers showed that the exclusion pressure for GLTP<br />

i.e. the pressure where GLTP no longer can penetrate the monolayer, was clearly the highest<br />

with DPPS (30.7 mN/m) of all lipids analyzed, even negatively charged sulfatide (Nylund et<br />

al., 2007). GLTP is positively charged at neutral pH <strong>and</strong> penetrates negatively charged<br />

monolayer lipid films at higher surface pressures than those of neutral lipid films. However,<br />

DPPS is also more easily penetrated than the <strong>glycolipid</strong> sulfatide. The relatively larger<br />

penetration of GLTP into the lipid monolayers at low surface pressures does not necessarily<br />

indicate a direct interaction of the protein with the lipid molecules, as GLTP alone is surfaceactive<br />

<strong>and</strong> can spontaneously migrate from the subphase to a lipid free interface. This shows<br />

that the extraction of GLTP can be enhanced by alterations in the matrix lipid, independent of<br />

any additional proteins. The hydration around the GSL is likely to be the driving force for the<br />

effects seen by PE <strong>and</strong> PA. Saturated-chain PEs (12:0, 14:0 & 16:0) are not miscible with<br />

Accepted Manuscript<br />

chain-mixed GlcCer either above or below T m of the phospholipid. However, the unsaturated<br />

POPE, OPPE <strong>and</strong> SOPE are all miscible with GlcCer at physiological temperatures (Quinn,<br />

2011;Quinn, 2012). It is likely that GlcCer <strong>and</strong> GalCer interact more favorable with POPE<br />

than with POPC in monolayers because of the smaller <strong>and</strong> less hydration of the PE head. The<br />

hydration of GlcCer <strong>and</strong> GalCer more resembles PE than PC. In the monolayer system one<br />

leaflet is missing, this inevitably allows the lipids to move more freely out from the<br />

membrane plane, less hydrophobic forces keep the lipid stronger into the membrane.<br />

12<br />

Page 13 of 34


5.3 GLTP membrane binding, desorption mechanism <strong>and</strong> lipid release<br />

Trp-142 has been confirmed to be essential in the GLTP membrane binding process. In<br />

FAPP2 <strong>and</strong> HET-C2 the conserved amino acids are Trp-447 <strong>and</strong> Phe-149 respectively.<br />

Several other hydrophobic amino acids take part in the docking of GLTP onto the membrane<br />

surface (Malinina et al., 2004;Rao et al., 2005;Airenne et al., 2006;West et al., 2006;Neumann<br />

et al., 2008;Zhai et al., 2009;Kamlekar et al., 2010;Kenoth et al., 2010). A coloring of the<br />

amino acids according to their hydrophobic scale is presented in Fig. 3B. The most<br />

hydrophobic amino acids are shown in red, <strong>and</strong> amino acids with lower hydrophobicity in<br />

lighter red shades, white being fully hydrophilic in nature (Eisenberg et al., 1984). Examining<br />

the changes to the crystal structure of GLTP after the <strong>glycolipid</strong> is bound to the hydrophobic<br />

tunnel reveals several movements in two interhelical loops <strong>and</strong> two different helices. Alphahelix<br />

2 moves along its axis <strong>and</strong> outwards <strong>and</strong> alpha-helix 6 moves outwards. The net result of<br />

these movements is an expansion of a hydrophobic tunnel to accommodate the lipid, either<br />

with one or two chains inside. A putative membrane binding domain with high<br />

hydrophobicity is indicated with a black arrow in Fig. 3B. There are only small changes in<br />

this domain between the apo-GLTP <strong>and</strong> the lig<strong>and</strong>-bound form. The domain is formed by a<br />

group of tryptophan, tyrosine, isoleucine <strong>and</strong> nonpolar residues on the protein surface, in<br />

close proximity to the <strong>glycolipid</strong>-binding site. The small change in the hydrophobic domain<br />

suggests that after GSL binding also other parts of the protein are responsible for the release<br />

of GLTP from the membrane.<br />

The mechanism how the GSL is lifted from the membrane into the GLTP cavity is unclear.<br />

Perhaps the ability of the sphingosine chain to bind to the outer grow of GLTP will allow the<br />

amide-linked chain to enter the hydrophobic tunnel, specifically recognized <strong>and</strong> bound by the<br />

amino acids of the sugar binding pocket of GLTP. We speculate that the mechanism could be<br />

Accepted Manuscript<br />

that the membrane-binding domain of GLTP generates a disturbance in the hydrogen-bond<br />

network between the water molecules <strong>and</strong> the membrane lipid head groups, allowing an acyl<br />

chain from the lipids in the membrane to move upward towards the hydrophobic amino acids,<br />

out from the hydrophobic interior, onto the GLTP surface. If the acyl chain belongs to a lipid<br />

that is also recognized by the sugar-binding pocket of GLTP, a lipid binding takes place. If<br />

the chain belongs to a lipid that is not recognized, it simply moves back into the membrane<br />

core, <strong>and</strong> the next lipid is allowed to be tested by GLTP. Once a <strong>glycolipid</strong> binding has<br />

occurred <strong>and</strong> the <strong>glycolipid</strong> has been captured by GLTP, the second chain would then be<br />

13<br />

Page 14 of 34


pulled along <strong>and</strong> would enter the hydrophobic tunnel. Once GLTP lifts of the interface, as a<br />

consequence of the helical movements <strong>and</strong> structural changes in the membrane-binding<br />

domain, the sphingosine chain would be forced away form the aqueous environment into a<br />

more energetically favorable milieu inside GLTP. Very little is known about the release of the<br />

bound <strong>glycolipid</strong> from GLTP, once it encounters new membranes. For instance, the binding<br />

of FAPP2 carrying glucosylceramide is stronger than an empty apo-FAPP2 (D'Angelo et al.,<br />

2013), if this is the case also for GLTP is at present unknown.<br />

6. Subcellular Targeting of GLTP <strong>and</strong> FAPP2<br />

There are numerous sequences, motifs <strong>and</strong> domains within proteins that help them find their<br />

right destinations in the cell. Both protein-protein interactions as well as protein-lipid<br />

interactions are important in guiding cellular components to various subcellular<br />

compartments. The C1 <strong>and</strong>/or C2 domains of for example protein kinase C, diacylglycerol<br />

kinases <strong>and</strong> some Ca 2+ -dependent phospholipases, bind most commonly to phospholipids in<br />

cell membranes (Cho, 2001). Another domain, the steroidogenic acute regulatory protein<br />

(StAR)-related lipid transfer (START) domain, is important in lipid trafficking <strong>and</strong><br />

metabolism <strong>and</strong> the different START domains identify several lipids, including cholesterol<br />

<strong>and</strong> ceramide (Alpy <strong>and</strong> Tomasetto, 2005). The Zn 2+ -binding FYVE domains (about 70<br />

residues) are, on the other h<strong>and</strong>, very specific for phosphatidylinositol-3-phosphates (PI3P)<br />

(Gaullier et al., 1998;Patki et al., 1998) <strong>and</strong> are responsible for endosomal targeting<br />

(Stenmark et al., 1996). Another group of phosphoinositides-binding domains are the PX<br />

domains, found in a diverse set of proteins (Simonsen <strong>and</strong> Stenmark, 2001). Many of the PX<br />

domain-containing proteins also preferentially recognize PI3P <strong>and</strong> localize to the endosomes<br />

(Tessier <strong>and</strong> Woodgett, 2006). The PH domains are very divergent in their sequences, but<br />

Accepted Manuscript<br />

mediate several protein-protein interactions as well as recognizing a variety of phosphorylated<br />

derivatives of phosphatidylinositol (Scheffzek <strong>and</strong> Welti, 2012). The specific binding of PH<br />

domains to distinct phosphoinositides, located at different sites in the cell, are especially<br />

important amongst lipid-binding proteins, since this recruitment decide the protein subcellular<br />

localization <strong>and</strong> may regulate protein activity (Levine <strong>and</strong> Munro, 2002). Although there are a<br />

large variety of domains regulating protein trafficking <strong>and</strong> activity, more targeting signals will<br />

for sure be found in the future. So this raise the important question, how can the cytoplasmic<br />

proteins GLTP <strong>and</strong> FAPP2 find their specific sites of action?<br />

14<br />

Page 15 of 34


6.1 GLTP<br />

GLTP is a cytoplasmic protein <strong>and</strong> results obtained from different experimental methods<br />

suggest that GLTP does not reside within the endoplasmic reticulum (ER), the nucleus, the<br />

mitochondria or the Golgi (Tuuf <strong>and</strong> Mattjus, 2007). However, analysis of the amino acid<br />

sequence of GLTP from different organisms, indicate that there still might exist motifs or<br />

sequences within GLTP that would enable targeting of the protein to different compartments<br />

in the cell. In 2009 a motif in GLTP was characterized, a motif the authors referred to as the<br />

FFAT-like motif (Tuuf et al., 2009). Previously it has been shown that several lipid-binding<br />

proteins possess a conserved motif designated FFAT (two phenylalanines (FF) in an acidic<br />

tract) (Loewen et al., 2003), a targeting signal which directs them to the cytosolic surface of<br />

the ER for association with the vesicle-associated membrane protein-associated proteins<br />

(VAMP-associated proteins or VAPs) (Wyles et al., 2002;Loewen et al., 2003;Kawano et al.,<br />

2006). The VAPs have been imlicated in several important processes in the cell, like vesicle<br />

transport (Skehel et al., 1995), the unfolded protein response (Kanekura et al., 2006) <strong>and</strong> in<br />

the neurodegenerative disorder Amyothrophic lateral sclerosis, type 8 (Nishimura et al.,<br />

2004). Two of the isoforms of the VAPs, VAP-A <strong>and</strong> VAP-B, are mainly located in the ER<br />

(Weir et al., 2001;Nishimura et al., 2004), <strong>and</strong> it is proposed that these integral membrane<br />

proteins localize to specific ER contact sites, which could act as meeting points for several<br />

lipid transfer proteins (Levine <strong>and</strong> Loewen, 2006). The consensus sequence of the FFAT<br />

motif, almost exclusively found in lipid-binding proteins, is reported to contain the residues<br />

EFFDAxE appearing in an acidic environment (Loewen et al., 2003). In the same study it was<br />

demonstrated that minor substitutions in some of the amino acids are allowed, while other<br />

residues are essential for an interaction to occur with the VAPs. This conclusion is in<br />

agreement with more recent structural <strong>and</strong> experimental studies, where it is demonstrated that<br />

Accepted Manuscript<br />

many amino acid substitutions in the FFAT motif can indeed be tolerated (Kaiser et al.,<br />

2005;Mikitova <strong>and</strong> Levine, 2012). The FFAT-like motif identified in GLTP contains the<br />

amino acids 32 PFFDCLG 38 , a region which is also responsible for VAP-A-binding in vitro<br />

(Tuuf et al., 2009) (Fig. 1B). This FFAT-like motif is closely related to EFFDAxE <strong>and</strong><br />

contains the two phenylalanines (F) <strong>and</strong> the aspartate (D) residue, together with proline,<br />

cysteine, leucine <strong>and</strong> glycine. The fact that the FFAT-like motif is exposed on the surface of<br />

the protein (Airenne et al., 2006) also indicates that it is structurally available for a interaction<br />

with the VAPs (Fig. 2A). However, in a recent analysis on proteins with FFAT-like motifs the<br />

15<br />

Page 16 of 34


authors proposed that the FFAT-like motif in GLTP might not be optimal situated for<br />

interaction with VAP (Mikitova <strong>and</strong> Levine, 2012). This is due to 32 PFFDCLG 38 in GLTP is<br />

residing in a α-helix <strong>and</strong> not in an β-str<strong>and</strong>-like conformation, a structure that has previously<br />

been found in FFAT-containing proteins in VAP-FFAT-binding studies (Kaiser et al.,<br />

2005;Furuita et al., 2010). However, <strong>and</strong> interestingly, another FFAT-like-motif in the GLTP<br />

amino acid sequence has been identified (Mikitova <strong>and</strong> Levine, 2012). This sequence is weak,<br />

contains the residues 22 PFLEAVS 28 <strong>and</strong> lies upstream the region 32 PFFDCLG 38 previously<br />

found in GLTP (Tuuf et al., 2009) (Fig. 1B). It was suggested that both the FFAT-like motifs<br />

might act in concert, <strong>and</strong> consequently, a stronger VAP interaction would be obtained. Other<br />

proteins, with FFAT-like motifs even further from the consensus sequence, have been<br />

reported to interact with the VAPs under physiological conditions (Saita et al.,<br />

2009;Saravanan et al., 2009), which could imply that GLTP also might associate with the<br />

VAPs in vivo <strong>and</strong> not just in vitro. However, whether GLTP actually interacts with the VAPs<br />

in biological systems still remains unclear.<br />

Although GLTP is suggested to have a cytosolic distribution (Tuuf <strong>and</strong> Mattjus, 2007), it<br />

should not be excluded that GLTP might interact with various organelles during different<br />

stages in the cell cycle, upon certain cellular signals or that the association of GLTP with<br />

organelle membranes might be weak. In a large phosphoproteomic study on phosphorylation<br />

sites using mass spectrometric methods, there was a site identified in the GLTP sequence that<br />

was phosphorylated, threonine 69 (Olsen et al., 2010). This might raise the possibility that the<br />

subcellular localization of GLTP <strong>and</strong>/or interactions with other proteins might be regulated by<br />

phosphorylation/dephosphorylation events. The phosphorylation status of ceramide transfer<br />

protein <strong>and</strong> oxysterol-binding protein (OSBP), which also interact with the VAPs, can affect<br />

the targeting of the proteins to different organelles (Kumagai et al., 2007;Fugmann et al.,<br />

Accepted Manuscript<br />

2007;Saito et al., 2008;Tomishige et al., 2009;Nhek et al., 2010). Other possible posttranslational<br />

modifications of GLTP that could act as regulatory mechanisms could involve<br />

ubiquitylation or perhaps attachment of various lipids. Interestingly, a recent study shows that<br />

GLTP can be ubiquitylated (Wagner et al., 2011).<br />

6.2 FAPP2<br />

The FAPPs possess a PH domain, which is important for targeting of the protein. This domain<br />

specifically recognizes PI4P <strong>and</strong> the GTPase ADP-ribosylation factor (ARF) on the TGN<br />

16<br />

Page 17 of 34


(Godi et al., 2004) (Fig. 1A). The same mechanism is utilized by OSBP in yeast, that uses its<br />

PH domain to associate with both ARF1 <strong>and</strong> PI4P in the Golgi complex (Levine <strong>and</strong> Munro,<br />

2002) . In vitro studies on the GlcCer transport ability of FAPP2 showed that if ARF1 <strong>and</strong><br />

PI4P were added to acceptor membranes, the transport of GlcCer mediated by FAPP2<br />

between lipid donors to acceptors was enhanced (D'Angelo et al., 2007). It is the PI4P that<br />

dictates the targeting <strong>and</strong> upon GlcCer binding, FAPP2 translocates to the PI4P binding site<br />

on TGN (D'Angelo et al., 2013). However, the phospatidylinositol-4-OH kinase-β is recruited<br />

to the Golgi by the ARFs, which consequently links PI4P production <strong>and</strong> ARF localization<br />

together (Godi et al., 1999). The binding of the FAPP1 PH domain to ARF1 <strong>and</strong> PI4P has<br />

been analyzed (He et al., 2011). It appears that PI4P is absolutely necessary for the PH<br />

domain to be inserted into the membrane, <strong>and</strong> that the PI4P binding is pH-dependent, showing<br />

a stronger binding at pH 6.5 than 7.4. Furthermore, in the same study it was demonstrated that<br />

although the binding sites of ARF1 <strong>and</strong> PI4P are in close proximity, they can bind<br />

independently of each other <strong>and</strong> on the same time. These features of the PH domain<br />

contribute to the specific targeting of the FAPPs to the Golgi complex. FAPP2 with a bound<br />

GlcCer is targeted to the TGN via its PH domain <strong>and</strong> is capable to bind to PI4P <strong>and</strong> small<br />

GTPase ARF1 proteins (Godi et al., 2004). The mechanism how the bound GlcCer would<br />

increase the affinity of FAPP2 towards PI4P in the TGN remains unclear. Presumable, when<br />

GlcCer is bound to the GLTP domain of FAPP2, the protein adopts another conformation,<br />

different than the apo-form, allowing for an association by the PH domain to the membranes<br />

in the TGN. Once the GlcCer is released, the apo-FAPP2 leaves the membrane, <strong>and</strong> the<br />

affinity to bind PI4P by the PH domain again becomes lower. How the release of GlcCer<br />

takes place is not known (D'Angelo et al., 2013). GlcCer needs to be located on the inner<br />

leaflet of the TGN membranes in order to become further glycosylated to LacCer <strong>and</strong> further<br />

to Gb 3 by the enzymes residing in the lumen of the Golgi. It is tempting to speculate that<br />

when GlcCer exits the hydrophobic cavity of FAPP2, it would directly be pushed, with the<br />

Accepted Manuscript<br />

glucose head group first, directly to the inner leaflet. This mechanism would render a GlcCer<br />

flipping from the outer to the inner leaflet of the TGN unnecessary.<br />

FAPP2 in mammals also contain a weak FFAT-like motif, defined as TFFSTMN (Mikitova<br />

<strong>and</strong> Levine, 2012) (Fig. 1B). However, results from experiments from the same study showed<br />

that FAPP2 did not target the VAPs on ER as wildtype, but if the serine at position 4 was<br />

pseudophosphorylated, it showed a weak targeting to the VAPs. To our knowledge, it has not<br />

yet been demonstrated that FAPP2 actually interacts with the VAPs in vivo. In addition, there<br />

17<br />

Page 18 of 34


are several phosphorylation sites in FAPP2 that has been identified, however, whether these<br />

modifications are of importance for the function or subcellular localization of FAPP2 remains<br />

to be determined.<br />

7. Biological functions of GLTP <strong>and</strong> FAPP2<br />

The substrate specificity of both GLTP <strong>and</strong> FAPP2 suggests that the members of the GLTP<br />

superfamily are vital for regulating the GSL metabolism. Amongst the lipid transfer proteins,<br />

exist both genuine lipid transporters (Hanada et al., 2003) <strong>and</strong>/or sensors of various lipids<br />

(Litvak et al., 2005;Vihervaara et al., 2011). What about the proteins that are specific for<br />

glycosphingolipids? The biological function of GLTP is still not clear. However, implications<br />

have been made that GLTP might catalyze the transfer of GlcCer from the Golgi to the PM<br />

(Warnock et al., 1994;Halter et al., 2007). Other reports propose that GLTP might in addition<br />

function as an intracellular sensor of GSLs. Malakhova <strong>and</strong> coworkers mutated the lig<strong>and</strong>ing<br />

site in GLTP <strong>and</strong> looked at the ability of GLTP to acquire <strong>and</strong> release its substrate<br />

(Malakhova et al., 2005). The authors speculated that in addition to acting as a <strong>glycolipid</strong><br />

carrier, GLTP might function as a GlcCer sensor or a presenter of <strong>glycolipid</strong>s to protein<br />

receptors, since the results showed that wildtype GLTP was not keen to release the lig<strong>and</strong> to<br />

POPC membranes. Furthermore, GLTP appears to sense changes in the amount of newly<br />

synthesized GlcCer (Kjellberg <strong>and</strong> Mattjus, 2013). When GLTP is overexpressed in HeLa<br />

cells, de novo synthesis of GlcCer, but not LacCer or GalCer, is enhanced (Tuuf <strong>and</strong> Mattjus,<br />

2007). Downregulation of GLTP resulted in no changes in de novo syntheses of the<br />

sphingolipids, which made the authors suggest that GLTP is not necessarily involved in<br />

GlcCer transport, bur rather has a sensor function (Tuuf <strong>and</strong> Mattjus, 2007). Together, these<br />

studies indicate that GlcCer is the relevant lipid substrate for GLTP. This goes well with<br />

Accepted Manuscript<br />

GlcCer being the only <strong>glycolipid</strong> synthesized at the cytoplasmic side of organelle membranes<br />

(Futerman <strong>and</strong> Pagano, 1991;Jeckel et al., 1992). Another proposed function of GLTP is as a<br />

regulator of adjusting levels of cellular ceramide <strong>and</strong>, consequently, also the production of<br />

GSLs (Zou et al., 2011).<br />

FAPP2 has been implicated in several biological processes in the cell. Initially, FAPP2 was<br />

suggested to be involved in the transport of cargo between the TGN <strong>and</strong> the PM (Godi et al.,<br />

2004;Vieira et al., 2005). FAPP2 has also a tubulation activity <strong>and</strong> is suggested to be<br />

18<br />

Page 19 of 34


esponsible for tubulating the membranes in the TGN through its PH domain, which gives<br />

further evidence for a role of FAPP2 in Golgi-PM trafficking (Cao et al., 2009). The PH<br />

domain of FAPP1 is furthermore able to tubulate membranes (Lenoir et al., 2010). Another<br />

major role of FAPP2 was demonstrated by D´Angelo <strong>and</strong> coworkers, when they discovered<br />

that FAPP2 was an important factor in the glycosphingolipid synthetic pathways (D'Angelo et<br />

al., 2007). FAPP2 was responsible for transporting GlcCer from the cytosolic side of cis-<br />

Golgi to the later Golgi compartments, where the more complex GSLs are synthesized<br />

(Lannert et al., 1998). This transport was shown to be non-vesicular <strong>and</strong> require both the<br />

presence of ARF <strong>and</strong> PI4P on the later Golgi (D'Angelo et al., 2007). In a recent work it has<br />

been shown that GlcCer can undergo two fates. After GlcCer is synthesized in can either flip<br />

into the lumen of Golgi <strong>and</strong> be a precursor molecule for LacCer <strong>and</strong> GM 3 or alternatively, it<br />

can be transported by FAPP2 through the cytosol to TGN for globoside production, Gb 3<br />

(D'Angelo et al., 2013). This interesting <strong>and</strong> novel finding shows how the sorting <strong>and</strong><br />

branching of GlcCer by two different transport mechanisms, the non-vesicular (FAPP2) <strong>and</strong><br />

the vesicular, will decide which GSL end product is produced. However, FAPP2 has also<br />

been shown to mediate a GlcCer-transport in another direction, from the cis-Golgi back to the<br />

ER (Halter et al., 2007). Here it is suggested that after GlcCer reaches the ER, it would flipflop<br />

into the lumen of ER <strong>and</strong> be transported through Golgi by vesicle transport to the TGN<br />

where it could be processed to more complex GSLs. The flip-flop of GlcCer into the lumen of<br />

ER might be catalyzed by a ATP-independent phospholipid flippase (Chalat et al., 2012). A<br />

possible VAP-interaction, as previously discussed in chapter 6.2, might enable ER targeting<br />

of FAPP2 (Mikitova <strong>and</strong> Levine, 2012).<br />

8. Conclusion<br />

What is known about the depth of GLTP penetration into the membrane surface (Fig. 4)?<br />

Accepted Manuscript<br />

Experimental evidence indicated that the GLTP binding to the membrane is nonperturbing<br />

<strong>and</strong> not deep (Rao et al., 2005;Nylund et al., 2007). A theoretical orientation of proteins in<br />

membranes (OPM) analyses indicate a membrane penetration depth for GLTP of 3.4 Å with a<br />

tilt angle of 86° for the apo-GLTP <strong>and</strong> 3.9 Å <strong>and</strong> 87° for GLTP with a LacCer bound. It<br />

would be interesting to relate this predicted penetration depth with the membrane curvature,<br />

head group hydration <strong>and</strong> the depth of the GSLs in the surrounding lipid microenvironment,<br />

to mechanistically underst<strong>and</strong> how the GLTP lipid extraction from the membrane takes place.<br />

19<br />

Page 20 of 34


The <strong>glycolipid</strong> depth is dependent on the acyl chain length <strong>and</strong> saturation of the ceramide<br />

backbone, <strong>and</strong> consequently, the lipid species is directly affecting the exposure of the<br />

saccharide moieties into the plane of the bilayer. This aglycone (non-sugar part of the<br />

<strong>glycolipid</strong>) modulation of <strong>glycolipid</strong>s will strongly affect how they are recognized for<br />

instance by their transfer proteins (Lingwood, 1996;Lingwood, 2011). It is clear that all<br />

membrane GSLs are not distributed equally over the cellular membranes <strong>and</strong> their amount<br />

only controlled by their respective synthesis <strong>and</strong> breakdown machineries. It is also<br />

fundamental to know whether the GSLs are transported by vesicular means or by proteins.<br />

The function of the <strong>glycolipid</strong> <strong>transferring</strong> proteins is slowly emerging while new members<br />

are found, like the CPTP protein. This further adds complexity to the intracellular lipid<br />

trafficking picture.<br />

Acknowledgements<br />

We thank Anders Backman for help with the images.<br />

References<br />

References<br />

Abe, A. 1990. Primary structure of <strong>glycolipid</strong> transfer protein from pig brain. J. Biol. Chem.<br />

265, 9634-9637.<br />

Abe, A. <strong>and</strong> T.Sasaki. 1989a. Formation of an intramolecular disulfide bond of <strong>glycolipid</strong><br />

transfer protein. Biochim. Biophys. Acta 985, 45-50.<br />

Abe, A. <strong>and</strong> T.Sasaki. 1989b. Sulfhydryl groups in <strong>glycolipid</strong> transfer protein: formation of an<br />

intramolecular disulfide bond <strong>and</strong> oligomers by Cu2+-catalyzed oxidation. Biochim.<br />

Biophys. Acta 985, 38-44.<br />

Abe, A., K.Yamada, <strong>and</strong> T.Sasaki. 1982. A protein purified from pig brain accelerates the<br />

inter-membranous translocation of mono- <strong>and</strong> dihexosylceramides, but not the<br />

translocation of phospholipids. Biochem. Biophys. Res. Commun. 104, 1386-1393.<br />

Airenne, T.T., H.Kidron, Y.Nymalm, M.Nylund, G.West, P.Mattjus, <strong>and</strong> T.A.Salminen.<br />

2006. Structural evidence for adaptive lig<strong>and</strong> binding of <strong>glycolipid</strong> transfer protein. J.<br />

Mol. Biol. 355, 224-236.<br />

Alpy, F. <strong>and</strong> C.Tomasetto. 2005. Give lipids a START: the StAR-related lipid transfer<br />

(START) domain in mammals. J. Cell Sci. 118, 2791-2801.<br />

Bloj, B. <strong>and</strong> D.B.Zilversmit. 1981. Accelerated transfer of neutral glycosphingolipids <strong>and</strong><br />

ganglioside GM1 by a purified lipid transfer protein. J. Biol. Chem. 256, 5988-5991.<br />

Brodersen, P., M.Petersen, H.M.Pike, B.Olszak, S.Skov, N.Odum, L.B.Jorgensen,<br />

R.E.Brown, <strong>and</strong> J.Mundy. 2002. Knockout of Arabidopsis accelerated-cell-death11<br />

Accepted Manuscript<br />

20<br />

Page 21 of 34


encoding a sphingosine transfer protein causes activation of programmed cell death <strong>and</strong><br />

defense. Genes Dev. 16, 490-502.<br />

Brown, R.E. <strong>and</strong> P.Mattjus. 2007. Glycolipid transfer proteins. Biochim. Biophys. Acta 1771,<br />

746-760.<br />

Brown, R.E., F.A.Stephenson, T.Markello, Y.Barenholz, <strong>and</strong> T.E.Thompson. 1985. Properties<br />

of a specific <strong>glycolipid</strong> transfer protein from bovine brain. Chem. Phys. Lipids 38, 79-93.<br />

Cao, X., U.Coskun, M.Rossle, S.B.Buschhorn, M.Grzybek, T.R.Dafforn, M.Lenoir,<br />

M.Overduin, <strong>and</strong> K.Simons. 2009. Golgi protein FAPP2 tubulates membranes. Proc. Natl.<br />

Acad. Sci. U. S. A 106, 21121-21125.<br />

Chalat, M., I.Menon, Z.Turan, <strong>and</strong> A.K.Menon. 2012. Reconstitution of glucosylceramide<br />

flip-flop across endoplasmic reticulum: implications for mechanism of glycosphingolipid<br />

biosynthesis. J. Biol. Chem. 287, 15523-15532.<br />

Cho, W. 2001. Membrane targeting by C1 <strong>and</strong> C2 domains. J. Biol. Chem. 276, 32407-32410.<br />

D'Angelo, G., E.Polishchuk, T.G.Di, M.Santoro, C.A.Di, A.Godi, G.West, J.Bielawski,<br />

C.C.Chuang, A.C.van der Spoel, F.M.Platt, Y.A.Hannun, R.Polishchuk, P.Mattjus, <strong>and</strong><br />

M.A.De Matteis. 2007. Glycosphingolipid synthesis requires FAPP2 transfer of<br />

glucosylceramide. Nature 449, 62-67.<br />

D'Angelo, G., L.R.Rega, <strong>and</strong> M.A.De Matteis. 2012. Connecting vesicular transport with lipid<br />

synthesis: FAPP2. Biochim. Biophys. Acta 1821, 1089-1095.<br />

D'Angelo, G., T.Uemura, C.C.Chuang, E.Polishchuk, M.Santoro, H.Ohvo-Rekila, T.Sato,<br />

T.G.Di, A.Varriale, S.D'Auria, T.Daniele, F.Capuani, L.Johannes, P.Mattjus, M.Monti,<br />

P.Pucci, R.L.Williams, J.E.Burke, F.M.Platt, A.Harada, <strong>and</strong> M.A.De Matteis. 2013.<br />

Vesicular <strong>and</strong> non-vesicular transport feed distinct glycosylation pathways in the Golgi<br />

1. Nature 501, 116-120.<br />

D'Angelo, G., M.Vicinanza, <strong>and</strong> M.A.De Matteis. 2008. Lipid-transfer proteins in<br />

biosynthetic pathways. Curr. Opin. Cell Biol. 20, 360-370.<br />

Dowler, S., R.A.Currie, D.G.Campbell, M.Deak, G.Kular, C.P.Downes, <strong>and</strong> D.R.Alessi.<br />

2000. Identification of pleckstrin-homology-domain-containing proteins with novel<br />

phosphoinositide-binding specificities. Biochem. J. 351, 19-31.<br />

Eisenberg, D., E.Schwarz, M.Komaromy, <strong>and</strong> R.Wall. 1984. Analysis of membrane <strong>and</strong><br />

surface protein sequences with the hydrophobic moment plot. J. Mol. Biol. 179, 125-142.<br />

Fantini, J., N.Garmy, <strong>and</strong> N.Yahi. 2006. Prediction of <strong>glycolipid</strong>-binding domains from the<br />

amino acid sequence of lipid raft-associated proteins: application to HpaA, a protein<br />

involved in the adhesion of Helicobacter pylori to gastrointestinal cells<br />

4. Biochemistry 45, 10957-10962.<br />

Fugmann, T., A.Hausser, P.Schoffler, S.Schmid, K.Pfizenmaier, <strong>and</strong> M.A.Olayioye. 2007.<br />

Regulation of secretory transport by protein kinase D-mediated phosphorylation of the<br />

ceramide transfer protein. J. Cell Biol. 178, 15-22.<br />

Furuita, K., J.Jee, H.Fukada, M.Mishima, <strong>and</strong> C.Kojima. 2010. Electrostatic interaction<br />

between oxysterol-binding protein <strong>and</strong> VAMP-associated protein A revealed by NMR <strong>and</strong><br />

mutagenesis studies. J. Biol. Chem. 285, 12961-12970.<br />

Futerman, A.H. <strong>and</strong> R.E.Pagano. 1991. Determination of the intracellular sites <strong>and</strong> topology<br />

of glucosylceramide synthesis in rat liver. Biochem. J. 280 ( Pt 2), 295-302.<br />

Gaullier, J.M., A.Simonsen, A.D'Arrigo, B.Bremnes, H.Stenmark, <strong>and</strong> R.Aasl<strong>and</strong>. 1998.<br />

FYVE fingers bind PtdIns(3)P. Nature 394, 432-433.<br />

Godi, A., C.A.Di, A.Konstantakopoulos, T.G.Di, D.R.Alessi, G.S.Kular, T.Daniele, P.Marra,<br />

J.M.Lucocq, <strong>and</strong> M.A.De Matteis. 2004. FAPPs control Golgi-to-cell-surface membrane<br />

traffic by binding to ARF <strong>and</strong> PtdIns(4)P. Nat. Cell Biol. 6, 393-404.<br />

Accepted Manuscript<br />

21<br />

Page 22 of 34


Godi, A., P.Pertile, R.Meyers, P.Marra, T.G.Di, C.Iurisci, A.Luini, D.Corda, <strong>and</strong> M.A.De<br />

Matteis. 1999. ARF mediates recruitment of PtdIns-4-OH kinase-beta <strong>and</strong> stimulates<br />

synthesis of PtdIns(4,5)P2 on the Golgi complex. Nat. Cell Biol. 1, 280-287.<br />

Guan, R.-J., M.Wang, X.-Q.Liu, <strong>and</strong> D.-C.Wang. 2001. Optimization of soluble protein<br />

crystallization with detergents. Journal of Crystal Growth 231, 273-279.<br />

Haapalainen, A.M., D.M.van Aalten, G.Merilainen, J.E.Jalonen, P.Pirila, R.K.Wierenga,<br />

J.K.Hiltunen, <strong>and</strong> T.Glumoff. 2001. Crystal structure of the lig<strong>and</strong>ed SCP-2-like domain<br />

of human peroxisomal multifunctional enzyme type 2 at 1.75 A resolution. J. Mol. Biol.<br />

313, 1127-1138.<br />

Halter, D., S.Neumann, S.M.van Dijk, J.Wolthoorn, A.M.de Maziere, O.V.Vieira, P.Mattjus,<br />

J.Klumperman, G.van Meer, <strong>and</strong> H.Sprong. 2007. Pre- <strong>and</strong> post-Golgi translocation of<br />

glucosylceramide in glycosphingolipid synthesis. J. Cell Biol. 179, 101-115.<br />

Hanada, K., K.Kumagai, S.Yasuda, Y.Miura, M.Kawano, M.Fukasawa, <strong>and</strong> M.Nishijima.<br />

2003. Molecular machinery for non-vesicular trafficking of ceramide. Nature 426, 803-<br />

809.<br />

He, J., J.L.Scott, A.Heroux, S.Roy, M.Lenoir, M.Overduin, R.V.Stahelin, <strong>and</strong><br />

T.G.Kutateladze. 2011. Molecular basis of phosphatidylinositol 4-phosphate <strong>and</strong> ARF1<br />

GTPase recognition by the FAPP1 pleckstrin homology (PH) domain. J. Biol. Chem. 286,<br />

18650-18657.<br />

Jeckel, D., A.Karrenbauer, K.N.Burger, G.van Meer, <strong>and</strong> F.Wiel<strong>and</strong>. 1992. Glucosylceramide<br />

is synthesized at the cytosolic surface of various Golgi subfractions. J. Cell Biol. 117,<br />

259-267.<br />

Kaiser, S.E., J.H.Brickner, A.R.Reilein, T.D.Fenn, P.Walter, <strong>and</strong> A.T.Brunger. 2005.<br />

Structural basis of FFAT motif-mediated ER targeting. Structure. 13, 1035-1045.<br />

Kamlekar, R.K., Y.Gao, R.Kenoth, J.G.Molotkovsky, F.G.Prendergast, L.Malinina, D.J.Patel,<br />

W.S.Wessels, S.Y.Venyaminov, <strong>and</strong> R.E.Brown. 2010. Human GLTP: Three distinct<br />

functions for the three tryptophans in a novel peripheral amphitropic fold. Biophys. J. 99,<br />

2626-2635.<br />

Kamlekar, R.K., D.K.Simanshu, Y.G.Gao, R.Kenoth, H.M.Pike, F.G.Prendergast,<br />

L.Malinina, J.G.Molotkovsky, S.Y.Venyaminov, D.J.Patel, <strong>and</strong> R.E.Brown. 2013. The<br />

<strong>glycolipid</strong> transfer protein (GLTP) domain of phosphoinositol 4-phosphate adaptor<br />

protein-2 (FAPP2): structure drives preference for simple neutral glycosphingolipids.<br />

Biochim. Biophys. Acta 1831, 417-427.<br />

Kanekura, K., I.Nishimoto, S.Aiso, <strong>and</strong> M.Matsuoka. 2006. Characterization of amyotrophic<br />

lateral sclerosis-linked P56S mutation of vesicle-associated membrane protein-associated<br />

protein B (VAPB/ALS8). J. Biol. Chem. 281, 30223-30233.<br />

Kawano, M., K.Kumagai, M.Nishijima, <strong>and</strong> K.Hanada. 2006. Efficient trafficking of<br />

ceramide from the endoplasmic reticulum to the Golgi apparatus requires a VAMPassociated<br />

protein-interacting FFAT motif of CERT. J. Biol. Chem. 281, 30279-30288.<br />

Kenoth, R., D.K.Simanshu, R.K.Kamlekar, H.M.Pike, J.G.Molotkovsky, L.M.Benson,<br />

H.R.Bergen, III, F.G.Prendergast, L.Malinina, S.Y.Venyaminov, D.J.Patel, <strong>and</strong><br />

R.E.Brown. 2010. Structural determination <strong>and</strong> tryptophan fluorescence of heterokaryon<br />

incompatibility C2 protein (HET-C2), a fungal <strong>glycolipid</strong> transfer protein (GLTP),<br />

provide novel insights into <strong>glycolipid</strong> specificity <strong>and</strong> membrane interaction by the GLTP<br />

fold. J. Biol. Chem. 285, 13066-13078.<br />

Kjellberg, M.A. <strong>and</strong> P.Mattjus. 2013. Glycolipid transfer protein expression is affected by<br />

glycosphingolipid synthesis. PLoS. One. 8, e70283.<br />

Kumagai, K., M.Kawano, F.Shinkai-Ouchi, M.Nishijima, <strong>and</strong> K.Hanada. 2007. Interorganelle<br />

trafficking of ceramide is regulated by phosphorylation-dependent cooperativity between<br />

the PH <strong>and</strong> START domains of CERT. J. Biol. Chem. 282, 17758-17766.<br />

Accepted Manuscript<br />

22<br />

Page 23 of 34


Lannert, H., K.Gorgas, I.Meissner, F.T.Wiel<strong>and</strong>, <strong>and</strong> D.Jeckel. 1998. Functional organization<br />

of the Golgi apparatus in glycosphingolipid biosynthesis. Lactosylceramide <strong>and</strong><br />

subsequent glycosphingolipids are formed in the lumen of the late Golgi. J. Biol. Chem.<br />

273, 2939-2946.<br />

Lenoir, M., U.Coskun, M.Grzybek, X.Cao, S.B.Buschhorn, J.James, K.Simons, <strong>and</strong><br />

M.Overduin. 2010. Structural basis of wedging the Golgi membrane by FAPP pleckstrin<br />

homology domains. EMBO Rep. 11, 279-284.<br />

Levine, T. <strong>and</strong> C.Loewen. 2006. Inter-organelle membrane contact sites: through a glass,<br />

darkly. Curr. Opin. Cell Biol. 18, 371-378.<br />

Levine, T.P. <strong>and</strong> S.Munro. 2002. Targeting of Golgi-specific pleckstrin homology domains<br />

involves both PtdIns 4-kinase-dependent <strong>and</strong> -independent components. Curr. Biol. 12,<br />

695-704.<br />

Li, X.M., M.L.Malakhova, X.Lin, H.M.Pike, T.Chung, J.G.Molotkovsky, <strong>and</strong> R.E.Brown.<br />

2004. Human <strong>glycolipid</strong> transfer protein: probing conformation using fluorescence<br />

spectroscopy. Biochemistry 43, 10285-10294.<br />

Lin, X., P.Mattjus, H.M.Pike, A.J.Windebank, <strong>and</strong> R.E.Brown. 2000. Cloning <strong>and</strong> expression<br />

of <strong>glycolipid</strong> transfer protein from bovine <strong>and</strong> porcine brain. J. Biol. Chem. 275, 5104-<br />

5110.<br />

Lingwood, C.A. 1996. Aglycone modulation of <strong>glycolipid</strong> receptor function. Glycoconj. J. 13,<br />

495-503.<br />

Lingwood, C.A. 2011. Glycosphingolipid functions. Cold Spring Harb. Perspect. Biol. 3.<br />

Litvak, V., N.Dahan, S.Ramach<strong>and</strong>ran, H.Sabanay, <strong>and</strong> S.Lev. 2005. Maintenance of the<br />

diacylglycerol level in the Golgi apparatus by the Nir2 protein is critical for Golgi<br />

secretory function. Nat. Cell Biol. 7, 225-234.<br />

Loewen, C.J., A.Roy, <strong>and</strong> T.P.Levine. 2003. A conserved ER targeting motif in three families<br />

of lipid binding proteins <strong>and</strong> in Opi1p binds VAP. EMBO J. 22, 2025-2035.<br />

Malakhova, M.L., L.Malinina, H.M.Pike, A.T.Kanack, D.J.Patel, <strong>and</strong> R.E.Brown. 2005. Point<br />

mutational analysis of the lig<strong>and</strong>ing site in human <strong>glycolipid</strong> transfer protein:<br />

Functionality of the complex. J. Biol. Chem. 280, 26312-26320.<br />

Malinina, L., M.L.Malakhova, A.T.Kanack, M.Lu, R.Abagyan, R.E.Brown, <strong>and</strong> D.J.Patel.<br />

2006. The lig<strong>and</strong>ing of <strong>glycolipid</strong> transfer protein is controlled by <strong>glycolipid</strong> acyl<br />

structure. PLoS. Biol. 4, e362.<br />

Malinina, L., M.L.Malakhova, A.Teplov, R.E.Brown, <strong>and</strong> D.J.Patel. 2004. Structural basis for<br />

glycosphingolipid transfer specificity. Nature 430, 1048-1053.<br />

Mattjus, P. 2009. Glycolipid transfer proteins <strong>and</strong> membrane interaction. Biochim. Biophys.<br />

Acta 1788, 267-272.<br />

Mattjus, P., A.Kline, H.M.Pike, J.G.Molotkovsky, <strong>and</strong> R.E.Brown. 2002. Probing for<br />

preferential interactions among sphingolipids in bilayer vesicles using the <strong>glycolipid</strong><br />

transfer protein. Biochemistry 41, 266-273.<br />

Mattjus, P., H.M.Pike, J.G.Molotkovsky, <strong>and</strong> R.E.Brown. 2000. Charged membrane surfaces<br />

impede the protein-mediated transfer of glycosphingolipids between phospholipid<br />

bilayers. Biochemistry 39, 1067-1075.<br />

Metz, R.J. <strong>and</strong> N.S.Radin. 1980. Glucosylceramide uptake protein from spleen cytosol. J.<br />

Biol. Chem. 255, 4463-4467.<br />

Metz, R.J. <strong>and</strong> N.S.Radin. 1982. Purification <strong>and</strong> properties of a cerebroside transfer protein.<br />

J. Biol. Chem. 257, 12901-12907.<br />

Mikitova, V. <strong>and</strong> T.P.Levine. 2012. Analysis of the key elements of FFAT-like motifs<br />

identifies new proteins that potentially bind VAP on the ER, including two AKAPs <strong>and</strong><br />

FAPP2. PLoS. One. 7, e30455.<br />

Accepted Manuscript<br />

23<br />

Page 24 of 34


Neumann, S., M.Opacic, R.W.Wechselberger, H.Sprong, <strong>and</strong> M.R.Egmond. 2008. Glycolipid<br />

transfer protein: clear structure <strong>and</strong> activity, but enigmatic function. Adv. Enzyme Regul.<br />

48, 137-151.<br />

Nhek, S., M.Ngo, X.Yang, M.M.Ng, S.J.Field, J.M.Asara, N.D.Ridgway, <strong>and</strong> A.Toker. 2010.<br />

Regulation of oxysterol-binding protein Golgi localization through protein kinase D-<br />

mediated phosphorylation. Mol. Biol. Cell 21, 2327-2337.<br />

Nishimura, A.L., M.Mitne-Neto, H.C.Silva, A.Richieri-Costa, S.Middleton, D.Cascio, F.Kok,<br />

J.R.Oliveira, T.Gillingwater, J.Webb, P.Skehel, <strong>and</strong> M.Zatz. 2004. A mutation in the<br />

vesicle-trafficking protein VAPB causes late-onset spinal muscular atrophy <strong>and</strong><br />

amyotrophic lateral sclerosis. Am. J. Hum. Genet. 75, 822-831.<br />

Nylund, M., C.Fortelius, E.K.Palonen, J.G.Molotkovsky, <strong>and</strong> P.Mattjus. 2007. Membrane<br />

curvature effects on <strong>glycolipid</strong> transfer protein activity. Langmuir 23, 11726-11733.<br />

Nylund, M., M.A.Kjellberg, J.G.Molotkovsky, H.S.Byun, R.Bittman, <strong>and</strong> P.Mattjus. 2006.<br />

Molecular features of phospholipids that affect <strong>glycolipid</strong> transfer protein-mediated<br />

galactosylceramide transfer between vesicles. Biochim. Biophys. Acta 1758, 807-812.<br />

Nylund, M. <strong>and</strong> P.Mattjus. 2005. Protein mediated <strong>glycolipid</strong> transfer is inhibited FROM<br />

sphingomyelin membranes but enhanced TO sphingomyelin containing raft like<br />

membranes. Biochim. Biophys. Acta 1669, 87-94.<br />

Ohvo-Rekilä, H. <strong>and</strong> P.Mattjus. 2011. Monitoring <strong>glycolipid</strong> transfer protein activity <strong>and</strong><br />

membrane interaction with the surface plasmon resonance technique. Biochim. Biophys.<br />

Acta 1808, 47-54.<br />

Olsen, J.V., M.Vermeulen, A.Santamaria, C.Kumar, M.L.Miller, L.J.Jensen, F.Gnad, J.Cox,<br />

T.S.Jensen, E.A.Nigg, S.Brunak, <strong>and</strong> M.Mann. 2010. Quantitative phosphoproteomics<br />

reveals widespread full phosphorylation site occupancy during mitosis. Sci. Signal. 3, ra3.<br />

Patki, V., D.C.Lawe, S.Corvera, J.V.Virbasius, <strong>and</strong> A.Chawla. 1998. A functional PtdIns(3)Pbinding<br />

motif. Nature 394, 433-434.<br />

Petersen, N.H., L.V.McKinney, H.Pike, D.Hofius, A.Zakaria, P.Brodersen, M.Petersen,<br />

R.E.Brown, <strong>and</strong> J.Mundy. 2008. Human GLTP <strong>and</strong> mutant forms of ACD11 suppress cell<br />

death in the Arabidopsis acd11 mutant. FEBS J. 275, 4378-4388.<br />

Quinn, P.J. 2011. The structure of complexes between phosphatidylethanolamine <strong>and</strong><br />

glucosylceramide: a matrix for membrane rafts. Biochim. Biophys. Acta 1808, 2894-<br />

2904.<br />

Quinn, P.J. 2012. Lipid-lipid interactions in bilayer membranes: married couples <strong>and</strong> casual<br />

liaisons. Prog. Lipid Res. 51, 179-198.<br />

Rao, C.S., T.Chung, H.M.Pike, <strong>and</strong> R.E.Brown. 2005. Glycolipid transfer protein interaction<br />

with bilayer vesicles: modulation by changing lipid composition. Biophys. J. 89, 4017-<br />

4028.<br />

Saita, S., M.Shirane, T.Natume, S.Iemura, <strong>and</strong> K.I.Nakayama. 2009. Promotion of neurite<br />

extension by protrudin requires its interaction with vesicle-associated membrane proteinassociated<br />

protein. J. Biol. Chem. 284, 13766-13777.<br />

Saito, S., H.Matsui, M.Kawano, K.Kumagai, N.Tomishige, K.Hanada, S.Echigo, S.Tamura,<br />

<strong>and</strong> T.Kobayashi. 2008. Protein phosphatase 2Cepsilon is an endoplasmic reticulum<br />

integral membrane protein that dephosphorylates the ceramide transport protein CERT to<br />

enhance its association with organelle membranes. J. Biol. Chem. 283, 6584-6593.<br />

Samygina, V.R., B.Ochoa-Lizarralde, A.N.Popov, A.Cabo-Bilbao, F.Goni-de-Cerio,<br />

J.G.Molotkovsky, D.J.Patel, R.E.Brown, <strong>and</strong> L.Malinina. 2013. Structural insights into<br />

lipid-dependent reversible dimerization of human GLTP. Acta Crystallogr. D. Biol.<br />

Crystallogr. 69, 603-616.<br />

Accepted Manuscript<br />

24<br />

Page 25 of 34


Samygina, V.R., A.N.Popov, A.Cabo-Bilbao, B.Ochoa-Lizarralde, F.Goni-de-Cerio, X.Zhai,<br />

J.G.Molotkovsky, D.J.Patel, R.E.Brown, <strong>and</strong> L.Malinina. 2011. Enhanced selectivity for<br />

sulfatide by engineered human <strong>glycolipid</strong> transfer protein. Structure. 19, 1644-1654.<br />

Saravanan, R.S., E.Slabaugh, V.R.Singh, L.J.Lapidus, T.Haas, <strong>and</strong> F.Br<strong>and</strong>izzi. 2009. The<br />

targeting of the oxysterol-binding protein ORP3a to the endoplasmic reticulum relies on<br />

the plant VAP33 homolog PVA12. Plant J. 58, 817-830.<br />

Saupe, S., C.Descamps, B.Turcq, <strong>and</strong> J.Begueret. 1994. Inactivation of the Podospora<br />

anserina vegetative incompatibility locus het-c, whose product resembles a <strong>glycolipid</strong><br />

transfer protein, drastically impairs ascospore production. Proc. Natl. Acad. Sci. U. S. A<br />

91, 5927-5931.<br />

Scheffzek, K. <strong>and</strong> S.Welti. 2012. Pleckstrin homology (PH) like domains - versatile modules<br />

in protein-protein interaction platforms. FEBS Lett. 586, 2662-2673.<br />

Schnabel, D., M.Schroder, <strong>and</strong> K.S<strong>and</strong>hoff. 1991. Mutation in the sphingolipid activator<br />

protein 2 in a patient with a variant of Gaucher disease. FEBS Lett. 284, 57-59.<br />

Schulze, H., T.Kolter, <strong>and</strong> K.S<strong>and</strong>hoff. 2009. Principles of lysosomal membrane degradation:<br />

Cellular topology <strong>and</strong> biochemistry of lysosomal lipid degradation. Biochim. Biophys.<br />

Acta 1793, 674-683.<br />

Simanshu, D.K., R.K.Kamlekar, D.S.Wijesinghe, X.Zou, X.Zhai, S.K.Mishra,<br />

J.G.Molotkovsky, L.Malinina, E.H.Hinchcliffe, C.E.Chalfant, R.E.Brown, <strong>and</strong> D.J.Patel.<br />

2013. Non-vesicular trafficking by a ceramide-1-phosphate transfer protein regulates<br />

eicosanoids. Nature.<br />

Simonsen, A. <strong>and</strong> H.Stenmark. 2001. PX domains: attracted by phosphoinositides. Nat. Cell<br />

Biol. 3, E179-E182.<br />

Skehel, P.A., K.C.Martin, E.R.K<strong>and</strong>el, <strong>and</strong> D.Bartsch. 1995. A VAMP-binding protein from<br />

Aplysia required for neurotransmitter release. Science 269, 1580-1583.<br />

Stenmark, H., R.Aasl<strong>and</strong>, B.H.Toh, <strong>and</strong> A.D'Arrigo. 1996. Endosomal localization of the<br />

autoantigen EEA1 is mediated by a zinc-binding FYVE finger. J. Biol. Chem. 271,<br />

24048-24054.<br />

Tessier, M. <strong>and</strong> J.R.Woodgett. 2006. Role of the Phox homology domain <strong>and</strong> phosphorylation<br />

in activation of serum <strong>and</strong> glucocorticoid-regulated kinase-3. J. Biol. Chem. 281, 23978-<br />

23989.<br />

Tidhar, R. <strong>and</strong> A.H.Futerman. 2013. The complexity of sphingolipid biosynthesis in the<br />

endoplasmic reticulum. Biochim. Biophys. Acta 1833, 2511-2518.<br />

Tomishige, N., K.Kumagai, J.Kusuda, M.Nishijima, <strong>and</strong> K.Hanada. 2009. Casein kinase<br />

I{gamma}2 down-regulates trafficking of ceramide in the synthesis of sphingomyelin.<br />

Mol. Biol. Cell 20, 348-357.<br />

Tuuf, J. <strong>and</strong> P.Mattjus. 2007. Human <strong>glycolipid</strong> transfer protein-Intracellular localization <strong>and</strong><br />

effects on the sphingolipid synthesis. Biochim. Biophys. Acta 1771, 1353-1363.<br />

Tuuf, J., L.Wistbacka, <strong>and</strong> P.Mattjus. 2009. The <strong>glycolipid</strong> transfer protein interacts with the<br />

vesicle-associated membrane protein-associated protein VAP-A. Biochem. Biophys. Res.<br />

Commun. 388, 395-399.<br />

Vieira, O.V., P.Verkade, A.Manninen, <strong>and</strong> K.Simons. 2005. FAPP2 is involved in the<br />

transport of apical cargo in polarized MDCK cells. J. Cell Biol. 170, 521-526.<br />

Vihervaara, T., M.Jansen, R.L.Uronen, Y.Ohsaki, E.Ikonen, <strong>and</strong> V.M.Olkkonen. 2011.<br />

Cytoplasmic oxysterol-binding proteins: sterol sensors or transporters? Chem. Phys.<br />

Lipids 164, 443-450.<br />

Wagner, S.A., P.Beli, B.T.Weinert, M.L.Nielsen, J.Cox, M.Mann, <strong>and</strong> C.Choudhary. 2011. A<br />

proteome-wide, quantitative survey of in vivo ubiquitylation sites reveals widespread<br />

regulatory roles. Mol. Cell Proteomics. 10, M111.<br />

Accepted Manuscript<br />

25<br />

Page 26 of 34


Warnock, D.E., M.S.Lutz, W.A.Blackburn, W.W.Young, Jr., <strong>and</strong> J.U.Baenziger. 1994.<br />

Transport of newly synthesized glucosylceramide to the plasma membrane by a non-<br />

Golgi pathway. Proc. Natl. Acad. Sci. U. S. A 91, 2708-2712.<br />

Weir, M.L., H.Xie, A.Klip, <strong>and</strong> W.S.Trimble. 2001. VAP-A binds promiscuously to both v-<br />

<strong>and</strong> tSNAREs. Biochem. Biophys. Res. Commun. 286, 616-621.<br />

West, G., M.Nylund, J.P.Slotte, <strong>and</strong> P.Mattjus. 2006. Membrane interaction <strong>and</strong> activity of<br />

the <strong>glycolipid</strong> transfer protein. Biochim. Biophys. Acta 1748, 1732-1742.<br />

West, G., Y.Nymalm, T.T.Airenne, H.Kidron, P.Mattjus, <strong>and</strong> T.T.Salminen. 2004.<br />

Crystallization <strong>and</strong> X-ray analysis of bovine <strong>glycolipid</strong> transfer protein. Acta Crystallogr.<br />

D. Biol. Crystallogr. 60, 703-705.<br />

West, G., L.Viitanen, C.Alm, P.Mattjus, T.A.Salminen, <strong>and</strong> J.Edqvist. 2008. Identification of<br />

a glycosphingolipid transfer protein GLTP1 in Arabidopsis thaliana. FEBS J. 275, 3421-<br />

3437.<br />

Wong, M., R.E.Brown, Y.Barenholz, <strong>and</strong> T.E.Thompson. 1984. Glycolipid transfer protein<br />

from bovine brain. Biochemistry 23, 6498-6505.<br />

Wright, C.S., Q.Zhao, <strong>and</strong> F.Rastinejad. 2003. Structural analysis of lipid complexes of GM2-<br />

activator protein. J. Mol. Biol. 331, 951-964.<br />

Wyles, J.P., C.R.McMaster, <strong>and</strong> N.D.Ridgway. 2002. Vesicle-associated membrane proteinassociated<br />

protein-A (VAP-A) interacts with the oxysterol-binding protein to modify<br />

export from the endoplasmic reticulum. J. Biol. Chem. 277, 29908-29918.<br />

Yamada, K., A.Abe, <strong>and</strong> T.Sasaki. 1986. Glycolipid transfer protein from pig brain transfers<br />

<strong>glycolipid</strong>s with beta-linked sugars but not with alpha-linked sugars at the sugar-lipid<br />

linkage. Biochim. Biophys. Acta 879, 345-349.<br />

Yamada, K., A.Abe, <strong>and</strong> T.Sasaki. 1985. Specificity of the <strong>glycolipid</strong> transfer protein from<br />

pig brain. J. Biol. Chem. 260, 4615-4621.<br />

Yamada, K. <strong>and</strong> T.Sasaki. 1982a. A rat brain cytosol protein which accelerates the<br />

translocation of galactosylceramide, lactosylceramide <strong>and</strong> glucosylceramide between<br />

membranes. Biochim. Biophys. Acta 687, 195-203.<br />

Yamada, K. <strong>and</strong> T.Sasaki. 1982b. Rat liver <strong>glycolipid</strong> transfer protein. A protein which<br />

facilitates the translocation of mono- <strong>and</strong> dihexosylceramides from donor to acceptor<br />

liposomes. J. Biochem. (Tokyo) 92, 457-464.<br />

Zhai, X., M.L.Malakhova, H.M.Pike, L.M.Benson, H.R.Bergen, III, I.P.Sugar, L.Malinina,<br />

D.J.Patel, <strong>and</strong> R.E.Brown. 2009. Glycolipid acquisition by human <strong>glycolipid</strong> transfer<br />

protein dramatically alters intrinsic tryptophan fluorescence: insights into <strong>glycolipid</strong><br />

binding affinity. J. Biol. Chem. 284, 13620-13628.<br />

Zhai, X., W.E.Momsen, D.A.Malakhov, I.A.Boldyrev, M.M.Momsen, J.G.Molotkovsky,<br />

H.L.Brockman, <strong>and</strong> R.E.Brown. 2013. GLTP-fold interaction with planar<br />

phosphatidylcholine surfaces is synergistically stimulated by phosphatidic acid <strong>and</strong><br />

phosphatidylethanolamine. J. Lipid Res. 54, 1103-1113.<br />

Zou, X., T.Chung, X.Lin, M.L.Malakhova, H.M.Pike, <strong>and</strong> R.E.Brown. 2008. Human<br />

<strong>glycolipid</strong> transfer protein (GLTP) genes: organization, transcriptional status <strong>and</strong><br />

evolution. BMC. Genomics 9, 72.<br />

Zou, X., Y.Gao, V.R.Ruvolo, T.L.Gardner, P.P.Ruvolo, <strong>and</strong> R.E.Brown. 2011. Human<br />

<strong>glycolipid</strong> transfer protein gene (GLTP) expression is regulated by Sp1 <strong>and</strong> Sp3:<br />

involvement of the bioactive sphingolipid ceramide. J. Biol. Chem. 286, 1301-1311.<br />

Accepted Manuscript<br />

26<br />

Page 27 of 34


Figure legends<br />

Fig. 1. A) Schematic presentation of GLTP, FAPP1 <strong>and</strong> FAPP2 sequences, showing the<br />

important PH <strong>and</strong> <strong>glycolipid</strong> binding domains. The binding sites of PI4P within the PH<br />

domain <strong>and</strong> the FFAT-like motifs within GLTP are highlighted. B) The consensus sequence<br />

of the FFAT motif is shown in bold (Loewen et al., 2003), together with the FFAT-like<br />

sequences in GLTP <strong>and</strong> FAPP2 in italic letters (Mikitova <strong>and</strong> Levine, 2012), <strong>and</strong> a second<br />

FFAT-like motif in the GLTP sequence written in normal letters (Tuuf et al., 2009).<br />

Figure 2. Structural features of GLTP. (A) A typical all alpha-helical GLTP fold with the 8<br />

helices colored as follows; helix 1 (dark blue), helix 2 (light blue), helix 3 (dark green), helix<br />

4 (light green), helix 5 (yellow), helix 6 (dark yellow), helix 7 (orange) <strong>and</strong> helix 8 (red). The<br />

location of the FFAT-like motif is in purple on helix 1. (B) The GSL sugar-binding pocket of<br />

GLTP (1BV7) with gangliosides GM 3 bound. Note that only the glucose residue of the<br />

ganglioside GM 3 head group is shown. The GSL molecule has the sphingosine chain directed<br />

outwards <strong>and</strong> a long amide linked acyl chain inside the tunnel. Amino acid residues Asp-48,<br />

Asn-52 (above) <strong>and</strong> His-140 (below) are shown in different shads of blue, <strong>and</strong> Trp-96 in red,<br />

are all responsible for stabilizing the <strong>glycolipid</strong>s inside the structure. Additional hydrophobic<br />

interactions between amino acids aligning the interior of the tunnel <strong>and</strong> with the acyl chain(s)<br />

of the lipid further locks the lipid in position. The yellow lines represent hydrogen bonds<br />

between the three amino acids <strong>and</strong> the GSL.<br />

Figure 3. (A) Structural presentation of sphingosine-in <strong>and</strong> sphingosine-out forms of the<br />

Accepted Manuscript<br />

glycosphingolipid binding of GLTP. The far end opening of the hydrophobic cavity can be<br />

seen in the left picture (black arrow). GalCer with N-linked linoleic acid (18:2 Δ 9,12 ) <strong>and</strong> with<br />

the sphingosine chain in a sphingosine-out conformation (blue) next to GalCer with N-linked<br />

oleic acid (18:1 Δ 9 ) in a sphingosine-in conformation (yellow). (B) Hydrophobicity scale<br />

coloring of GLTP amino acids, red most hydrophobic <strong>and</strong> white hydrophilic. A putative<br />

membrane interaction domain with high hydrophobicity is indicated with a black arrow (left<br />

picture. The right picture is an 180 degree rotation, revealing a red area originating from the<br />

amino acids lining the far end of the hydrophobic tunnel (white arrow).<br />

27<br />

Page 28 of 34


Figure 4. Schematic illustration of the hypothetical GLTP (in cyan) positioning at a<br />

phosphtidylcholine membrane surface with a glucosylceramide bound to GLTP (only one<br />

membrane leaflet is shown), the right picture is an 180 degree rotation. The yellow colored<br />

amino acids indicate the location of the FFAT-like domain in GLTP. The phosphatidylcholine<br />

molecules are rendered in the van der Waals volume. Lipid coloring: black, carbon; red,<br />

oxygen; orange, phosphorus <strong>and</strong> white, hydrogen. The approximate length of a membrane<br />

leaflet is indicated.<br />

Accepted Manuscript<br />

28<br />

Page 29 of 34


Accepted Manuscrip<br />

Figure 1<br />

Page 30 of 34


Figure 2<br />

Accepted Manuscript<br />

Page 31 of 34


Accepted Manuscrip<br />

Figure 3<br />

Page 32 of 34


Accepted Manuscrip<br />

Figure 4<br />

Page 33 of 34


Accepted Manuscrip<br />

Graphical Abstract<br />

Page 34 of 34

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!