Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2016 Feb 2.
Published in final edited form as: Nat Rev Drug Discov. 2008 Sep;7(9):747–758. doi: 10.1038/nrd2659

Targeting of tetraspanin proteins — potential benefits and strategies

Martin E Hemler 1
PMCID: PMC4737550  NIHMSID: NIHMS501852  PMID: 18758472

Abstract

The tetraspanin transmembrane proteins have emerged as key players in malignancy, the immune system, during fertilization and infectious disease processes. Tetraspanins engage in a wide range of specific molecular interactions, occurring through the formation of tetraspanin-enriched microdomains (TEMs). TEMs therefore serve as a starting point for understanding how tetraspanins affect cell signalling, adhesion, morphology, motility, fusion and virus infection. An abundance of recent evidence suggests that targeting tetraspanins, for example, by monoclonal antibodies, soluble large-loop proteins or RNAi technology, should be therapeutically beneficial.


There are 33 mammalian tetraspanin proteins1,2, which are listed in the HUGO Gene Nomenclature Committee web site (see Further Information) when searching for the term “TSPAN”. Mammalian tetraspanins, as well as tetraspanins from other species, are transmembrane proteins that are abundantly expressed and are defined by their similarity in size (20–30 kDa protein core), topology (small and large outer loops, short N-terminal and C-terminal tails, four transmembrane domains) and shared structural features (FIG. 1a). These include characteristic disulphide bonds and a CCG motif in the large outer loop, hydrophilic residues within transmembrane domains, and membrane-proximal palmitoylation sites3-6. The large outer loop — called the large extracellular loop (LEL) or extracellular domain 2 (EC2) — can be further divided into a constant region, containing conserved A, B and E helices, and a variable region containing sites for specific protein–protein interactions3,4,7. Structural data from uroplakin tetraspanins (UPK1A, UPK1B)8 and from molecular models of tetraspanins9,10 indicate close packing of the four transmembrane domain helices and an overall rod-shaped structure (FIG. 1b), which is suitable for the docking of partner proteins8. In support of this model, tetraspanin extracellular domains, transmembrane domains11-13 and intracellular membrane-proximal cysteines14 can all be in close contact with neighbouring proteins. Other similarly sized proteins also contain four transmembrane domains; for example, the L6 family proteins, the connexins and the PMP22/EMP/MP20/Claudin superfamily of proteins. Although these other proteins may be tetraspans, they are not members of the tetraspanin family because they lack sequence homology and key structural features of tetraspanins.

Figure 1. Tetraspanin structural features.

Figure 1

a | Shown in this ‘unfolded’ tetraspanin are membrane-proximal palmitoylations (red), hydrophilic residues within transmembrane domains (grey balls), and extracellular loops (EC1 and EC2). EC2, also called large extracellular loop (LEL) is divided into a constant region, containing A, B and E helices, and a variable region, containing the signature tetraspanin CCG motif, two conserved disulphide bonds (red) and a third loop and disulphide bond (dashed) that appears in some tetraspanins. b | This more realistic scheme emphasizes the close packing of the four transmembrane domains, the proximity of EC1 and EC2, and the overall rod shaped structure of tetraspanins. Disulphide bonds are not shown. This scheme is based on structural results seen for uroplakin tetraspanins UPK1A and UPK1B8, and modelled for other tetraspanins9,10.

At least a few different tetraspanins are expressed on nearly all cell and tissue types1. Details regarding human tetraspanin expression on specific normal and malignant cell and tissue types can be found at Oncomine Research (see Further Information) and elsewhere15-17. Genetic evidence in a number of species, including fungi, worms, flies, mice and humans, confirms that tetraspanins exert a wide-ranging influence on the nervous system, immune system, tumours, infectious disease processes, fertilization, and development in skin and other organ systems18-23. Besides determining cell morphology, tetraspanins also modulate cell motility, invasion, fusion, adhesion strengthening, signalling and protein trafficking5,17,19. This Review will discuss the functions of specific tetraspanins, their key molecular interactions and their prospects as therapeutic targets.

Tetraspanin microdomains

Tetraspanins protrude only 4–5 nm from the plasma membrane7,8. They do not typically serve as cell-surface receptors, although the hepatitis C virus protein E2 can bind to CD81 (REF. 24), the FimH protein in uropathogenic bacteria binds to tetraspanin UPK1A25, and ligands for tetraspanin CD9 have been suggested26. Tetraspanins are best known for their ability to organize laterally into tetraspanin-enriched microdomains (TEMs). Initially, TEMs were defined biochemically based on the tendency of tetraspanin proteins and their partners to remain associated under non-stringent detergent conditions23,27. Later, they were characterized by immunoelectron microscopy as units, with an area of ~0.2 μm2 and 0.6–0.7 μm spacing, sometimes appearing on the plasma membrane28. The presence of gangliosides and cholesterol27,29,30 helps to explain the resistance of TEMs to solubilization with detergent, while suggesting similarity to lipid rafts19. However, evidence involving cholesterol depletion, detergent solubilization, sucrose density analyses and palmitoylation-site mutation19,23,31-36 clearly establishes TEMs as discrete biochemical entities, which is in contrast to the rather ambiguously defined lipid rafts.

At the core of TEMs are tetraspanins engaging in direct protein–protein interactions with both transmembrane and intracellular proteins, including the immunoglobulin superfamily members EWI-2 (also known as IGSF8, CD316) and EWI-F (also known as C9P-1, FPRP), Claudin 1, epidermal growth factor receptor (EGFR) membrane-bound ligands, integrins and Syntenin-1, to carry out various functions (BOX 1). These primary complexes include tetraspanin homodimers, which have been captured by covalent crosslinking37 and by protein crystallization7. Tetraspanins also form primary complexes with several other types of molecules (TABLE 1).

Box 1. Tetraspanin primary complexes.

Tetraspanins are involved in direct protein–protein interactions with both transmembrane and intracellular proteins.

For example, tetraspanins CD9 and CD81 can directly associate with EWI-2 and EWI-F, a pair of related cell-surface proteins in the immunoglobulin superfamily, named for a conserved Glu-Trp-Ile motif. EWI-2 negatively regulates cell motility, morphology and spreading in several cell lines81,139,165. EWI-2 also impairs orthotopic tumour growth when expressed in a glioblastoma cell line (T. V. Kolesnikova et al., personal communication). It is doubtful that EWI-2 functions without being associated with CD9 and/or CD81 because CD81 (and possibly also CD9) facilitate the cell-surface expression of EWI-2 (REF. 81). Also, EWI-2 redirects CD81 and CD9 into filopodia81 and causes CD9 to switch from homo oligomerization to EWI-2 hetero association141. A truncated form of EWI-2, called EWI-2wint, associates with CD81 and blocks hepatitis C virus entry166.

In addition, CD9 directly associates with Claudin-1 when it is not present in tight junctions, leading to prolonged Claudin-1 half-life14. Other tetraspanins (CD81, CD151) may also associate physically and functionally with Claudin-1 (REFS 14,115), but it is not yet clear whether these associations are direct. Also, CD9 can associate with heparin-binding epidermal growth factor-like growth factor (HB-EGF), and other membrane-bound ligands for the epidermal growth factor receptor (EGFR), to promote juxtacrine signalling167-169. Furthermore, CD9 can modulate proteolytic release of soluble EGFR agonists, leading to either suppression169 or stimulation170 of paracrine signalling. Tetraspanin CD81 associates with and facilitates the biosynthesis of CD19 (REF. 89), a molecule that plays a pivotal role during signalling through the B-cell receptor171. CD81 also associates closely with the integrin α4β1 (REF. 138), and regulates α4β1 adhesion strengthening under shear flow conditions.90 Tetraspanin CD151 associates directly with laminin-binding integrins (α3β1, α6β1, α6β4, α7β1)137,172, with association occurring early in biosynthesis. Indeed, CD151 associates with the integrin α3 subunit in the endoplasmic reticulum, even before α3–β1 association11,12. CD151 contributes to laminin-binding integrin-dependent adhesion strengthening, cell spreading, motility and morphology11,23,32,173-176.

Tetraspanins not only associate directly with other transmembrane proteins, but also with intracellular proteins. For example, the C-terminal cytoplasmic tail of tetraspanin CD63 associates directly with a PDZ motif in Syntenin-1, leading to a negative regulation of CD63 endocytosis157. It remains to be determined whether PDZ-binding motifs in the C termini of several other tetraspanins will also bind to proteins containing PDZ domains.

Table 1. Directly associated tetraspanin partner proteins*.

Tetraspanin Partner protein Function Refs
CD9 CD9 CD9–CD9 dimer may be a basic
structural unit
37
EWI-2
(IGSF8, CD316)
Modulates integrin-dependent cell
motility and/or spreading
13,81,139,
177
EWI-F
(C9P-1, FPRP)
Functions unknown 178,179
Claudin-1 CD9 stabilizes expression of
non-junctional Claudin-1
14
HB-EGF CD9 upregulates both diphtheria
toxin binding and mitogenic
functions of HB-EGF
128,167
CD81 CD81 CD81–CD81 dimer may be a basic
structural unit
7,37,177
EWI-2 Modulates integrin-dependent
cell motility and/or spreading;
also, CD81 supports maturation
and surface expression of EWI-2
13,81,139,
177,180
EWI-2wint Inhibits hepatitis C virus interaction
with CD81
166
EWI-F Functions unknown 178,179
CD19 CD81 supports maturation and
surface expression of CD19
89,181
α4β1 Integrin CD81 supports α4β1 integrin
adhesion strengthening
90,138
CD151 CD151 CD151–CD151 dimer may be a
basic structural unit
37
α3β1 Integrin CD151 affects integrin-dependent
adhesion and motility
53,137,138,
147,182
α6β1 Integrin CD151 affects integrin-dependent
adhesion strengthening
176
α6β4 Integrin CD151 affects integrin- dependent
adhesion and motility
45,53,147,
172,173
α7β1 Integrin Functions not yet demonstrated 172
CD63 Syntenin-1 Endocytosis regulation 157
*

This table does not include tetraspanins specialized to function in the eye (Peripherin, ROM1) or urothelial membrane (uroplakin 1a, uroplakin 1b).

EWI-2wint, EWI-2 without its N-terminal immunoglobulin domain166; HB-EGF, heparin-binding epidermal growth factor-like growth factor.

Primary complexes then join into a network of looser secondary interactions, which include not only multiple distinct tetraspanins, and many other transmembrane proteins19,38, but also signalling enzymes such as conventional protein kinase C (PKC) isoforms39,40 and phosphatidylinositol 4-kinase41,42. Of the more than 100 tetraspanin partners that have been identified, most of them fall into the category of ‘secondary’ interacting partners, mainly because direct association has not been demonstrated and/or because association is lost under more stringent detergent conditions such as 1% Triton X-100 and digitonin, as compared with resistance to milder detergents such as 1% Brij 96 and 1% Brij 99 (REF. 19).

Notably, tetraspanins and many of their partner proteins (such as integrins, EWI proteins and Claudin-1) undergo protein palmitoylation, which does not contribute to primary interactions but helps to stabilize secondary interactions within TEMs32-34. Among the 8 β and 18 α subunits that comprise the integrin family, the only subunits that undergo palmitoylation are α3, α6, α7 and β4. Interestingly, it is these same subunits that form the most robust associations with tetraspanins. A practical consequence of the enrichment for palmitoylated proteins within TEMs is that metabolic labelling with 3H-palmitate is a particularly useful tool for studying TEMs43-45.

Cholesterol can physically and functionally associate with TEMs30,46. Treatment of cells with methyl-β-cyclodextrin to deplete cholesterol reduces CD81 oligomerization46, and with more extensive treatment intact TEMs can be released into the media in microvesicles14,31. In a natural process that may be related, multivesicular bodies assemble and are then shed to form 50–100 nm vesicles, known as exosomes47. It remains to be determined whether tetraspanins have a functional role in the production or maintenance of exosomes or other types of microvesicles. In one instance, exosomes have been shown to deliver a pro-angiogenic stimulatory signal that was dependent on the presence of the tetraspanin TSPAN8 (previously known as CO-029, TM4SF3)48, emphasizing that exosome tetraspanins can be functionally active.

Targeting tetraspanin-dependent functions

So far, there has been negligible drug discovery targeting of tetraspanins, perhaps because cell-surface tetraspanins do not typically interact with soluble ligands or counter-receptors on apposing cells. In addition, many of the 33 mammalian tetraspanins have been only minimally studied. However, as more reagents are generated (such as monoclonal antibodies (mAbs), small interfering RNAs (siRNAs) and transgenic mouse models) and new assay systems become available, further examples of tetraspanins as potential therapeutic targets should emerge. Nonetheless, several tetraspanins have already been implicated in a number of normal and pathological processes with the potential to be targeted therapeutically (TABLE 2).

Table 2. Tetraspanin functions amenable to targeting*.

Tetraspanin Activity Evidence Refs
Normal physiology and non-infectious pathologies
CD9 Oocyte fertilization KO, mAb,
RNAi, sLEL
145,
183185
CD37 B-cell leukemia and lymphoma growth h-mAb,
RIT-mAb
131,186
CD81 Oocyte fertilization KO 79,187
Spinal-cord injury mAb 101
Cocaine-induced locomotion RNAi 151
B-lymphocyte adhesion strengthening KO 90
CD151 Pathological angiogenesis KO 44
Wound healing KO 100
Tumour metastasis mAb, OE 51-53
Primary tumour growth RNAi 45
Platelet aggregation, spreading, clot
retraction
KO 95
TSPAN8 (CO-029) Tumour metastasis, consumption
coagulopathy
OE 61
TSPAN32 (TSSC6) Thrombus stabilization KO 96
Infectious-disease pathologies
CD9 Canine distemper virus-induced cell–cell
fusion
mAb 121,188
Diphtheria toxin co-receptor mAb, OE 128,189
HIV-1 infection of macrophages sLEL 106
Feline immunodeficiency virus infection mAb 126
CD63 HIV-1 infection of macrophages mAb, sLEL 105,106
Intracellular chlamydial development mAb 190
CD81 Support of hepatitis C virus infectivity KO, mAb,
RNAi, sLEL
112,114
Malaria sporozoite infectivity KO, mAb 127
HIV-1 infection of macrophages sLEL 106
HTLV-induced cell–cell fusion mAb 118
CD82 HTLV transmission, cell–cell fusion OE, mAb 118,119
CD151 Porcine RRS virus infection mAb, RNAi 150
HIV-1 infection of macrophages sLEL 106
*

This table is not intended to be an exhaustive list of tetraspanin functions, but focuses on those that might be worth targeting.

Functional evidence has been obtained either by knockout (KO) mouse experiments, monoclonal antibody (mAb) inhibition, RNAi intervention (either with siRNA or shRNA), soluble large extracellular loop (sLEL) incubation, or overexpression (OE) studies.

h-mAb, humanized mAb; RIT-mAb, mAb conjugated with radioisotope; RRS, reproductive and respiratory syndrome.

Normal physiology and non-infectious diseases

Tumour progression

Tetraspanin CD151 may contribute to tumour progression at multiple levels. For example, CD151 expression is significantly increased in prostate cancer compared with benign prostate hyperplasia49. Moreover, higher levels of CD151 are associated with poor prognosis in lung and prostate cancers49,50, and overexpression of CD151 promotes metastasis in colon carcinoma and fibrosarcoma cells51. In addition, CD151 expression is increased in a subset of human breast cancer samples, particularly those of high grade and/or triple negative basal-type45. Consistent with CD151 having a functional role in cancer development, its ablation from basal-like breast cancer cells impairs both ectopic and orthotopic growth in xenograft models45. The mechanistic consequences of CD151 ablation include subcellular redistribution of α6 integrins, reorganization of integrin partner proteins and altered migration and invasion accompanied by diminished adhesion-dependent signalling through the Rho GTPase subfamily member RAC1 and focal adhesion kinase FAK45. Interestingly, anti-CD151 antibodies have been shown to inhibit spontaneous and experimental metastasis and tumour cell intravasation, but not tumour growth, in chick embryo and mouse xenograft models, or in epidermoid carcinoma, fibrosarcoma or colon carcinoma cells51-53.

Primary tumour growth and metastasis are strongly supported by CD151, not only when it is present on tumour cells but also when present in the host animal. Lewis lung carcinoma cells implanted into CD151-null mice show markedly reduced primary tumour growth44 through a mechanism that may involve diminished tumour angiogenesis in the host mouse (see below). Furthermore, host CD151 contributes to experimental metastasis. For example, in CD151-null mice, lung melanoma colonies are diminished in number but not in size, suggesting that CD151 affects the initiation of metastatic growth but not proliferation (Y. Takeda et al., personal communication). In all of these studies, CD151 mostly acts in concert with laminin-binding integrins44,45. There may be an advantage in targeting CD151, rather than the laminin-binding integrins to which it associates, as the α3β1 and α6β4 integrins are needed for normal development of skin and kidney and other tissues in 129Sv and/or C57BL6 mice54-56. By contrast, CD151 is not needed for normal development in those mouse strains44,57, although it can affect kidney function when mice are at least partly bred into the FVB mouse strain58, and it may contribute to normal human kidney and skin development59. It remains to be seen whether CD151 targeting can ablate tumour progression without adversely affecting normal physiology.

Apart from CD151, the tetraspanin TSPAN8 is also associated with tumour progression. Overexpression of TSPAN8 supports pancreatic adenocarcinoma metastasis60,61, and TSPAN8 is increased on hepatocellular carcinomas that are poorly differentiated and prone to intrahepatic spreading62. In addition, TSPAN8 is upregulated in oesophageal carcinoma and promotes invasion and migration in oesophageal carcinoma cell lines63.

Tetraspanins may also act as tumour suppressors. For example, the presence of CD9 (REFS 64-66) impairs invasion, metastasis and/or survival in many cancer cell types. CD9 may also suppress primary growth of colon carcinomas67, whereas TSPAN13 (previously known as NET-6, TM4SF13) can suppress breast carcinoma proliferation and invasion in vitro68, and CD63 may suppress melanoma growth and metastasis69. Although the latter result has been called into question owing to a cell-line contamination70, another report shows that knockdown of CD63 elevates the invasive potential of melanoma cells71, consistent with CD63 having suppressor function. CD82 (previously known as KAI1) is perhaps the most well-known tetraspanin tumour suppressor72. It was initially reported to suppress prostate cancer metastasis73. Now, at least 11 different tumour types show an inverse correlation between CD82 expression and tumour invasion and/or metastasis72. Tumour suppression activity could be related to CD82-induced redistribution and/or downregulation of other crucial cell-surface molecules including gangliosides and EGFR74, urinary plasminogen activator receptor (UPAR)75 and the α6β1 integrin76.

Angiogenesis

CD151-null mice are defective in angiogenesis in vivo, in assays involving implantation of corneal pellets, subcutaneous Matrigel plugs and subcutaneously injected tumour cells44. However, vascular development appears normal in these mice44,57,58, emphasizing that the angiogenesis defect may be restricted to pathological conditions. In support of a pro-angiogenic role for CD151, overexpression of CD151 promotes neovascularization and improves blood perfusion in a rat hindlimb ischaemia model, probably by a mechanism that involves activation of the phosphotidylinositol 3-kinase (PI3K)–AKT signalling pathway77. The absence of CD151 does not affect angiogenesis in an in vivo oxygen-induced retinopathy model44, perhaps because retinal vascularization depends more on a specialized astrocytic template78 and less on a typical laminin-containing basement membrane. CD151, which is abundant at endothelial cell–cell junctions, supports several in vitro endothelial cell functions that are of relevance to angiogenesis. These include endothelial cell invasion, chemotactic migration, cable formation, Matrigel contraction, tube formation, sprouting, and signalling through AKT and RAC144.

Another tetraspanin, TSPAN8, has also been linked to angiogenesis. When TSPAN8 is expressed on a rat carcinoma cell line, it stimulates angiogenesis in vivo, with TSPAN8 itself also becoming upregulated on newly sprouting endothelium48.

Oocyte fertilization

When CD9 and CD81 are absent from oocytes, sperm–egg fusion is completely eliminated79. Tetraspanin absence does not affect the initial sperm binding but prevents subsequent sperm–egg fusion events79. Oocyte CD9 appears to function by optimizing the length, thickness, density and radius of curvature for oocyte microvilli, which may play an active role in capturing sperm cells80. Oocyte CD9 associates closely with its major protein partners (EWI-2 and EWI-F)80, which is relevant for two reasons. First, EWI-2 can drive the localization of CD9 and CD81 into filopodia81. Second, EWI proteins may associate directly with ezrin-radixin-moesin proteins, thereby linking CD9 and other TEM proteins to the actin cytoskeleton82. So, CD9–EWI protein complexes may work together in regulating oocyte microvilli structure and function. During sperm–egg fusion in humans, tetraspanins CD9, CD81 and CD151 may all have a role, with CD9 controlling the molecular organization of key proteins, such as integrin α6β1, on the oocyte surface83. With respect to inhibiting fertilization, tetraspanins may be the most attractive targets yet described84. Indeed, multiple strategies have already been used to target oocyte tetraspanins and to inhibit fertilization (see below).

Immune cell functions

Leukocyte subsets — T and B lymphocytes, granulocytes, monocytes, macrophages, dendritic cells and natural killer cells — each express eight or more different tetraspanins15. These tetraspanins associate directly or indirectly with numerous molecules — for example, T-cell receptors, B-cell receptors, major histocompatibility complex (MHC) class I, MHC class II, CD2, CD4, CD8, CD19, EWI-2, α4β1 and α6β1 integrins — that have crucial roles in the functions of immune cells38. Excellent reviews highlight the extensive and sometimes apparently conflicting roles played by tetraspanins on immune cells15,18,38. Because of these conflicting roles, and because of the unpredictability of targeting immune cells in general, it is not yet clear whether tetraspanin targeting on immune cells will be a productive endeavour. Nonetheless, a brief overview of some possibilities are discussed.

Antibody crosslinking of several tetraspanins (for example, CD9, CD53, CD81 and CD82) can co-stimulate T cells, suggesting positive effector roles38. However, mouse knockout models of four different tetraspanins (CD81, CD37, TSPAN32 and CD151) exhibit T-cell hyperproliferation in response to in vitro stimulation, which is consistent with endogenous tetraspanins exerting negative regulatory functions57,85-87. These tetraspanins may down-modulate T-cell triggering, either by recruitment of a negative regulatory phosphatase88 or by sequestering key molecules into TEMs.

On B cells, tetraspanin CD81 is required for molecular organization and efficient collaboration between the B cell receptor and its partners (CD21, CD19 and various signalling enzymes)18. Also, CD81 is required for maturation and surface expression of CD1989, a molecule with a key role during B-cell activation. In addition, CD81 associates with the α4β1 integrin on B cells, thereby strengthening α4β1 integrin adhesion under shear flow conditions90. Other tetraspanins also have key roles on B cells. For example, CD9 suppresses the motility of intraperitoneal B1 cells91, and CD37 supports T-cell dependent B-cell responses, especially at suboptimal doses of antigen86.

Tetraspanins may also contribute to antigen presentation. There is extensive evidence for tetraspanin associations with MHC class I and MHC class II molecules18,38, whereas CD81 (on T cells and antigen presenting cells) and CD82 (on T cells) both distribute into immune synapses92,93. Furthermore, multiple tetraspanins associate with a subset of MHC class II molecules that may have distinct antigen presenting properties35, and lateral associations between heterologous MHC class II molecules (I-A and I-E) depend on CD9 (REF. 94). However, there is not yet definitive evidence for a functional role for specific tetraspanins during antigen presentation. It is thought that alternative tetraspanins are able to compensate for the effects of removing individual tetraspanins.

Platelet functions

Tetraspanin CD151 is essential for normal platelet function. Platelets from CD151-null mice show defective platelet aggregation, impaired spreading on fibrinogen and delayed clot retraction95. CD151 functions by physically associating with αIIbβ3, the major platelet integrin, leading to impaired outside–in (that is, post-ligand occupancy) signalling by the integrin95. Another tetraspanin, TSPAN32 (previously known as TSSC6, PHEMX), also associates with αIIbβ3 integrin and modulates outside–in signalling, with TSPAN32-null mouse platelets similarly showing impaired clot retraction, aggregation and spreading on fibrinogen96. In addition, TSPAN32 is required for stable platelet thrombus formation during vascular injury96. Tetraspanin CD63 also associates with αIIbβ3 integrin and may regulate platelet spreading and signalling on fibrinogen97. However, the functions of αIIbβ3 integrins in CD63-null platelets have not yet been evaluated. The αIIbβ3 integrin is a well-established therapeutic target with antagonists widely used to treat coronary thrombosis98. It remains to be determined whether integrin-associated tetraspanins, which so far have a less dramatic role, will also be useful targets in platelet thrombus formation.

Several other studies, involving mAbs, suggest additional functional roles for various tetraspanins on platelets and other types of myeloid cells15. However, interpretations are complicated by expression of the immunoglobulin-γ receptor FcγRIIa and other Fc receptors on myeloid cells, which leads to antibody crosslinking and dual signalling effects99. Nonetheless, modulation of Fc receptor signalling by tetraspanins may be relevant to immune disorders99.

Other non-infectious disease functions

CD151-null mice show significantly impaired wound healing in a skin incision model, with deficient re-epithelialization being accompanied by laminin basement membrane disorganization and failure to upregulate α6 and β4 integrin subunits at the wound site100. CD81 has a key role during experimental spinal-cord injury in rats. Intraspinal infusion of an anti-CD81 antibody enhanced motor function, and spared tissue loss at the lesion site101. It is suggested that the anti-CD81 antibody inhibits microglial proliferation and/or macrophage recruitment at the injury site101.

Infectious diseases

HIV-1

Tetraspanins are involved at multiple steps in HIV-1 infection. After HIV-1 particles are captured by dendritic cells they are sorted through compartments containing tetraspanins (CD9, CD81, CD82), which redistribute into an ‘infectious synapse’, thus enabling transfer of virus particles to CD4+ T cells102. Alternatively, virus particles endocytosed by dendritic cells into tetraspanin-enriched compartments are released within vesicles resembling exosomes, which have an enhanced capability to infect CD4+ T cells103. Although it has not yet been demonstrated whether tetraspanins are required for transfer of virus particles from dendritic cells to CD4+ T cells, they do have a role during HIV-1-induced membrane fusion of CD4+ T cells. Anti-CD9 and anti-CD81 mAbs enhance syncytia formation in T lymphoblasts, apparently by blocking (that is, reversing) the inherent syncytium suppressing effect that is exerted by endogenous T-cell CD9 and CD81 molecules104.

Tetraspanins may also have an active role during HIV-1 infection of macrophages. For example, the entry of HIV-1 R5-tropic virus is inhibited by anti-CD63 mAb105, and both R5-tropic and X4-tropic viruses can be potently inhibited by soluble LEL/EC2 domains from multiple tetraspanins (CD63, CD81, CD82, CD53 and CD151)106. Once inside macrophages, HIV-1 assembles within a novel type of intracellular membrane domain containing CD81, CD9 and CD53, with CD63 then being recruited and incorporated into virions107. Finally, HIV-1 egress may occur through TEMs, as key endosomal sorting components (TSG101 and VPS28), involved in HIV-1 budding, are recruited to TEMs upon virus expression28. However, it remains to be seen whether specific tetraspanins are required for virus assembly and egress. In this regard, despite CD63 being involved in macrophage infection105,106, and being present at sites of virus assembly and budding108, it is not required for virus production in human macrophages109. As a consequence of assembly in tetraspanin-enriched compartments, newly released HIV-1 particles show incorporation of CD63 and multiple other tetraspanins. By an unknown mechanism, these embedded tetraspanins have the potential to down-modulate subsequent HIV-1 infection at a post-attachment entry step110.

Hepatitis C virus

Hepatitis C virus (HCV) binding to CD81 is necessary, but not sufficient for HCV infection111. Early studies focused on direct binding of the virus E2 glycoprotein to a site in the soluble LEL of CD81 (REF. 24). However, later studies show that the affinity of E2 for CD81 is not predictive of HCV infection and therefore potentially misleading. For example, CD81 mutations that substantially reduce E2 binding have minimal effect on HCV infection112, whereas those reducing HCV entry do not correspondingly reduce E2 binding113. Nonetheless, CD81 is required for infection and remains an attractive target. Multiple CD81 molecules, together with various other known and unknown hepatocyte surface proteins, are suggested to engage in a number of low-affinity interactions with HCV, which are necessary for the initial steps in virus binding112,114. Claudin-1, a membrane protein that physically associates with CD81 (REFS 14,115), may collaborate with CD81 during a later step in virus entry115,116. Although Claudin-1 is typically found in tight junctions, tetraspanin–claudin complexes are not in tight junctions14,117, which suggests that HCV might engage non-junctional claudins.

Other infectious disease pathologies

Antibodies to CD82 and CD81 block human T-cell lymphotropic virus type 1 (HTLV-1)-induced syncytium formation in T lymphoblastoid cells118. Overexpression of CD82 also inhibited syncytium formation, in addition to cell-to-cell virus transmission119. Both the HTLV-1 Gag protein, which directs virion assembly and release at the plasma membrane, and the Env protein, interact with tetraspanins CD82 (REFS 119,120) and CD81 (REF. 119). Hence, TEMs may participate both in HTLV-1-induced syncytia formation as well as virus assembly and release. Studies of the canine distemper virus (CDV) show that transfection of NIH-3T3 (mouse fibroblasts) or MBDK (bovine kidney) cells with CD9 is sufficient to enable infection121. Later studies show that CD9 acts not as a virus receptor, but as a regulator of CDV haemagglutinin-dependent cell–cell fusion122. These examples of tetraspanins facilitating virus-dependent fusion events are consistent with CD9 and CD81 regulating other types of cell–cell fusion, such as seen for gametes79, myoblasts123, monocytes during osteoclastogenesis124, and monocytes forming multi-nucleated giant cells125. CD9 may also support feline immunodeficiency virus infection. Again, the tetraspanin does not affect virus binding, but instead contributes to virus assembly and/or release126.

CD81 (but not CD9) is required in hepatocytes to allow invasion by sporozoites from malaria-causing parasites Plasmodium yoelii and Plasmodium falciparum127. CD81 functions together with cholesterol in the context of TEMs to enable Plasmodium sporozoite invasion46. No interaction between CD81 and sporozoite proteins has been observed, which suggests that CD81 may be playing a co-receptor role127. Tetraspanin CD9 is a co-receptor for diphtheria toxin. CD9 does not bind directly to the toxin, but forms a membrane complex with the diphtheria toxin receptor (heparin-binding epidermal growth factor-like growth factor; HB-EGF), to increase the number of effective binding sites128.

Strategies for targeting tetraspanins

Tetraspanins might be targeted using mAbs, recombinant soluble LELs, RNA interference (RNAi) or other approaches (FIG. 2).

Figure 2. Strategies for targeting tetraspanins.

Figure 2

a | Monoclonal antibodies (mAbs) might interfere with tetraspanins by blocking lateral interactions, although this has not yet been well documented. Also, they may sequester tetraspanins, leading to disruption of tetra spanin-enriched domains (TEMs), down-modulation or triggering of apoptosis. In addition, mAbs can be used to deliver a lethal hit to cells expressing particular tetraspanins. For example, the antibody might trigger apoptosis, complement-dependent cytotoxicity or antibody-dependent cellular cytotoxicity, or deliver a conjugated toxin or lethal radioisotope or nanoparticle. b | Recombinant soluble large extracelullar loops (sLELs) most probably insert into TEMs, to disrupt lateral interactions among tetraspanins, or between tetraspanins and their partner molecules. In a special case, hepatitis C virus (HCV) infection, the CD81 sLEL acts in trans by binding to the virus and preventing interaction with cellular CD81. c | RNAi strategies are well documented to demonstrate the functional importance of tetraspanins. d | Other approaches might be useful for tetraspanin targeting. For example, a small molecule could be designed to inhibit the interaction of a C-terminal tail with a PDZ-domain protein. A peptide specifically mimicking a tetraspanin transmembrane domain might be effective in disrupting lateral interactions. Interference with the appropriate protein acyl transferase in the Golgi would prevent tetraspanin palmitoylation, leading to an inability to assemble TEMs and resulting in impaired functions.

Monoclonal antibodies

As of 2007, over 400 therapeutic mAbs had entered commercially sponsored clinical development129. Although this list includes mAbs that target tetraspanin-associated proteins, such as epithelial cell adhesion molecule (EPCAM) and EGFR, mAbs targeting tetraspanins themselves are not listed129. However, a few promising in vivo results have so far been observed. For example, mAb targeting of CD81 improves recovery from spinal injury in rats101, intrauterine injection of an anti-CD9 mAb increased embryo implantation130, intra-tumour injection of anti-CD9 mAbs inhibited in vivo colon carcinoma growth67, and a modified anti-CD37 antibody depleted human B-cell chronic lymphocytic leukaemia (B-CLL) cells and improved survival in a mouse xenograft model131. Results listed in TABLE 2 indicate the potential for additional mAb targeting of a number of tetraspanins in different clinical settings.

Anti-CD81 mAbs can block HCV infection at least in part by blocking binding of the HCV E2 protein113,114. However, in most of the other examples listed in TABLE 2, it is not clear how anti-tetraspanin mAbs alter cell functions. Theoretically, anti-tetraspanin mAbs could block lateral associations with partner proteins. However, such an effect has not yet been well documented. It is more likely that treatment of cells with multivalent anti-tetraspanin mAbs causes formation of aggregates of TEM proteins, leading to disruption of TEMs and/or down-modulation of the target tetraspanin, perhaps together with key associated proteins. Indeed, it is suggested that the ability of anti-tetraspanin mAbs to perturb large patches of TEM proteins probably explains why mAb treatment effects often exceed the magnitude of effects seen upon knockdown or knockout of individual tetraspanin proteins. For example, anti-CD63 mAb strongly inhibits HIV-1 entry into macrophages105, but RNAi ablation of CD63 had no effect on HIV-1 infectivity and virus assembly109. Similarly, anti-CD81 antibodies affect numerous cellular functions132 such as lymphocyte proliferation133 and thymocyte development in fetal organ culture134, but CD81-null mice show relatively mild phenotypes, with normal lymphocyte proliferation and thymic development85,135.

Another mode of action is that the mAb may exert an agonist effect. In this regard, a few anti-tetraspanin antibodies enhance rather than inhibit tetraspanin-dependent functions. For example, an anti-CD151 antibody can stimulate cell adhesion, thereby preventing de-adhesion and immobilizing tumour cells53. In other examples, anti-CD9 and anti-CD81 antibodies may stimulate cell–cell fusion, either by an agonist effect or by blocking inherent anti-fusion properties of CD9 and CD81 (REFS 104,125). In addition, anti-CD9 mAbs amplify the inherent tumour suppressor function of CD9 (REF. 67).

Despite tetraspanins being relatively small molecules on the cell surface (only ~100 extracellular amino acids), antibodies to the same tetraspanin can have a range of different properties136. For example, some anti-CD151 antibodies recognize total protein, whereas others selectively recognize an epitope exposed only when CD151 is not bound to laminin-binding integrins136-138. Similarly, epitopes defined by anti-CD81 and anti-CD9 mAbs vary widely46,139,140, which is partly due to differential shielding of epitopes. As a practical benefit, specialized anti-tetraspanin mAbs may have utility in diagnosing the molecular state of tetraspanins in intact cells. For example, antibodies that selectively detect unbound CD151 might provide information regarding integrin association136-138, whereas an anti-CD9 antibody might also provide information regarding integrin activation and/or association140. A low-affinity mAb that selectively binds clustered CD9 can be used to assess the changes in the state of homo-oligomeric CD9 clustering on the cell surface141. Similarly, an antibody with preference for oligomerized CD81 can provide clues regarding the cell-surface organization of CD81 (REF. 46).

Besides blocking, clustering or otherwise disrupting specific molecular targets, mAbs also can trigger the removal of target cells. This can be achieved directly by the stimulation of apoptosis in carcinoma cell lines142 or in myotubes123 with specific anti-CD81 and/or anti-CD9 antibodies. Alternatively, antibody-coated cells may undergo complement-dependent cytotoxicity (CDC; lysis due to complement fixation) or antibody-dependent cell-mediated cytotoxicity (ADCC). In the latter case, apoptosis is triggered after antibody-coated cells are recognized by Fc receptors of natural killer cells. In this regard, a modified anti-CD37 antibody triggered efficient ADCC-dependent in vivo removal of B-CLL cells131. In fact, the therapeutic efficacy of anti-CD37 may surpass that of rituximab (an anti-CD20 antibody)131, which is currently used in the treatment of human B-CLL. Finally, mAbs can be coupled to a lethal substance, such as a toxin, radioisotope or nanoparticle, which can then be used to ablate cells bearing the mAb target molecule. Although an iodine[131]-labelled anti-CD37 mAb yielded promising results in patients with B-cell lymphoma143, there has been no recent continuation of these studies. It remains to be seen whether other tetraspanins will show the sufficient degree of disease-type, cell-type or tumour specificity needed for this approach.

Soluble EC2 loops

Recombinant soluble LEL/EC2 of tetraspanin CD9 inhibited sperm–egg fusion when pre-incubated with oocytes but not sperm cells. These results suggest that CD9 may act in cis on the oocyte surface, whereas sperm lack a trans acting counter-receptor for CD9 (REF. 144). Subsequent studies confirmed these results using recombinant soluble LELs from CD9 and CD81, with CD63 soluble LEL as a negative control145. In other studies, human (but not mouse) CD9 soluble LEL inhibited blood monocyte fusion125, and soluble LELs from CD9 and CD151 disrupted endothelial cell interactions with leukocytes146. Recombinant soluble LELs have also been used effectively to inhibit virus functions. Multiple soluble LELs (from CD63, CD9, CD81 and CD151) potently inhibit CCR5-tropic HIV-1 infection of macrophages106. The soluble LEL proteins appear to act on target cells by interfering with the uptake of virions106. In several of these studies, mutant soluble LELs, or soluble LELs from another species, function less effectively, thus serving as important controls. Together, these results, obtained in several different settings, emphasize the utility of soluble LEL reagents. Curiously, HIV-1 infection of macrophages is inhibited by multiple soluble LELs (CD63, CD9, CD81 and CD151) but only by a single anti-tetraspanin mAb (anti-CD63, but not anti-CD9 or anti-CD81). One interpretation of these results is that CD63 plays a more direct role in virus infection, thus explaining mAb selectivity. By contrast, soluble LELs from multiple other tetraspanins may interfere with endogenous lateral homo-tetraspanin and heterotetraspanin oligomerizations, thereby disrupting TEMs, which indirectly impairs the function of CD63.

Recombinant soluble LEL protein from CD81 also inhibits HCV infectivity. In this case, the mechanism is more obvious. During infection, the HCV E2 protein binds directly to cellular CD81, which serves a critical co-receptor function. CD81 soluble LEL binds to the virus E2 protein, thus preventing virus interaction with endogenous CD81 LEL112. However, recent work shows that CD81 soluble LEL is minimally effective as an inhibitor in HCV assays that more closely mimic in vivo HCV infection in humans. Although CD81 is still required at an early step in the infection of normal primary human hepatocytes, viral particles appear to be less accessible to inhibitory interaction with CD81 LEL114.

In summary, soluble LELs from multiple tetraspanins have been used in a number of settings to disrupt tetraspanin-dependent functions. Although the mechanism of action has not been precisely clarified in many of these assays, it is presumed that soluble LELs can disrupt lateral interactions essential for assembly and/or maintenance of TEMs. Hence, by exerting a combination of direct and indirect effects, soluble LELs may offer a broader range of inhibitory possibilities than mAbs. However, in the case of HCV, in which the LEL binds to virus instead of cells, anti-CD81 mAbs may be more potent114.

RNAi approaches

Another strategy for interfering with tetraspanins involves small interfering (siRNA) knockdown. Knockdown of CD9, CD151 and CD81 can be readily achieved using siRNA and/or small hairpin RNA (shRNA)14,146,147. This approach shows that endothelial cell CD9 and CD151 support lymphocyte transendothelial migration146, CD9 supports integrin-dependent adhesion and reduced dissemination in ovarian carcinoma cells148, and CD151 can regulate integrin-dependent cell adhesion and migration147. Additional RNAi studies show that osteoclastogenesis is inhibited by TSPAN13 but promoted by TSPAN5 (previously known as NET-4, TM4SF9)149. In infectious disease studies, silencing of CD81 with siRNA markedly inhibits (>90%) infection of primary normal human hepatocytes by HCV, much more effectively than soluble LEL from CD81 (REF. 114). Also, knockdown of CD151 significantly reduces susceptibility to porcine reproductive and respiratory syndrome virus infection in a model cell line150. The RNAi approach has already been used to inhibit tetraspanins in at least one in vivo study. Using an inducible shRNA vector, CD81 was targeted in rat brain, resulting in suppression of cocaine-induced locomotor activity151.

The area of RNAi therapeutics is now sufficiently advanced, such that siRNAs can be made potent, specific (avoiding off-target effects) and chemically stabilized, although delivery to specific sites still needs to be optimized152. Regarding delivery, the recent use of specific mAbs to target siRNA-containing nanoparticles opens up a wide realm of new possibilities in which effects of relatively lethal siRNAs can be restricted to specific cell or tissue types153. With a number of different RNAi therapeutics already in clinical trials152, this technology should be available for application to tetraspanin targets.

Other possibilities

Tetraspanin transmembrane domains are involved in crucial intramolecular and intermolecular interactions that are essential for correct intramolecular folding and for intermolecular associations with other transmembrane proteins in TEMs10,89,154,155. A study of G-protein-coupled receptors shows that structural analogues of individual transmembrane domains can serve as potent and specific receptor antagonists156. Hence, such an approach might also be applicable to tetraspanins. For example, transmembrane domain analogues that mimick single transmembrane domains in CD81 or CD151 conceivably could interfere with HCV infection or carcinoma progression, respectively.

The C-terminal cytoplasmic tail of tetraspanin CD63 interacts with a PDZ domain in Syntenin-1 to regulate CD63 endocytosis157. Several other tetraspanins also contain C-terminal three amino-acid sequences corresponding to potential PDZ domain binding motifs. PDZ domains, with their well-defined binding sites, are promising targets for drug discovery158. As additional tetraspanin–PDZ domain interactions are uncovered, prospects for PDZ intervention as a tetraspanin targeting strategy will be enhanced.

Other specific protein–protein interactions could also be targeted. The LELs of tetraspanins are being structurally characterized7-9, and key interaction sites are being mapped, such as between CD151 and integrins137. So, it should become more feasible to disrupt these interactions with small peptides or with other small molecules. In this regard, the interaction between CD81–soluble LEL and the HCV–E2 protein has been used to screen for small-molecule inhibitors, with the intention of blocking HCV infection159,160.

Palmitoylation of tetraspanins and partner proteins has a vital role during the assembly and/or maintenance of TEMs10,32-34,43. Palmitoylation of tetraspanin CD82 is particularly important for its association with HTLV-1 Gag protein, which may support virus assembly and release120. It is only recently that progress has been made in the identification of the protein acyl transferases that are responsible for protein palmitoylation161. Targeting relevant protein acyl transferases could provide a novel means of physically disrupting TEMs, thus leading to altered cellular functions. In this regard, a protein acyl transferase, called DHHC2, promotes tetraspanin palmitoylation while supporting TEM assembly162.

It is not yet clear how to take advantage of tetraspanin tumour suppressor functions therapeutically. However, in the case of the metastasis suppressor CD82, an interesting new possibility has emerged. Because CD82 is degraded by gp78 ubiquitin ligase, targeting of gp78 increases abundance of CD82 while reducing metastasis163. Whether other tetraspanin suppressors (for example, CD9, CD63 and TSPAN13) undergo post-translational down-regulation that could be similarly targeted remains to be determined. Another possibility for amplifying tetraspanin suppression is through adenoviral transduction methods. Adenoviruses encoding CD9 or CD82 were administered, via the trachea, into mice orthotopically preimplanted with Lewis lung carcinoma cells, resulting in a dramatic reduction in metastasis65. Curiously, host-cell expression of CD9 itself may facilitate adeno-associated virus serotype 2 infection164.

Despite their abundance in humans (33 tetraspanins), flies (36 tetraspanins), fish (66 tetraspanins), worms (21 tetraspanins) and elsewhere1,2, and despite the large numbers of unbiased genetic screens that have been carried out, relatively few tetraspanins have been identified as having essential functions. Furthermore, knockout of individual tetraspanins in mice has resulted in relatively mild phenotypes. One reason for the lack of more dramatic phenotypes could be that there is substantial compensation by the remaining tetraspanins. Consequently, it is predicted that engineering dominant negative tetraspanins will yield phenotypic effects that exceed those of individual knockouts, and will more closely resemble the effects seen using mAbs and RNAi strategies in developed organisms.

Summary and future prospects

So far, there have been few examples of tetraspanin targeting in vivo. Nonetheless, potential target opportunities abound as tetraspanins contribute to tumour progression, angiogenesis, fertilization, immune-cell functions, platelet clotting, infectious diseases and other processes. As more of the 33 mammalian tetraspanins are studied, and their functions and molecular interactions identified, the list of potential targets will grow as will possibilities for intervention.

Strategies for targeting tetraspanins include the use of mAbs and recombinant soluble LELs. The mAb and soluble LEL approaches can be complementary, as some processes are inhibited by one type of reagent but not the other. However, for both mAbs and soluble LELs, interpretation of results can be handicapped by not knowing the precise mechanism of action. It is often not clear which specific molecular interactions are being perturbed. Also, results can be a little unpredictable, as the same antibodies can stimulate or inhibit different types of cell fusion, and serve as both antagonists and agonists. The RNAi approach is clear in terms of mechanism and promising examples of both in vitro and in vivo applications have emerged.

The widespread distribution of many tetraspanins on numerous cell and tissue types is a potentially important obstacle to tetraspanin targeting. For example, agents that target CD9 on oocytes or CD151 on tumour cells might also affect the normal functions of platelets and endothelial cells, which also express abundant CD9 and CD151. Similarly, targeting of CD37 on B leukemic cells could also affect CD37-dependent functions of T cells. However, the development of cell-specific delivery systems may overcome this problem. Another issue is that of compensation and/or redundancy provided by non-targeted tetraspanins. To circumvent this, it may be necessary to focus on tetraspanins that have highly specific functions or to target multiple tetraspanins at once.

Finally, as tetraspanins function in the context of TEMs, a novel type of membrane microdomain, new strategies may need to be developed. For example, strategies aimed at non-protein elements in TEMs (such as gangliosides, cholesterol or palmitoylation), at specific transmembrane domain interactions, at specific cytoplasmic tail interactions, or at specific tetraspanin–partner combinations might be worth consideration.

Acknowledgements

The author gratefully acknowledges support from the National Institutes of Health grants GM38903 and CA42368.

Glossary

Integrins

A family of cell-surface transmembrane proteins (24 mammalian members), with αβ-heterodimeric structures, which function as cell adhesion molecules.

EWI proteins

A family of four cell-surface immunoglobulin superfamily proteins, sharing a conserved glutamine-tryptophan-isoleucine (EWI) motif.

Protein palmitoylation

Post-translational acylation of a protein, typically on an intracellular cysteine residue.

Exosomes

Vesicles of 50–100 nm, enriched 10-fold to 100-fold for tetraspanins, and shed from the multivesicular bodies of cells.

Monoclonal antibody

A specific antibody produced in large quantity by a single hybrid cell clone formed in the laboratory by the fusion of a B cell with a tumour cell.

Angiogenesis

The process by which new blood vessels grow from pre-existing blood vessels.

RNA interference (RNAi)

A form of post-transcriptional gene silencing in which expression or transfection of dsRNA induces degradation, by nucleases, of the homologous endogenous transcripts, resulting in reduction or loss of gene activity.

DHHC2

A member of a family of enzymes (24 members in mammals) containing a conserved aspartate-histidine-histidine-cysteine (DHHC) motif, responsible for the S-palmitoylation of proteins.

References

  • 1.Huang S, et al. The phylogenetic analysis of tetraspanins projects the evolution of cell–cell interactions from unicellular to multicellular organisms. Genomics. 2005;86:674–684. doi: 10.1016/j.ygeno.2005.08.004. [DOI] [PubMed] [Google Scholar]
  • 2.Garcia-Espana A, et al. Appearance of new tetraspanin genes during vertebrate evolution. Genomics. 2008;91:326–334. doi: 10.1016/j.ygeno.2007.12.005. [DOI] [PubMed] [Google Scholar]
  • 3.Stipp CS, Kolesnikova TV, Hemler ME. Functional domains in tetraspanin proteins. Trends Biochem. Sci. 2003;28:106–112. doi: 10.1016/S0968-0004(02)00014-2. [DOI] [PubMed] [Google Scholar]
  • 4.Seigneuret M, Delaguillaumie A, Lagaudriere-Gesbert C, Conjeaud H. Structure of the tetraspanin main extracellular domain. A partially conserved fold with a structurally variable domain insertion. J. Biol. Chem. 2001;276:40055–40064. doi: 10.1074/jbc.M105557200. [DOI] [PubMed] [Google Scholar]
  • 5.Boucheix C, Rubinstein E. Tetraspanins. Cell. Mol. Life Sci. 2001;58:1189–1205. doi: 10.1007/PL00000933. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Maecker HT, Todd SC, Levy S. The tetraspanin superfamily: molecular facilitators. FASEB J. 1997;11:428–442. [PubMed] [Google Scholar]
  • 7.Kitadokoro K, et al. CD81 extracellular domain 3D structure: insight into the tetraspanin superfamily structural motifs. EMBO J. 2001;20:12–18. doi: 10.1093/emboj/20.1.12. The first detailed structural information for a tetraspanin LEL.
  • 8.Min G, Wang H, Sun TT, Kong XP. Structural basis for tetraspanin functions as revealed by the cryo-EM structure of uroplakin complexes at 6-Å resolution. J. Cell Biol. 2006;173:975–983. doi: 10.1083/jcb.200602086. The most detailed structural information yet available for a tetraspanin and its partner protein.
  • 9.Seigneuret M. Complete predicted three-dimensional structure of the facilitator transmembrane protein and hepatitis C virus receptor CD81: conserved and variable structural domains in the tetraspanin superfamily. Biophys. J. 2006;90:212–227. doi: 10.1529/biophysj.105.069666. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Kovalenko OV, Metcalf DG, DeGrado WF, Hemler ME. Structural organization and interactions of transmembrane domains in tetraspanin proteins. BMC Struct. Biol. 2005;5:11. doi: 10.1186/1472-6807-5-11. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Kazarov AR, Yang X, Stipp CS, Sehgal B, Hemler ME. An extracellular site on tetraspanin CD151 determines α3 and α6 integrin-dependent cellular morphology. J. Cell Biol. 2002;158:1299–1309. doi: 10.1083/jcb.200204056. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Berditchevski, et al. Analysis of the CD151-α3β1 integrin and CD151-tetraspanin interactions by mutagenesis. J. Biol. Chem. 2001;276:41165–41174. doi: 10.1074/jbc.M104041200. [DOI] [PubMed] [Google Scholar]
  • 13.Charrin S, et al. EWI-2 is a new component of the tetraspanin web in hepatocytes and lymphoid cells. Biochem. J. 2003;373:409–421. doi: 10.1042/BJ20030343. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Kovalenko OV, Yang XH, Hemler ME. A novel cysteine cross-linking method reveals a direct association between claudin-1 and tetraspanin CD9. Mol. Cell Proteomics. 2007;6:1855–1867. doi: 10.1074/mcp.M700183-MCP200. [DOI] [PubMed] [Google Scholar]
  • 15.Wright MD, Moseley GW, van Spriel AB. Tetraspanin microdomains in immune cell signalling and malignant disease. Tissue Antigens. 2004;64:533–542. doi: 10.1111/j.1399-0039.2004.00321.x. [DOI] [PubMed] [Google Scholar]
  • 16.Kishimoto T, et al. Leukocyte Typing VI. Garland; New York: 1998. pp. 1–1342. [Google Scholar]
  • 17.Berditchevski F, Odintsova E. Tetraspanins as regulators of protein trafficking. Traffic. 2007;8:89–96. doi: 10.1111/j.1600-0854.2006.00515.x. [DOI] [PubMed] [Google Scholar]
  • 18.Levy S, Shoham T. The tetraspanin web modulates immune-signalling complexes. Nature Rev. Immunol. 2005;5:136–148. doi: 10.1038/nri1548. [DOI] [PubMed] [Google Scholar]
  • 19.Hemler ME. Tetraspanin proteins mediate cellular penetration, invasion and fusion events, and define a novel type of membrane microdomain. Annu. Rev. Cell Dev. Biol. 2003;19:397–422. doi: 10.1146/annurev.cellbio.19.111301.153609. [DOI] [PubMed] [Google Scholar]
  • 20.Rubinstein E, Ziyyat A, Wolf JP, Le Naour F, Boucheix C. The molecular players of sperm–egg fusion in mammals. Semin. Cell Dev. Biol. 2006;17:254–263. doi: 10.1016/j.semcdb.2006.02.012. [DOI] [PubMed] [Google Scholar]
  • 21.Lambou K, et al. Fungi have three tetraspanin families with distinct functions. BMC Genomics. 2008;9:63. doi: 10.1186/1471-2164-9-63. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Todres E, Nardi JB, Robertson HM. The tetraspanin superfamily in insects. Insect Mol. Biol. 2000;9:581–590. doi: 10.1046/j.1365-2583.2000.00222.x. [DOI] [PubMed] [Google Scholar]
  • 23.Hemler ME. Tetraspanin functions and associated microdomains. Nature Rev. Mol. Cell Biol. 2005;6:801–811. doi: 10.1038/nrm1736. [DOI] [PubMed] [Google Scholar]
  • 24.Pileri P, et al. Binding of hepatitis C virus to CD81. Science. 1998;282:938–941. doi: 10.1126/science.282.5390.938. First evidence for HCV binding to tetraspanin CD81.
  • 25.Wu XR, Sun TT, Medina JJ. In vitro binding of type 1-fimbriated Escherichia coli to uroplakins Ia and Ib: relation to urinary tract infections. Proc. Natl Acad. Sci. USA. 1996;93:9630–9635. doi: 10.1073/pnas.93.18.9630. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Ellerman DA, Ha C, Primakoff P, Myles DG, Dveksler GS. Direct binding of the ligand PSG17 to CD9 requires a CD9 site essential for sperm–egg fusion. Mol. Biol. Cell. 2003;14:5098–5103. doi: 10.1091/mbc.E03-04-0244. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Odintsova E, et al. Gangliosides play an important role in the organisation of CD82-enriched microdomains. Biochem. J. 2006;400:315–325. doi: 10.1042/BJ20060259. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Nydegger S, Khurana S, Krementsov DN, Foti M, Thali M. Mapping of tetraspanin-enriched microdomains that can function as gateways for HIV-1. J. Cell Biol. 2006;173:795–807. doi: 10.1083/jcb.200508165. An excellent characterization of TEMs on the surface of intact cells.
  • 29.Ono M, et al. GM3 ganglioside inhibits CD9-facilitated haptotactic cell motility: coexpression of GM3 and CD9 is essential in the downregulation of tumor cell motility and malignancy. Biochemistry. 2001;40:6414–6421. doi: 10.1021/bi0101998. [DOI] [PubMed] [Google Scholar]
  • 30.Charrin S, et al. A physical and functional link between cholesterol and tetraspanins. Eur. J. Immunol. 2003;33:2479–2489. doi: 10.1002/eji.200323884. Definitive crosslinking evidence for cholesterol being a component of TEMs.
  • 31.Claas C, Stipp CS, Hemler ME. Evaluation of prototype TM4SF protein complexes and their relation to lipid rafts. J. Biol. Chem. 2001;276:7974–7984. doi: 10.1074/jbc.M008650200. First paper to compare and distinguish TEMs and lipid rafts.
  • 32.Yang X, et al. Palmitoylation of tetraspanin proteins: modulation of CD151 lateral interactions, subcellular distribution, and integrin-dependent cell morphology. Mol. Biol. Cell. 2002;13:767–781. doi: 10.1091/mbc.01-05-0275. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Berditchevski F, Odintsova E, Sawada S, Gilbert E. Expression of the palmitoylation-deficient CD151 weakens the association of α3β1 integrin with the tetraspanin-enriched microdomains and affects integrin-dependent signalling. J. Biol. Chem. 2002;277:36991–37000. doi: 10.1074/jbc.M205265200. [DOI] [PubMed] [Google Scholar]
  • 34.Charrin S, et al. Differential stability of tetraspanin/tetraspanin interactions: role of palmitoylation. FEBS Lett. 2002;516:139–144. doi: 10.1016/s0014-5793(02)02522-x. [DOI] [PubMed] [Google Scholar]
  • 35.Kropshofer H, et al. Tetraspan microdomains distinct from lipid rafts enrich select peptide-MHC class II complexes. Nature Immunol. 2002;3:61–68. doi: 10.1038/ni750. [DOI] [PubMed] [Google Scholar]
  • 36.Le Naour F, Andre M, Boucheix C, Rubinstein E. Membrane microdomains and proteomics: lessons from tetraspanin microdomains and comparison with lipid rafts. Proteomics. 2006;6:6447–6454. doi: 10.1002/pmic.200600282. [DOI] [PubMed] [Google Scholar]
  • 37.Kovalenko OV, Yang X, Kolesnikova TV, Hemler ME. Evidence for specific tetraspanin homodimers: inhibition of palmitoylation makes cysteine residues available for cross-linking. Biochem. J. 2004;377:407–417. doi: 10.1042/BJ20031037. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Tarrant JM, Robb L, van Spriel AB, Wright MD. Tetraspanins: molecular organisers of the leukocyte surface. Trends Immunol. 2003;24:610–617. doi: 10.1016/j.it.2003.09.011. [DOI] [PubMed] [Google Scholar]
  • 39.Zhang XA, Bontrager AL, Hemler ME. TM4SF proteins associate with activated PKC and Link PKC to specific β1 integrins. J. Biol. Chem. 2001;276:25005–25013. doi: 10.1074/jbc.M102156200. [DOI] [PubMed] [Google Scholar]
  • 40.Chattopadhyay N, Wang Z, Ashman LK, Brady-Kalnay SM, Kreidberg JA. α3β1 integrin-CD151, a component of the cadherin-catenin complex, regulates PTPmu expression and cell–cell adhesion. J. Cell Biol. 2003;163:1351–1362. doi: 10.1083/jcb.200306067. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Berditchevski F, Tolias KF, Wong K, Carpenter CL, Hemler ME. A novel link between integrins, TM4SF proteins (CD63, CD81) and phosphatidylinositol 4-kinase. J. Biol. Chem. 1997;272:2595–2598. doi: 10.1074/jbc.272.5.2595. [DOI] [PubMed] [Google Scholar]
  • 42.Yauch RL, Hemler ME. Specific interactions among transmembrane 4 superfamily (TM4SF) proteins and phosphatidylinositol 4-kinase. Biochem. J. 2000;351:629–637. [PMC free article] [PubMed] [Google Scholar]
  • 43.Yang X, et al. Palmitoylation supports assembly and function of integrin–tetraspanin complexes. J. Cell Biol. 2004;167:1231–1240. doi: 10.1083/jcb.200404100. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Takeda Y, et al. Deletion of tetraspanin Cd151 results in decreased pathologic angiogenesis in vivo and in vitro. Blood. 2007;109:1524–1532. doi: 10.1182/blood-2006-08-041970. First in vivo evidence for a tetraspanin affecting angiogenesis.
  • 45.Yang XH, et al. CD151 accelerates breast cancer by regulating α6 integrin functions, signaling, and molecular organization. Cancer Res. 2008;68:3204–3213. doi: 10.1158/0008-5472.CAN-07-2949. First in vivo evidence for CD151 accelerating progression of breast cancer.
  • 46.Silvie O, et al. Cholesterol contributes to the organization of tetraspanin-enriched microdomains and to CD81 -dependent infection by malaria sporozoites. J. Cell Sci. 2006;119:1992–2002. doi: 10.1242/jcs.02911. Use of a novel anti-CD81 mAb to demonstrate a coordinated role for cholesterol and CD81, on hepatocytes, during malaria infection.
  • 47.Escola JM, et al. Selective enrichment of tetraspan proteins on the internal vesicles of multivesicular endosomes and on exosomes secreted by human B-lymphocytes. J. Biol. Chem. 1998;273:20121–20127. doi: 10.1074/jbc.273.32.20121. [DOI] [PubMed] [Google Scholar]
  • 48.Gesierich S, Berezovskiy I, Ryschich E, Zoller M. Systemic induction of the angiogenesis switch by the tetraspanin D6.1A/CO-029. Cancer Res. 2006;66:7083–7094. doi: 10.1158/0008-5472.CAN-06-0391. [DOI] [PubMed] [Google Scholar]
  • 49.Ang J, Lijovic M, Ashman LK, Kan K, Frauman AG. CD151 protein expression predicts the clinical outcome of low-grade primary prostate cancer better than histologic grading: a new prognostic indicator? Cancer Epidemiol. Biomarkers Prev. 2004;13:1717–1721. [PubMed] [Google Scholar]
  • 50.Tokuhara T, et al. Clinical significance of CD151 gene expression in non-small cell lung cancer. Clin. Cancer Res. 2001;7:4109–4114. [PubMed] [Google Scholar]
  • 51.Kohno M, Hasegawa H, Miyake M, Yamamoto T, Fujita S. CD151 enhances cell motility and metastasis of cancer cells in the presence of focal adhesion kinase. Int. J. Cancer. 2002;97:336–343. doi: 10.1002/ijc.1605. [DOI] [PubMed] [Google Scholar]
  • 52.Testa JE, Brooks PC, Lin JM, Quigley JP. Eukaryotic expression cloning with an antimetastatic monoclonal antibody identifies a tetraspanin (PETA-3/CD151) as an effector of human tumor cell migration and metastasis. Cancer Res. 1999;59:3812–3820. [PubMed] [Google Scholar]
  • 53.Zijlstra A, Lewis J, Degryse B, Stuhlmann H, Quigley JP. The inhibition of tumor cell intravasation and subsequent metastasis via regulation of in vivo tumor cell motility by the tetraspanin CD151. Cancer Cell. 2008;13:221–234. doi: 10.1016/j.ccr.2008.01.031. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Kreidberg JA, et al. α3β1 integrin has a crucial role in kidney and lung organogenesis. Development. 1996;122:3537–3547. doi: 10.1242/dev.122.11.3537. [DOI] [PubMed] [Google Scholar]
  • 55.Georges-Labouesse EN, et al. Absence of the α-6 integrin leads to epidermolysis bullosa and neonatal death in mice. Nature Genet. 1996;13:370–373. doi: 10.1038/ng0796-370. [DOI] [PubMed] [Google Scholar]
  • 56.Feltri ML, Arona M, Scherer SS, Wrabetz L. Cloning and sequence of the cDNA encoding the β4 integrin subunit in rat peripheral nerve. Gene. 1997;186:299–304. doi: 10.1016/s0378-1119(96)00725-1. [DOI] [PubMed] [Google Scholar]
  • 57.Wright MD, et al. Characterization of mice lacking the tetraspanin superfamily member CD151. Mol. Cell Biol. 2004;24:5978–5988. doi: 10.1128/MCB.24.13.5978-5988.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58.Sachs N, et al. Kidney failure in mice lacking the tetraspanin CD151. J. Cell Biol. 2006;175:33–39. doi: 10.1083/jcb.200603073. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59.Karamatic Crew V, et al. CD151, the first member of the tetraspanin (TM4) superfamily detected on erythrocytes, is essential for the correct assembly of human basement membranes in kidney and skin. Blood. 2004;104:2217–2223. doi: 10.1182/blood-2004-04-1512. [DOI] [PubMed] [Google Scholar]
  • 60.Herlevsen M, Schmidt DS, Miyazaki K, Zoller M. The association of the tetraspanin D6.1A with the α6β4 integrin supports cell motility and liver metastasis formation. J. Cell Sci. 2003;116:4373–4390. doi: 10.1242/jcs.00760. [DOI] [PubMed] [Google Scholar]
  • 61.Claas C, et al. Association between rat homologue of CO-029, a metastasis-associated tetraspanin molecule and consumption coagulopathy. J. Cell Biol. 1998;141:267–280. doi: 10.1083/jcb.141.1.267. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.Kanetaka K, et al. Overexpression of tetraspanin CO-029 in hepatocellular carcinoma. J. Hepatol. 2001;35:637–642. doi: 10.1016/s0168-8278(01)00183-0. [DOI] [PubMed] [Google Scholar]
  • 63.Zhou Z, et al. TM4SF3 promotes esophageal carcinoma metastasis via upregulating ADAM12m expression. Clin. Exp.Metastasis. 2008;25:537–548. doi: 10.1007/s10585-008-9168-0. [DOI] [PubMed] [Google Scholar]
  • 64.Ikeyama S, Koyama M, Yamaoko M, Sasada R, Miyake M. Suppression of cell motility and metastasis by transfection with human motility-related protein (MRP-1/CD9) DNA. J. Exp. Med. 1993;177:1231–1237. doi: 10.1084/jem.177.5.1231. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65.Takeda T, et al. Adenoviral transduction of MRP-1/CD9 and KAI1/CD82 inhibits lymph node metastasis in orthotopic lung cancer model. Cancer Res. 2007;67:1744–1749. doi: 10.1158/0008-5472.CAN-06-3090. In vivo demonstration of potential therapeutic use of tetraspanin tumour suppressor activity.
  • 66.Saito Y, et al. Absence of CD9 enhances adhesion-dependent morphologic differentiation, survival, and matrix metalloproteinase-2 production in small cell lung cancer cells. Cancer Res. 2006;66:9557–9565. doi: 10.1158/0008-5472.CAN-06-1131. [DOI] [PubMed] [Google Scholar]
  • 67.Ovalle S, et al. The tetraspanin CD9 inhibits the proliferation and tumorigenicity of human colon carcinoma cells. Int. J. Cancer. 2007;121:2140–2152. doi: 10.1002/ijc.22902. [DOI] [PubMed] [Google Scholar]
  • 68.Huang H, Sossey-Alaoui K, Beachy SH, Geradts J. The tetraspanin superfamily member NET-6 is a new tumor suppressor gene. J. Cancer Res. Clin. Oncol. 2007;133:761–769. doi: 10.1007/s00432-007-0221-1. [DOI] [PubMed] [Google Scholar]
  • 69.Radford KJ, Mallesch J, Hersey P. Suppression of human melanoma cell growth and metastasis by the melanoma-associated antigen CD63 (ME491) Int. J. Cancer. 1995;62:631–635. doi: 10.1002/ijc.2910620523. [DOI] [PubMed] [Google Scholar]
  • 70.Moseley GW, Elliott J, Wright MD, Partridge LJ, Monk PN. Interspecies contamination of the KM3 cell line: implications for CD63 function in melanoma metastasis. Int. J. Cancer. 2003;105:613–616. doi: 10.1002/ijc.11144. [DOI] [PubMed] [Google Scholar]
  • 71.Jang HI, Lee H. A decrease in the expression of CD63 tetraspanin protein elevates invasive potential of human melanoma cells. Exp. Mol. Med. 2003;35:317–323. doi: 10.1038/emm.2003.43. [DOI] [PubMed] [Google Scholar]
  • 72.Liu WM, Zhang XA. KAI1/CD82, a tumor metastasis suppressor. Cancer Lett. 2006;240:183–194. doi: 10.1016/j.canlet.2005.08.018. [DOI] [PubMed] [Google Scholar]
  • 73.Dong J-T, et al. KAI 1, a metastasis suppressor gene for prostate cancer on human chromosome 11p11.2. Science. 1995;268:884–886. doi: 10.1126/science.7754374. [DOI] [PubMed] [Google Scholar]
  • 74.Odintsova E, Voortman J, Gilbert E, Berditchevski F. Tetraspanin CD82 regulates compartmentalisation and ligand-induced dimerization of EGFR. J. Cell Sci. 2003;116:4557–4566. doi: 10.1242/jcs.00793. [DOI] [PubMed] [Google Scholar]
  • 75.Bass R, et al. Regulation of urokinase receptor proteolytic function by the tetraspanin CD82. J. Biol. Chem. 2005;280:14811–14818. doi: 10.1074/jbc.M414189200. [DOI] [PubMed] [Google Scholar]
  • 76.He B, et al. Tetraspanin CD82 attenuates cellular morphogenesis through down-regulating integrin α6-mediated cell adhesion. J. Biol. Chem. 2004;280:3346–3354. doi: 10.1074/jbc.M406680200. [DOI] [PubMed] [Google Scholar]
  • 77.Zheng ZZ, Liu ZX. Activation of the phosphatidylinositol 3-kinase/protein kinase Akt pathway mediates CD151-induced endothelial cell proliferation and cell migration. Int. J. Biochem. Cell Biol. 2006;39:340–348. doi: 10.1016/j.biocel.2006.09.001. [DOI] [PubMed] [Google Scholar]
  • 78.Dorrell MI, Aguilar E, Friedlander M. Retinal vascular development is mediated by endothelial filopodia, a preexisting astrocytic template and specific R-cadherin adhesion. Invest. Ophthalmol. Vis. Sci. 2002;43:3500–3510. [PubMed] [Google Scholar]
  • 79.Rubinstein E, et al. Reduced fertility of female mice lacking CD81. Dev. Biol. 2006;290:351–358. doi: 10.1016/j.ydbio.2005.11.031. [DOI] [PubMed] [Google Scholar]
  • 80.Runge KE, et al. Oocyte CD9 is enriched on the microvillar membrane and required for normal microvillar shape and distribution. Dev. Biol. 2007;304:317–325. doi: 10.1016/j.ydbio.2006.12.041. [DOI] [PubMed] [Google Scholar]
  • 81.Stipp CS, Kolesnikova TV, Hemler ME. EWI-2 regulates α3β1 integrin-dependent cell functions on laminin-5. J. Cell Biol. 2003;163:1167–1177. doi: 10.1083/jcb.200309113. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 82.Sala-Valdes M, et al. EWI-2 and EWI-F link the tetraspanin web to the actin cytoskeleton through their direct association with ezrin-radixin-moesin proteins. J. Biol. Chem. 2006;281:19665–19675. doi: 10.1074/jbc.M602116200. [DOI] [PubMed] [Google Scholar]
  • 83.Ziyyat A, et al. CD9 controls the formation of clusters that contain tetraspanins and the integrin α6β1, which are involved in human and mouse gamete fusion. J. Cell Sci. 2006;119:416–424. doi: 10.1242/jcs.02730. [DOI] [PubMed] [Google Scholar]
  • 84.Primakoff P, Myles DG. Cell–cell membrane fusion during mammalian fertilization. FEBS Lett. 2007;581:2174–2180. doi: 10.1016/j.febslet.2007.02.021. [DOI] [PubMed] [Google Scholar]
  • 85.Miyazaki T, Muller U, Campbell KS. Normal development but differentially altered proliferative responses of lymphocytes in mice lacking CD81. EMBO J. 1997;16:4217–4225. doi: 10.1093/emboj/16.14.4217. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 86.Knobeloch KP, et al. Targeted inactivation of the tetraspanin CD37 impairs T-cell-dependent B-cell response under suboptimal costimulatory conditions. Mol. Cell Biol. 2000;20:5363–5369. doi: 10.1128/mcb.20.15.5363-5369.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 87.Tarrant JM, et al. The absence of Tssc6, a member of the tetraspanin superfamily, does not affect lymphoid development but enhances in vitro T-cell proliferative responses. Mol. Cell Biol. 2002;22:5006–5018. doi: 10.1128/MCB.22.14.5006-5018.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 88.van Spriel AB, et al. A regulatory role for CD37 in T cell proliferation. J. Immunol. 2004;172:2953–2961. doi: 10.4049/jimmunol.172.5.2953. [DOI] [PubMed] [Google Scholar]
  • 89.Shoham T, Rajapaksa R, Kuo CC, Haimovich J, Levy S. Building of the tetraspanin web: distinct structural domains of CD81 function in different cellular compartments. Mol. Cell Biol. 2006;26:1373–1385. doi: 10.1128/MCB.26.4.1373-1385.2006. Comprehensive domain-swapping reveals distinct functions for different domains within CD81.
  • 90.Feigelson SW, Grabovsky V, Shamri R, Levy S, Alon R. The CD81 tetraspanin facilitates instantaneous leukocyte VLA-4 adhesion strengthening to vascular cell adhesion molecule 1 (VCAM-1) under shear flow. J. Biol. Chem. 2003;278:51203–51212. doi: 10.1074/jbc.M303601200. Strongly supports the concept of tetraspanins as modulators of integrin-dependent adhesion strengthening, rather than initial ligand binding.
  • 91.Ha SA, et al. Regulation of B1 cell migration by signals through Toll-like receptors. J. Exp. Med. 2006;203:2541–2550. doi: 10.1084/jem.20061041. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 92.Mittelbrunn M, Yanez-Mo M, Sancho D, Ursa A, Sanchez-Madrid F. Cutting Edge: dynamic redistribution of tetraspanin CD81 at the central zone of the immune synapse in both T lymphocytes and APC. J. Immunol. 2002;169:6691–6695. doi: 10.4049/jimmunol.169.12.6691. [DOI] [PubMed] [Google Scholar]
  • 93.Mazurov D, Heidecker G, Derse D. HTLV-1 Gag protein associates with CD82 tetraspanin microdomains at the plasma membrane. Virology. 2006;346:194–204. doi: 10.1016/j.virol.2005.10.033. [DOI] [PubMed] [Google Scholar]
  • 94.Unternaehrer JJ, Chow A, Pypaert M, Inaba K, Mellman I. The tetraspanin CD9 mediates lateral association of MHC class II molecules on the dendritic cell surface. Proc. Natl Acad. Sci. USA. 2007;104:234–239. doi: 10.1073/pnas.0609665104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 95.Lau LM, et al. The tetraspanin superfamily member, CD151 regulates outside-in integrin αIIbβ3 signalling and platelet function. Blood. 2004;104:2368–2375. doi: 10.1182/blood-2003-12-4430. [DOI] [PubMed] [Google Scholar]
  • 96.Goschnick MW, et al. Impaired “outside-in” integrin αIIbβ3 signaling and thrombus stability in TSSC6-deficient mice. Blood. 2006;108:1911–1918. doi: 10.1182/blood-2006-02-004267. [DOI] [PubMed] [Google Scholar]
  • 97.Israels SJ, McMillan-Ward EM. CD63 modulates spreading and tyrosine phosphorylation of platelets on immobilized fibrinogen. Thromb. Haemost. 2005;93:311–318. doi: 10.1160/TH04-08-0503. [DOI] [PubMed] [Google Scholar]
  • 98.Tricoci P, Peterson ED. The evolving role of glycoprotein IIb/IIIa inhibitor therapy in contemporary care of acute coronary syndrome patients. J. Interv. Cardiol. 2006;19:449–455. doi: 10.1111/j.1540-8183.2006.00182.x. [DOI] [PubMed] [Google Scholar]
  • 99.Moseley GW. Tetraspanin–Fc receptor interactions. Platelets. 2005;16:3–12. doi: 10.1080/09537100400004363. [DOI] [PubMed] [Google Scholar]
  • 100.Cowin AJ, et al. Wound healing is defective in mice lacking tetraspanin CD151. J. Invest. Dermatol. 2006;126:680–689. doi: 10.1038/sj.jid.5700142. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 101.Dijkstra S, et al. Intraspinal administration of an antibody against CD81 enhances functional recovery and tissue sparing after experimental spinal cord injury. Exp. Neurol. 2006;202:57–66. doi: 10.1016/j.expneurol.2006.05.011. One of the few examples of an anti-tetraspanin mAb used successfully in vivo in a disease model.
  • 102.Garcia E, et al. HIV-1 trafficking to the dendritic cell–T-cell infectious synapse uses a pathway of tetraspanin sorting to the immunological synapse. Traffic. 2005;6:488–501. doi: 10.1111/j.1600-0854.2005.00293.x. [DOI] [PubMed] [Google Scholar]
  • 103.Wiley RD, Gummuluru S. Immature dendritic cell-derived exosomes can mediate HIV-1 trans infection. Proc. Natl Acad. Sci. USA. 2006;103:738–743. doi: 10.1073/pnas.0507995103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 104.Gordon-Alonso M, et al. Tetraspanins CD9 and CD81 modulate HIV-1-induced membrane fusion. J. Immunol. 2006;177:5129–5137. doi: 10.4049/jimmunol.177.8.5129. [DOI] [PubMed] [Google Scholar]
  • 105.von Lindern JJ, et al. Potential role for CD63 in CCR5-mediated human immunodeficiency virus type 1 infection of macrophages. J. Virol. 2003;77:3624–3633. doi: 10.1128/JVI.77.6.3624-3633.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 106.Ho SH, et al. Recombinant extracellular domains of tetraspanin proteins are potent inhibitors of the infection of macrophages by human immunodeficiency virus type 1. J. Virol. 2006;80:6487–6496. doi: 10.1128/JVI.02539-05. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 107.Deneka M, Pelchen-Matthews A, Byland R, Ruiz-Mateos E, Marsh M. In macrophages, HIV-1 assembles into an intracellular plasma membrane domain containing the tetraspanins CD81, CD9, and CD53. J. Cell Biol. 2007;177:329–341. doi: 10.1083/jcb.200609050. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 108.Pelchen-Matthews A, Kramer B, Marsh M. Infectious HIV-1 assembles in late endosomes in primary macrophages. J. Cell Biol. 2003;162:443–455. doi: 10.1083/jcb.200304008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 109.Ruiz-Mateos E, Pelchen-Matthews A, Deneka M, Marsh M. CD63 is not required for the production of infectious human immunodeficiency virus type 1 in human macrophages. J. Virol. 2008;82:4751–4761. doi: 10.1128/JVI.02320-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 110.Sato K, et al. Modulation of human immunodeficiency virus type 1 infectivity through incorporation of tetraspanin proteins. J. Virol. 2008;82:1021–1033. doi: 10.1128/JVI.01044-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 111.Cocquerel L, Voisset C, Dubuisson J. Hepatitis C virus entry: potential receptors and their biological functions. J. Gen. Virol. 2006;87:1075–1084. doi: 10.1099/vir.0.81646-0. [DOI] [PubMed] [Google Scholar]
  • 112.Flint M, et al. Diverse CD81 proteins support hepatitis C virus infection. J. Virol. 2006;80:11331–11342. doi: 10.1128/JVI.00104-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 113.Bertaux C, Dragic T. Different domains of CD81 mediate distinct stages of hepatitis C virus pseudoparticle entry. J. Virol. 2006;80:4940–4948. doi: 10.1128/JVI.80.10.4940-4948.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 114.Molina S, et al. Serum-derived hepatitis C virus infection of primary human hepatocytes is tetraspanin CD81 dependent. J. Virol. 2008;82:569–574. doi: 10.1128/JVI.01443-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 115.Harris HJ, et al. CD81 and Claudin 1 co-receptor association: a role in hepatitis C virus entry. J. Virol. 2008;82:5007–5020. doi: 10.1128/JVI.02286-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 116.Evans MJ, et al. Claudin-1 is a hepatitis C virus co-receptor required for a late step in entry. Nature. 2007;446:801–805. doi: 10.1038/nature05654. [DOI] [PubMed] [Google Scholar]
  • 117.Yang W, et al. Correlation of the tight junction-like distribution of Claudin-1 to the cellular tropism of hepatitis C virus. J. Biol. Chem. 2008;283:8643–8653. doi: 10.1074/jbc.M709824200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 118.Fukudome K, et al. Identification of membrane antigen C33 recognized by monoclonal antibodies inhibitory to human T-cell leukemia virus type 1 (HTLV-1)-induced syncytium formation: altered glycosylation of C33 antigen in HTLV-1-positive T cells. J. Virol. 1992;66:1394–1401. doi: 10.1128/jvi.66.3.1394-1401.1992. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 119.Pique C, et al. Interaction of CD82 tetraspanin proteins with HTLV-1 envelope glycoproteins inhibits cell-to-cell fusion and virus transmission. Virology. 2000;276:455–465. doi: 10.1006/viro.2000.0538. [DOI] [PubMed] [Google Scholar]
  • 120.Mazurov D, Heidecker G, Derse D. The inner loop of tetraspanins CD82 and CD81 mediates interactions with human T cell lymphotrophic virus type 1 Gag protein. J. Biol. Chem. 2007;282:3896–3903. doi: 10.1074/jbc.M607322200. [DOI] [PubMed] [Google Scholar]
  • 121.Loffler S, et al. CD9, a tetraspan transmembrane protein, renders cells susceptible to canine distemper virus. J. Virol. 1997;71:42–49. doi: 10.1128/jvi.71.1.42-49.1997. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 122.Singethan K, et al. CD9-dependent regulation of Canine distemper virus-induced cell–cell fusion segregates with the extracellular domain of the haemagglutinin. J. Gen. Virol. 2006;87:1635–1642. doi: 10.1099/vir.0.81629-0. [DOI] [PubMed] [Google Scholar]
  • 123.Tachibana I, Hemler ME. Role of transmembrane-4 superfamily (TM4SF) proteins CD9 and CD81 in muscle cell fusion and myotube maintenance. J. Cell Biol. 1999;146:893–904. doi: 10.1083/jcb.146.4.893. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 124.Tanio Y, Yamazaki H, Kunisada T, Miyake K, Hayashi SI. CD9 molecule expressed on stromal cells is involved in osteoclastogenesis. Exp. Hematol. 1999;27:853–859. doi: 10.1016/s0301-472x(99)00011-9. [DOI] [PubMed] [Google Scholar]
  • 125.Takeda Y, et al. Tetraspanins CD9 and CD81 function to prevent the fusion of mononuclear phagocytes. J. Cell Biol. 2003;161:945–956. doi: 10.1083/jcb.200212031. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 126.de Parseval A, Lerner DL, Borrow P, Willett B, Elder JH. Blocking of feline immunodeficiency virus infection by a monoclonal antibody to CD9 is via inhibition of virus release rather than interference with receptor binding. J. Virol. 1997;71:5742–5749. doi: 10.1128/jvi.71.8.5742-5749.1997. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 127.Silvie O, et al. Hepatocyte CD81 is required for Plasmodium falciparum and Plasmodium yoelii sporozoite infectivity. Nature Med. 2003;9:93–96. doi: 10.1038/nm808. [DOI] [PubMed] [Google Scholar]
  • 128.Iwamoto R, et al. Heparin-binding EGF-like growth factor, which acts as a diphtheria toxin receptor, forms a complex with membrane protein DRAP27/CD9, which upregulates functional receptors and diphtheria toxin sensitivity. EMBO J. 1994;13:2322–2330. doi: 10.1002/j.1460-2075.1994.tb06516.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 129.Reichert JM, Valge-Archer VE. Development trends for monoclonal antibody cancer therapeutics. Nature Rev. Drug Discov. 2007;6:349–356. doi: 10.1038/nrd2241. [DOI] [PubMed] [Google Scholar]
  • 130.Liu WM, et al. Tetraspanin CD9 regulates invasion during mouse embryo implantation. J. Mol. Endocrinol. 2006;36:121–130. doi: 10.1677/jme.1.01910. [DOI] [PubMed] [Google Scholar]
  • 131.Zhao X, et al. Targeting CD37-positive lymphoid malignancies with a novel engineered small modular immunopharmaceutical. Blood. 2007;110:2569–2577. doi: 10.1182/blood-2006-12-062927. One of the few examples of successful tetraspanin targeting in vivo.
  • 132.Levy S, Todd SC, Maecker HT. CD81 (TAPA-1): a molecule involved in signal transduction and cell adhesion in the immune system. Annu. Rev. Immunol. 1998;16:89–109. doi: 10.1146/annurev.immunol.16.1.89. [DOI] [PubMed] [Google Scholar]
  • 133.Oren R, Takahashi S, Doss C, Levy R, Levy S. TAPA-1, the target of an antiproliferative antibody, defines a new family of transmembrane proteins. Mol. Cell Biol. 1990;10:4007–4015. doi: 10.1128/mcb.10.8.4007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 134.Boismenu R, Rhein M, Fischer WH, Havran WL. A role for CD81 in early T cell development. Science. 1996;271:198–200. doi: 10.1126/science.271.5246.198. [DOI] [PubMed] [Google Scholar]
  • 135.Tsitsikov EN, Gutierrez-Ramos JC, Geha RS. Impaired CD19 expression and signaling, enhanced antibody response to type II T independent antigen and reduction of B-1 cells in CD81-deficient mice. Proc. Natl Acad. Sci. USA. 1997;94:10844–10849. doi: 10.1073/pnas.94.20.10844. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 136.Geary SM, Cambareri AC, Sincock PM, Fitter S, Ashman LK. Differential tissue expression of epitopes of the tetraspanin CD151 recognised by monoclonal antibodies. Tissue Antigens. 2001;58:141–153. doi: 10.1034/j.1399-0039.2001.580301.x. [DOI] [PubMed] [Google Scholar]
  • 137.Yauch RL, Kazarov AR, Desai B, Lee RT, Hemler ME. Direct extracellular contact between integrin α3β1 and TM4SF protein CD151. J. Biol. Chem. 2000;275:9230–9238. doi: 10.1074/jbc.275.13.9230. [DOI] [PubMed] [Google Scholar]
  • 138.Serru V, et al. Selective tetraspan-integrin complexes (CD81/α4β1, CD151/α3β1, CD151/α6β1) under conditions disrupting tetraspan interactions. Biochem. J. 1999;340:103–111. [PMC free article] [PubMed] [Google Scholar]
  • 139.Kolesnikova TV, et al. EWI-2 modulates lymphocyte integrin α4β1 functions. Blood. 2004;103:3013–3019. doi: 10.1182/blood-2003-07-2201. [DOI] [PubMed] [Google Scholar]
  • 140.Gutierrez-Lopez MD, et al. A functionally relevant conformational epitope on the CD9 tetraspanin depends on the association with activated β1 integrin. J. Biol. Chem. 2003;278:208–218. doi: 10.1074/jbc.M207805200. [DOI] [PubMed] [Google Scholar]
  • 141.Yang XH, et al. Contrasting Effects of EWI proteins, integrins, and protein palmitoylation on cell surface CD9 organization. J. Biol. Chem. 2006;281:12976–12985. doi: 10.1074/jbc.M510617200. [DOI] [PubMed] [Google Scholar]
  • 142.Murayama Y, et al. CD9-mediated activation of the p46 Shc isoform leads to apoptosis in cancer cells. J. Cell Sci. 2004;117:3379–3388. doi: 10.1242/jcs.01201. [DOI] [PubMed] [Google Scholar]
  • 143.Press OW, et al. Treatment of refractory non-Hodgkin’s lymphoma with radiolabeled MB-1 (anti-CD37) antibody. J. Clin. Oncol. 1989;7:1027–1038. doi: 10.1200/JCO.1989.7.8.1027. [DOI] [PubMed] [Google Scholar]
  • 144.Zhu GZ, et al. Residues SFQ (173–175) in the large extracellular loop of CD9 are required for gamete fusion. Development. 2002;129:1995–2002. doi: 10.1242/dev.129.8.1995. [DOI] [PubMed] [Google Scholar]
  • 145.Higginbottom A, et al. Structural requirements for the inhibitory action of the CD9 large extracellular domain in sperm/oocyte binding and fusion. Biochem. Biophys. Res. Commun. 2003;311:208–214. doi: 10.1016/j.bbrc.2003.09.196. [DOI] [PubMed] [Google Scholar]
  • 146.Barreiro O, et al. Endothelial tetraspanin microdomains regulate leukocyte firm adhesion during extravasation. Blood. 2005;105:2852–2861. doi: 10.1182/blood-2004-09-3606. [DOI] [PubMed] [Google Scholar]
  • 147.Winterwood NE, Varzavand A, Meland MN, Ashman LK, Stipp CS. A critical role for tetraspanin CD151 in α3β1 and α6β4 integrin-dependent tumor cell functions on laminin-5. Mol. Biol. Cell. 2006;17:2707–2721. doi: 10.1091/mbc.E05-11-1042. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 148.Furuya M, et al. Down-regulation of CD9 in human ovarian carcinoma cell might contribute to peritoneal dissemination: morphologic alteration and reduced expression of β1 integrin subsets. Cancer Res. 2005;65:2617–2625. doi: 10.1158/0008-5472.CAN-04-3123. [DOI] [PubMed] [Google Scholar]
  • 149.Iwai K, Ishii M, Ohshima S, Miyatake K, Saeki Y. Expression and function of transmembrane-4 superfamily (tetraspanin) proteins in osteoclasts: reciprocal roles of Tspan-5 and NET-6 during osteoclastogenesis. Allergol. Int. 2007;56:457–463. doi: 10.2332/allergolint.O-07-488. [DOI] [PubMed] [Google Scholar]
  • 150.Shanmukhappa K, Kim JK, Kapil S. Role of CD151, A tetraspanin, in porcine reproductive and respiratory syndrome virus infection. Virol. J. 2007;4:62. doi: 10.1186/1743-422X-4-62. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 151.Bahi A, Boyer F, Kolira M, Dreyer JL. In vivo gene silencing of CD81 by lentiviral expression of small interference RNAs suppresses cocaine-induced behaviour. J. Neurochem. 2005;92:1243–1255. doi: 10.1111/j.1471-4159.2004.02961.x. [DOI] [PubMed] [Google Scholar]
  • 152.de Fougerolles A, Vornlocher HP, Maraganore J, Lieberman J. Interfering with disease: a progress report on siRNA-based therapeutics. Nature Rev. Drug Discov. 2007;6:443–453. doi: 10.1038/nrd2310. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 153.Peer D, Park EJ, Morishita Y, Carman CV, Shimaoka M. Systemic leukocyte-directed siRNA delivery revealing cyclin D1 as an anti-inflammatory target. Science. 2008;319:627–630. doi: 10.1126/science.1149859. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 154.Cannon KS, Cresswell P. Quality control of transmembrane domain assembly in the tetraspanin CD82. EMBO J. 2001;20:2443–2453. doi: 10.1093/emboj/20.10.2443. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 155.Goldberg AF, et al. An intramembrane glutamic acid governs peripherin/rds function for photoreceptor disk morphogenesis. Invest. Ophthalmol. Vis. Sci. 2007;48:2975–2986. doi: 10.1167/iovs.07-0049. [DOI] [PubMed] [Google Scholar]
  • 156.Tarasova NI, Rice WG, Michejda CJ. Inhibition of G-protein-coupled receptor function by disruption of transmembrane domain interactions. J. Biol. Chem. 1999;274:34911–34915. doi: 10.1074/jbc.274.49.34911. [DOI] [PubMed] [Google Scholar]
  • 157.Latysheva N, et al. Syntenin-1 is a new component of tetraspanin-enriched microdomains: mechanisms and consequences of the interaction of syntenin-1 with CD63. Mol. Cell Biol. 2006;26:7707–7718. doi: 10.1128/MCB.00849-06. First demonstration of a specific tetraspanin interaction with a PDZ domain-containing protein.
  • 158.Dev KK. Making protein interactions druggable: targeting PDZ domains. Nature Rev. Drug Discov. 2004;3:1047–1056. doi: 10.1038/nrd1578. [DOI] [PubMed] [Google Scholar]
  • 159.Holzer M, et al. Identification of terfenadine as an inhibitor of human CD81-receptor HCV-E2 interaction: synthesis and structure optimization. Molecules. 2008;13:1081–1110. doi: 10.3390/molecules13051081. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 160.VanCompernolle SE, et al. Small molecule inhibition of hepatitis C virus E2 binding to CD81. Virology. 2003;314:371–380. doi: 10.1016/s0042-6822(03)00406-9. [DOI] [PubMed] [Google Scholar]
  • 161.Mitchell DA, Vasudevan A, Linder ME, Deschenes RJ. Protein palmitoylation by a family of DHHC protein S-acyltransferases. J. Lipid Res. 2006;47:1118–1127. doi: 10.1194/jlr.R600007-JLR200. [DOI] [PubMed] [Google Scholar]
  • 162.Sharma C, Yang XH, Hemler ME. DHHC2 affects palmitoylation and stability of tetraspanins CD9 and CD151. Mol. Biol. Cell. 2008;19:3415–3425. doi: 10.1091/mbc.E07-11-1164. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 163.Tsai YC, et al. The ubiquitin ligase gp78 promotes sarcoma metastasis by targeting KAI1 for degradation. Nature Med. 2007;13:1504–1509. doi: 10.1038/nm1686. [DOI] [PubMed] [Google Scholar]
  • 164.Kurzeder C, et al. CD9 promotes adeno-associated virus type 2 infection of mammary carcinoma cells with low cell surface expression of heparan sulphate proteoglycans. Int. J. Mol. Med. 2007;19:325–333. [PubMed] [Google Scholar]
  • 165.Zhang XA, Lane WS, Charrin S, Rubinstein E, Liu L. EWI2/PGRL associates with the metastasis suppressor KAI1/CD82 and inhibits the migration of prostate cancer cells. Cancer Res. 2003;63:2665–2674. [PubMed] [Google Scholar]
  • 166.Rocha-Perugini V, et al. The CD81 partner EWI-2wint inhibits hepatitis C virus entry. PLoS ONE. 2008;3:e1866. doi: 10.1371/journal.pone.0001866. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 167.Higashiyama S, et al. The membrane protein CD9/DRAP27 potentiates the juxtacrine growth factor activity of the membrane-anchored heparin-binding EGF-like growth factor. J. Cell Biol. 1995;128:929–938. doi: 10.1083/jcb.128.5.929. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 168.Inui S, et al. Possible role of coexpression of CD9 with membrane-anchored heparin-binding EGF-like growth factor and amphiregulin in cultured human keratinocyte growth. J. Cell Physiol. 1997;171:291–298. doi: 10.1002/(SICI)1097-4652(199706)171:3<291::AID-JCP7>3.0.CO;2-J. [DOI] [PubMed] [Google Scholar]
  • 169.Shi W, Fan H, Shum L, Derynck R. The tetraspanin CD9 associates with transmembrane TGF-α and regulates TGF-α-induced EGF receptor activation and cell proliferation. J. Cell Biol. 2000;148:591–602. doi: 10.1083/jcb.148.3.591. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 170.Yan Y, Shirakabe K, Werb Z. The metalloprotease Kuzbanian (ADAM10) mediates the transactivation of EGF receptor by G. protein-coupled receptors. J. Cell Biol. 2002;158:221–226. doi: 10.1083/jcb.200112026. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 171.Fearon DT, Carter RH. The CD19/CR2/TAPA-1 complex of B lymphocytes: linking natural and acquired immunity. Annu. Rev. Immunol. 1995;13:127–149. doi: 10.1146/annurev.iy.13.040195.001015. [DOI] [PubMed] [Google Scholar]
  • 172.Sterk LM, et al. Association of the tetraspanin CD151 with the laminin-binding integrins α3β1, α6β1, α6β4 and α7β1 in cells in culture and in vivo. J. Cell Sci. 2002;115:1161–1173. doi: 10.1242/jcs.115.6.1161. Excellent use of diagnostic anti-CD151 antibodies to distinguish integrin-associated and non-associated CD151.
  • 173.Sincock PM, et al. PETA-3/CD151, a member of the transmembrane 4 superfamily, is localised to the plasma membrane and endocytic system of endothelial cells, associates with multiple integrins and modulates cell function. J. Cell Sci. 1999;112:833–844. doi: 10.1242/jcs.112.6.833. [DOI] [PubMed] [Google Scholar]
  • 174.Yanez-Mo M, et al. Regulation of endothelial cell motility by complexes of tetraspan molecules CD81/TAPA-1 and CD151/PETA-3 with α3β1 integrin localized at endothelial lateral junctions. J. Cell Biol. 1998;141:791–804. doi: 10.1083/jcb.141.3.791. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 175.Zhang XA, et al. Function of the tetraspanin CD151–α6β1 integrin complex during cellular morphogenesis. Mol. Biol. Cell. 2002;13:1–11. doi: 10.1091/mbc.01-10-0481. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 176.Lammerding J, Kazarov AR, Huang H, Lee RT, Hemler ME. Tetraspanin CD151 regulates α6β1 integrin adhesion strengthening. Proc. Natl Acad. Sci. USA. 2003;100:7616–7621. doi: 10.1073/pnas.1337546100. Ligand-coated magnetic beads used to show that CD151 affects ligand detachment rather than attachment.
  • 177.Stipp CS, Kolesnikova TV, Hemler ME. EWI-2 is a major CD9 and CD81 partner, and member of a novel Ig protein subfamily. J. Biol. Chem. 2001;276:40545–40554. doi: 10.1074/jbc.M107338200. [DOI] [PubMed] [Google Scholar]
  • 178.Stipp CS, Orlicky D, Hemler ME. FPRP: A major, highly stoichiometric, highly specific CD81 and CD9-associated protein. J. Biol. Chem. 2001;276:4853–4862. doi: 10.1074/jbc.M009859200. [DOI] [PubMed] [Google Scholar]
  • 179.Charrin S, et al. The major CD9 and CD81 molecular partner. Identification and characterization of the complexes. J. Biol. Chem. 2001;276:14329–14337. doi: 10.1074/jbc.M011297200. [DOI] [PubMed] [Google Scholar]
  • 180.Clark KL, Zeng Z, Langford AL, Bowen SM, Todd SC. Pgrl is a major CD81-associated protein on lymphocytes and distinguishes a new family of cell surface proteins. J. Immunol. 2001;167:5115–5121. doi: 10.4049/jimmunol.167.9.5115. [DOI] [PubMed] [Google Scholar]
  • 181.Horvath G, et al. CD19 is linked to the integrin-associated tetraspans CD9, CD81, and CD82. J. Biol. Chem. 1998;273:30537–30543. doi: 10.1074/jbc.273.46.30537. [DOI] [PubMed] [Google Scholar]
  • 182.Yauch RL, Berditchevski F, Harler MB, Reichner J, Hemler ME. Highly stoichiometric, stable and specific association of integrin α3β1 with CD151 provides a major link to phosphatidylinositol 4-kinase and may regulate cell migration. Mol. Biol. Cell. 1998;9:2751–2765. doi: 10.1091/mbc.9.10.2751. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 183.Miyado K, et al. Requirement of CD9 on the egg plasma membrane for fertilization. Science. 2000;287:321–324. doi: 10.1126/science.287.5451.321. [DOI] [PubMed] [Google Scholar]
  • 184.Le Naour F, Rubinstein E, Jasmin C, Prenant M, Boucheix C. Severely reduced female fertility in CD9-deficient mice. Science. 2000;287:319–321. doi: 10.1126/science.287.5451.319. [DOI] [PubMed] [Google Scholar]
  • 185.Kaji K, et al. The gamete fusion process is defective in eggs of Cd9-deficient mice. Nature Genet. 2000;24:279–282. doi: 10.1038/73502. [DOI] [PubMed] [Google Scholar]
  • 186.Kaminski MS, et al. Imaging, dosimetry, and radioimmunotherapy with iodine 131-labeled anti-CD37 antibody in B-cell lymphoma. J. Clin. Oncol. 1992;10:1696–1711. doi: 10.1200/JCO.1992.10.11.1696. [DOI] [PubMed] [Google Scholar]
  • 187.Tanigawa M, et al. Possible involvement of CD81 in acrosome reaction of sperm in mice. Mol. Reprod. Dev. 2008;75:150–155. doi: 10.1002/mrd.20709. [DOI] [PubMed] [Google Scholar]
  • 188.Schmid E, et al. Antibodies to CD9, a tetraspan transmembrane protein, inhibit canine distemper virus-induced cell–cell fusion but not virus–cell fusion. J. Virol. 2000;74:7554–7561. doi: 10.1128/jvi.74.16.7554-7561.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 189.Iwamoto R, Senoh H, Okada Y, Uchida T, Mekada E. An antibody that inhibits the binding of diphtheria toxin to cells revealed the association of a 27-kD membrane protein with the diphtheria toxin receptor. J. Biol. Chem. 1991;266:20463–20469. [PubMed] [Google Scholar]
  • 190.Beatty WL. Trafficking from CD63-positive late endocytic multivesicular bodies is essential for intracellular development of Chlamydia trachomatis. J. Cell Sci. 2006;119:350–359. doi: 10.1242/jcs.02733. [DOI] [PubMed] [Google Scholar]

RESOURCES