CD58

Last updated
CD58 molecule
Identifiers
SymbolCD58
Alt. symbolsLFA3
NCBI gene 965
HGNC 1688
OMIM 153420
RefSeq NM_001779
UniProt P19256
Other data
Locus Chr. 1 p13
Search for
Structures Swiss-model
Domains InterPro

CD58, or lymphocyte function-associated antigen 3 (LFA-3), is a cell adhesion molecule expressed on Antigen Presenting Cells (APCs), particularly macrophages, and other tissue cells. [1] [2] [3]

Contents

CD58 binds to CD2 (LFA-2) [4] [5] on T cells and is important in strengthening the adhesion and recognition between the T cells and Professional Antigen Presenting Cells, facilitating signal transduction necessary for an immune response. This adhesion occurs as part of the transitory initial encounters between T cells and Antigen Presenting Cells before T cell activation, when T cells are roaming the lymph nodes looking at the surface of APCs for peptide:MHC complexes the T-cell receptors are reactive to.

Polymorphisms in the CD58 gene are associated with increased risk for multiple sclerosis. [6] Genomic region containing the single-nucleotide polymorphism rs1335532, associated with high risk of multiple sclerosis, has enhancer properties and can significantly boost the CD58 promoter activity in lymphoblast cells. The protective (C) rs1335532 allele creates functional binding site for ASCL2 transcription factor, a target of the Wnt signaling pathway. [7]

CD58 plays a role in the regulation of colorectal tumor-initiating cells (CT-ICs). Thus, cells that express CD58 have become a cell of interest in tumorigenesis. [8] Mutations of CD58 have been linked to immune evasion observed in some lymphomas and studies are underway to analyze how its involvement directly affects classical Hodgkin lymphoma (cHL). [9]

Introduction

CD58, lymphocyte-function antigen 3 (LFA-3), is a glycoprotein that plays a vital role in the body's immune response. The natural ligand to CD58, CD2, is most commonly found on the surfaces of both T cells and Natural Killer cells (T/NK cells). [3] During an immune response, the interactions between the CD2 and CD58 glycoproteins allows for the activation and proliferation of both T and Natural Killer cells (T/NK cells), enhancing cell adhesion. [3] Furthermore, upon activation, a succession of intracellular signaling within T and Natural Killer cells and other target cells occurs, enhancing further cell recognition. [3] Overall, CD58-CD2 interactions are intricate and involved in a variety of immune regulatory responses, including antiviral, inflammation in numerous autoimmune diseases, and immune rejections in organ transplants. [3]

CD58 is expressed on a variety of different cells, including hematopoietic and nonhematopoietic cells. [10] More specifically, CD58 is expressed on cell surfaces, allowing for effector-target adhesion sequentially to antigen recognition. [11] This adhesion allows for proper T cell activation via correct cell signaling.

CD58 and CD2 interaction

The composition of CD2 and CD58 share many similarities. Both extracellular domains have similar amino acid sequences which aid in cell adhesion. [12] This allows for a high affinity of the extracellular amino-terminal sequence on CD2 to bind with CD58, which has a capacity to bind to CD2 on T cells, on target cells. [12] [13] For a regulatory T cell to become activated, the recognition of an antigen located within a major histocompatibility complex (MHC) protein by the TcR, or T cell receptor, is insufficient. [14] Proliferation of regulatory T cells requires the TcR recognition and other co-stimulatory signals. [15] The binding of CD2-CD58 allows for the formation of a co-stimulatory signal, contributing to further regulatory T cell proliferation and regulation of T cell responses via signaling transduction. [15] [16]

Structure and localization of CD58

The CD58 glycoprotein can be found in two different protein isoforms, each on the cell surface. [17] These include transmembrane and GPI-anchored form. [17] It has been found that in both isoforms, CD58 is able to interact with a variety of different kinases, and is not dependent on only one form. [17] Rather, each isoform is able to associate more effectively with different kinases. [18] Each form, transmembrane and GPI-anchored, can be found in different parts of the cell membrane. The GPI-anchored isoform is mostly found in lipid rafts while the transmembrane isoform is mainly found in nonraft domains. [18] Despite this, the transmembrane CD58 form can trigger independent signaling without the need for the GPI-anchored isoform. [18] Transmembrane CD58 has a structure that consists of six N-linked glycosylation sites in the extracellular domain, a hydrophobic transmembrane domain, and finally a short cytoplasmic domain. [19] GPI-anchored CD58 has a similar extracellular domain, but no hydrophobic transmembrane domain or cytoplasmic domain. [19] Rather, it is linked to the cell membrane via a GPI tail. It is estimated that the CD58 structure is made of approximately 44-68% carbohydrate. [19] The structure of CD58 also plays a role in cell adhesion. A study found that effective cell adhesion was dependent on the density of CD58. [20] Comparing the GPI-anchored and transmembrane isoforms, the GPI-anchored is much more efficient during cell adhesion, and on average, takes much less time than the transmembrane isoform. [20] Regardless, the structure of both the GPI-anchor and transmembrane CD58 are crucial in overall function. While the GPI-anchor enhances cell adhesion, the transmembrane isoform is more efficient in cell signal transduction. [3]

Multiple sclerosis

Multiple Sclerosis (MS) is autoimmune disease that effects the central nervous system (CNS). In an individual with MS, the immune system attacks the myelin sheath, which are crucial for covering nerve fibers and allowing proper communication with the brain and the rest of the body. [21] A genomic association study suggested that there is a risk of developing MS in individuals with allelic variation in the CD58 gene coding region. [22] Further research done on the topic suggested that there is a strong association betweenCD58 single-nucleotide polymorphism (SNP) rs12044852 and the onset of MS. [23] Another study focused on the (SNP) rs1414273 in the microRNA-548ac stem-loop region of the CD58 gene. [24] More specifically, the SNP was found to have an influence on Drosha cleavage activity, which can cause uncoupling of the expression of CD58 and microRNA-548ac production. [24] The data from the study also showed carriers of the allele rs1414273 showed an overall decrease in CD58 mRNA levels. [24] However, the carriers of the allele did exhibit an increase in the levels of hsa-miR-548ac. [24] There is an influence between CD58 and susceptibility to MS. On a similar note, a genome wide association study found that the SNP rs1335532 was associated with a decrease in the susceptibility of developing MS. [25] In addition, it was found that in individuals with MS had an increase in CD58 mRNA. [25] This was because the region where rs1335532 resides had certain properties that increased the activity of CD58 in lymphoblasts. [25] The protective rs1335532 allele also targeted the Wnt signaling pathway by creating a binding site for ASCL2, a transcription factor and target of the Wnt signaling pathway. [25] In immune cells like monocytes and primary B-cells, the Wnt signaling pathway activation causes an increase in CD58 promotor activity via a strong binding site of ASCL2. [25] A reduced expression of CD58 is a possible risk for developing MS.

Rheumatoid arthritis

Rheumatoid arthritis (RA) is an autoimmune disease that mainly affects an individuals' joints, but can affect and cause problems in different tissues. [26] A study that used enzyme-linked immunosorbent assay (ELISA) to measure sCD58 (soluble form of CD58) in individuals with RA and normal controls (NC) to determine if there was a correlation between sCD58 levels and RA. [27] It was found that sCD58 levels were significantly lower in the individuals with RA compared to those in the control (NC). [27] The sCD58 levels in the synovial fluid (SF) of the individuals with RA were also lower than the control subjects. [27] A decrease in sCD58 production could cause a decrease in CD2-CD58 adhesion, leading to an increase in T cells. [28] Continued inflammation would also be an effect of the decrease in sCD58. [28]

Related Research Articles

<span class="mw-page-title-main">CD31</span> Mammalian protein found in Homo sapiens

Platelet endothelial cell adhesion molecule (PECAM-1) also known as cluster of differentiation 31 (CD31) is a protein that in humans is encoded by the PECAM1 gene found on chromosome17q23.3. PECAM-1 plays a key role in removing aged neutrophils from the body.

<span class="mw-page-title-main">ICAM-1</span> Mammalian protein found in Homo sapiens

ICAM-1 also known as CD54 is a protein that in humans is encoded by the ICAM1 gene. This gene encodes a cell surface glycoprotein which is typically expressed on endothelial cells and cells of the immune system. It binds to integrins of type CD11a / CD18, or CD11b / CD18 and is also exploited by rhinovirus as a receptor for entry into respiratory epithelium.

<span class="mw-page-title-main">Immunological synapse</span> Interface between lymphocyte and target cell

In immunology, an immunological synapse is the interface between an antigen-presenting cell or target cell and a lymphocyte such as a T/B cell or Natural Killer cell. The interface was originally named after the neuronal synapse, with which it shares the main structural pattern. An immunological synapse consists of molecules involved in T cell activation, which compose typical patterns—activation clusters. Immunological synapses are the subject of much ongoing research.

<span class="mw-page-title-main">Integrin alpha X</span> Mammalian protein found in Homo sapiens

CD11c, also known as Integrin, alpha X (ITGAX), is a gene that encodes for CD11c.

<span class="mw-page-title-main">Integrin alpha L</span> Mammalian protein found in Homo sapiens

Integrin, alpha L , also known as ITGAL, is a protein that in humans is encoded by the ITGAL gene. CD11a functions in the immune system. It is involved in cellular adhesion and costimulatory signaling. It is the target of the drug efalizumab.

<span class="mw-page-title-main">PTPRC</span> Mammalian protein found in Homo sapiens

Protein tyrosine phosphatase, receptor type, C also known as PTPRC is an enzyme that, in humans, is encoded by the PTPRC gene. PTPRC is also known as CD45 antigen, which was originally called leukocyte common antigen (LCA).

<span class="mw-page-title-main">CD2</span> Cell adhesion molecule found on the surface of T cells and natural killer

CD2 is a cell adhesion molecule found on the surface of T cells and natural killer (NK) cells. It has also been called T-cell surface antigen T11/Leu-5, LFA-2, LFA-3 receptor, erythrocyte receptor and rosette receptor.

Lymphocyte function-associated antigen 1 (LFA-1) is an integrin found on lymphocytes and other leukocytes. LFA-1 plays a key role in emigration, which is the process by which leukocytes leave the bloodstream to enter the tissues. LFA-1 also mediates firm arrest of leukocytes. Additionally, LFA-1 is involved in the process of cytotoxic T cell mediated killing as well as antibody mediated killing by granulocytes and monocytes. As of 2007, LFA-1 has 6 known ligands: ICAM-1, ICAM-2, ICAM-3, ICAM-4, ICAM-5, and JAM-A. LFA-1/ICAM-1 interactions have recently been shown to stimulate signaling pathways that influence T cell differentiation. LFA-1 belongs to the integrin superfamily of adhesion molecules.

<span class="mw-page-title-main">CD22</span> Lectin molecule

CD22, or cluster of differentiation-22, is a molecule belonging to the SIGLEC family of lectins. It is found on the surface of mature B cells and to a lesser extent on some immature B cells. Generally speaking, CD22 is a regulatory molecule that prevents the overactivation of the immune system and the development of autoimmune diseases.

<span class="mw-page-title-main">CD146</span> Protein-coding gene in the species Homo sapiens

CD146 also known as the melanoma cell adhesion molecule (MCAM) or cell surface glycoprotein MUC18, is a 113kDa cell adhesion molecule currently used as a marker for endothelial cell lineage. In humans, the CD146 protein is encoded by the MCAM gene.

<span class="mw-page-title-main">CD52</span> Mammalian protein found in Homo sapiens

CAMPATH-1 antigen, also known as cluster of differentiation 52 (CD52), is a glycoprotein that in humans is encoded by the CD52 gene.

<span class="mw-page-title-main">Integrin beta 2</span> Mammalian protein found in Homo sapiens

In molecular biology, CD18 is an integrin beta chain protein that is encoded by the ITGB2 gene in humans. Upon binding with one of a number of alpha chains, CD18 is capable of forming multiple heterodimers, which play significant roles in cellular adhesion and cell surface signaling, as well as important roles in immune responses. CD18 also exists in soluble, ligand binding forms. Deficiencies in CD18 expression can lead to adhesion defects in circulating white blood cells in humans, reducing the immune system's ability to fight off foreign invaders.

<span class="mw-page-title-main">ICAM3</span> Mammalian protein found in Homo sapiens

Intercellular adhesion molecule 3 (ICAM3) also known as CD50, is a protein that in humans is encoded by the ICAM3 gene. The protein is constitutively expressed on the surface of leukocytes, which are also called white blood cells and are part of the immune system. ICAM3 mediates adhesion between cells by binding to specific integrin receptors. It plays an important role in the immune cell response through its facilitation of interactions between T cells and dendritic cells, which allows for T cell activation. ICAM3 also mediates the clearance of cells undergoing apoptosis by attracting and binding macrophages, a type of cell that breaks down infected or dying cells through a process known as phagocytosis, to apoptotic cells.

<span class="mw-page-title-main">ICAM2</span> Protein-coding gene in the species Homo sapiens

Intercellular adhesion molecule 2 (ICAM2), also known as CD102, is a human gene, and the protein resulting from it.

<span class="mw-page-title-main">CD69</span>

CD69 is a human transmembrane C-Type lectin protein encoded by the CD69 gene. It is an early activation marker that is expressed in hematopoietic stem cells, T cells, and many other cell types in the immune system. It is also implicated in T cell differentiation as well as lymphocyte retention in lymphoid organs.

<span class="mw-page-title-main">CD97</span> Mammalian protein found in Homo sapiens

Cluster of differentiation 97 is a protein also known as BL-Ac[F2] encoded by the ADGRE5 gene. CD97 is a member of the adhesion G protein-coupled receptor (GPCR) family. Adhesion GPCRs are characterized by an extended extracellular region often possessing N-terminal protein modules that is linked to a TM7 region via a domain known as the GPCR-Autoproteolysis INducing (GAIN) domain.

<i>CD82</i> (gene) Mammalian protein found in Homo sapiens

CD82, or KAI1, is a human protein encoded by the CD82 gene.

<span class="mw-page-title-main">CD48</span> Protein-coding gene in humans

CD48 antigen also known as B-lymphocyte activation marker (BLAST-1) or signaling lymphocytic activation molecule 2 (SLAMF2) is a protein that in humans is encoded by the CD48 gene.

<span class="mw-page-title-main">CD226</span> Protein-coding gene in the species Homo sapiens

CD226, PTA1 or DNAM-1 is a ~65 kDa immunoglobulin-like transmembrane glycoprotein expressed on the surface of natural killer cells, NK T cell, B cells, dendritic cells, hematopoietic precursor cells, platelets, monocytes and T cells.

<span class="mw-page-title-main">CD6</span>

CD6 is a human protein encoded by the CD6 gene.

References

  1. Barbosa JA, Mentzer SJ, Kamarck ME, Hart J, Biro PA, Strominger JL, Burakoff SJ (April 1986). "Gene mapping and somatic cell hybrid analysis of the role of human lymphocyte function-associated antigen-3 (LFA-3) in CTL-target cell interactions". Journal of Immunology. 136 (8): 3085–91. doi:10.4049/jimmunol.136.8.3085. PMID   3514752. S2CID   22866345.
  2. Wallich R, Brenner C, Brand Y, Roux M, Reister M, Meuer S (March 1998). "Gene structure, promoter characterization, and basis for alternative mRNA splicing of the human CD58 gene". Journal of Immunology. 160 (6): 2862–71. doi:10.4049/jimmunol.160.6.2862. PMID   9510189. S2CID   8489739.
  3. 1 2 3 4 5 6 Zhang, Yalu; Liu, Qiaofei; Yang, Sen; Liao, Quan (2021-06-08). "CD58 Immunobiology at a Glance". Frontiers in Immunology. 12: 705260. doi: 10.3389/fimmu.2021.705260 . ISSN   1664-3224. PMC   8218816 . PMID   34168659.
  4. Selvaraj P, Plunkett ML, Dustin M, Sanders ME, Shaw S, Springer TA (1987). "The T lymphocyte glycoprotein CD2 binds the cell surface ligand LFA-3". Nature. 326 (6111): 400–3. Bibcode:1987Natur.326..400S. doi:10.1038/326400a0. PMID   2951597. S2CID   4334290.
  5. Wang JH, Smolyar A, Tan K, Liu JH, Kim M, Sun ZY, et al. (June 1999). "Structure of a heterophilic adhesion complex between the human CD2 and CD58 (LFA-3) counterreceptors". Cell. 97 (6): 791–803. doi: 10.1016/S0092-8674(00)80790-4 . PMID   10380930. S2CID   6264220.
  6. De Jager PL, Baecher-Allan C, Maier LM, Arthur AT, Ottoboni L, Barcellos L, et al. (March 2009). "The role of the CD58 locus in multiple sclerosis". Proceedings of the National Academy of Sciences of the United States of America. 106 (13): 5264–9. Bibcode:2009PNAS..106.5264D. doi: 10.1073/pnas.0813310106 . PMC   2664005 . PMID   19237575.
  7. Mitkin NA, Muratova AM, Korneev KV, Pavshintsev VV, Rumyantsev KA, Vagida MS, et al. (October 2018). "Protective C allele of the single-nucleotide polymorphism rs1335532 is associated with strong binding of Ascl2 transcription factor and elevated CD58 expression in B-cells". Biochimica et Biophysica Acta (BBA) - Molecular Basis of Disease. 1864 (10): 3211–3220. doi: 10.1016/j.bbadis.2018.07.008 . PMID   30006149.
  8. Xu S, Wen Z, Jiang Q, Zhu L, Feng S, Zhao Y, et al. (March 2015). "CD58, a novel surface marker, promotes self-renewal of tumor-initiating cells in colorectal cancer". Oncogene. 34 (12): 1520–31. doi: 10.1038/onc.2014.95 . PMID   24727892. S2CID   22035160.
  9. Schneider M, Schneider S, Zühlke-Jenisch R, Klapper W, Sundström C, Hartmann S, et al. (October 2015). "Alterations of the CD58 gene in classical Hodgkin lymphoma". Genes, Chromosomes & Cancer. 54 (10): 638–45. doi:10.1002/gcc.22276. PMID   26194173. S2CID   28291445.
  10. Krensky, A. M.; Sanchez-Madrid, F.; Robbins, E.; Nagy, J. A.; Springer, T. A.; Burakoff, S. J. (August 1983). "The functional significance, distribution, and structure of LFA-1, LFA-2, and LFA-3: cell surface antigens associated with CTL-target interactions". Journal of Immunology. 131 (2): 611–616. doi:10.4049/jimmunol.131.2.611. ISSN   0022-1767. PMID   6345670. S2CID   24722957.
  11. Krensky, A. M.; Robbins, E.; Springer, T. A.; Burakoff, S. J. (May 1984). "LFA-1, LFA-2, and LFA-3 antigens are involved in CTL-target conjugation". Journal of Immunology. 132 (5): 2180–2182. doi:10.4049/jimmunol.132.5.2180. ISSN   0022-1767. PMID   6201533. S2CID   11856034.
  12. 1 2 Sewell, W. A.; Palmer, R. W.; Spurr, N. K.; Sheer, D.; Brown, M. H.; Bell, Y.; Crumpton, M. J. (1988). "The human LFA-3 gene is located at the same chromosome band as the gene for its receptor CD2". Immunogenetics. 28 (4): 278–282. doi:10.1007/BF00345506. ISSN   0093-7711. PMID   2458315. S2CID   11197336.
  13. Selvaraj, P.; Plunkett, M. L.; Dustin, M.; Sanders, M. E.; Shaw, S.; Springer, T. A. (March 26 – April 1, 1987). "The T lymphocyte glycoprotein CD2 binds the cell surface ligand LFA-3". Nature. 326 (6111): 400–403. Bibcode:1987Natur.326..400S. doi:10.1038/326400a0. ISSN   0028-0836. PMID   2951597. S2CID   4334290.
  14. Bierer, B. E.; Hahn, W. C. (August 1993). "T cell adhesion, avidity regulation and signaling: a molecular analysis of CD2". Seminars in Immunology. 5 (4): 249–261. doi:10.1006/smim.1993.1029. ISSN   1044-5323. PMID   7693022.
  15. 1 2 Van Seventer, G. A.; Shimizu, Y.; Horgan, K. J.; Luce, G. E.; Webb, D.; Shaw, S. (July 1991). "Remote T cell co-stimulation via LFA-1/ICAM-1 and CD2/LFA-3: demonstration with immobilized ligand/mAb and implication in monocyte-mediated co-stimulation". European Journal of Immunology. 21 (7): 1711–1718. doi:10.1002/eji.1830210719. ISSN   0014-2980. PMID   1711977. S2CID   352060.
  16. Miller, G. T.; Hochman, P. S.; Meier, W.; Tizard, R.; Bixler, S. A.; Rosa, M. D.; Wallner, B. P. (1993-07-01). "Specific interaction of lymphocyte function-associated antigen 3 with CD2 can inhibit T cell responses". The Journal of Experimental Medicine. 178 (1): 211–222. doi:10.1084/jem.178.1.211. ISSN   0022-1007. PMC   2191085 . PMID   7686212.
  17. 1 2 3 Itzhaky, D.; Raz, N.; Hollander, N. (1998-05-01). "The glycosylphosphatidylinositol-anchored form and the transmembrane form of CD58 associate with protein kinases". Journal of Immunology. 160 (9): 4361–4366. doi:10.4049/jimmunol.160.9.4361. ISSN   0022-1767. PMID   9574540. S2CID   16730539.
  18. 1 2 3 Ariel, Ortal; Kukulansky, Tova; Raz, Nava; Hollander, Nurit (June 2004). "Distinct membrane localization and kinase association of the two isoforms of CD58". Cellular Signalling. 16 (6): 667–673. doi:10.1016/j.cellsig.2003.08.015. ISSN   0898-6568. PMID   15093607.
  19. 1 2 3 Wallner, B. P.; Frey, A. Z.; Tizard, R.; Mattaliano, R. J.; Hession, C.; Sanders, M. E.; Dustin, M. L.; Springer, T. A. (1987-10-01). "Primary structure of lymphocyte function-associated antigen 3 (LFA-3). The ligand of the T lymphocyte CD2 glycoprotein". The Journal of Experimental Medicine. 166 (4): 923–932. doi:10.1084/jem.166.4.923. ISSN   0022-1007. PMC   2188720 . PMID   3309127.
  20. 1 2 Chan, P. Y.; Lawrence, M. B.; Dustin, M. L.; Ferguson, L. M.; Golan, D. E.; Springer, T. A. (October 1991). "Influence of receptor lateral mobility on adhesion strengthening between membranes containing LFA-3 and CD2". The Journal of Cell Biology. 115 (1): 245–255. doi:10.1083/jcb.115.1.245. ISSN   0021-9525. PMC   2289925 . PMID   1717480.
  21. "Multiple sclerosis - Symptoms and causes". Mayo Clinic. Retrieved 2023-04-18.
  22. De Jager, Philip L.; Baecher-Allan, Clare; Maier, Lisa M.; Arthur, Ariel T.; Ottoboni, Linda; Barcellos, Lisa; McCauley, Jacob L.; Sawcer, Stephen; Goris, An; Saarela, Janna; Yelensky, Roman; Price, Alkes; Leppa, Virpi; Patterson, Nick; de Bakker, Paul I. W. (2009-03-31). "The role of the CD58 locus in multiple sclerosis". Proceedings of the National Academy of Sciences of the United States of America. 106 (13): 5264–5269. Bibcode:2009PNAS..106.5264D. doi: 10.1073/pnas.0813310106 . ISSN   1091-6490. PMC   2664005 . PMID   19237575.
  23. Torbati, Sara; Karami, Fatemeh; Ghaffarpour, Majid; Zamani, Mahdi (2015). "Association of CD58 Polymorphism with Multiple Sclerosis and Response to Interferon ß Therapy in A Subset of Iranian Population". Cell Journal. 16 (4): 506–513. doi:10.22074/cellj.2015.505. ISSN   2228-5806. PMC   4297489 . PMID   25685741.
  24. 1 2 3 4 Hecker, Michael; Boxberger, Nina; Illner, Nicole; Fitzner, Brit; Schröder, Ina; Winkelmann, Alexander; Dudesek, Ales; Meister, Stefanie; Koczan, Dirk; Lorenz, Peter; Thiesen, Hans-Jürgen; Zettl, Uwe Klaus (February 2019). "A genetic variant associated with multiple sclerosis inversely affects the expression of CD58 and microRNA-548ac from the same gene". PLOS Genetics. 15 (2): e1007961. doi:10.1371/journal.pgen.1007961. ISSN   1553-7404. PMC   6382214 . PMID   30730892.
  25. 1 2 3 4 5 Mitkin, Nikita A.; Muratova, Alisa M.; Korneev, Kirill V.; Pavshintsev, Vsevolod V.; Rumyantsev, Konstantin A.; Vagida, Murad S.; Uvarova, Aksinya N.; Afanasyeva, Marina A.; Schwartz, Anton M.; Kuprash, Dmitry V. (October 2018). "Protective C allele of the single-nucleotide polymorphism rs1335532 is associated with strong binding of Ascl2 transcription factor and elevated CD58 expression in B-cells". Biochimica et Biophysica Acta (BBA) - Molecular Basis of Disease. 1864 (10): 3211–3220. doi:10.1016/j.bbadis.2018.07.008. ISSN   1879-260X. PMID   30006149. S2CID   51625452.
  26. "Rheumatoid Arthritis (RA) | Arthritis | CDC". www.cdc.gov. 2020-07-27. Retrieved 2023-04-19.
  27. 1 2 3 Hoffmann, J. C.; Räuker, H. J.; Krüger, H.; Bayer, B.; Zeidler, H. (1996). "Decreased levels of a soluble form of the human adhesion receptor CD58 (LFA-3) in sera and synovial fluids of patients with rheumatoid arthritis". Clinical and Experimental Rheumatology. 14 (1): 23–29. ISSN   0392-856X. PMID   8697653.
  28. 1 2 Hoffmann, J. C.; Bayer, B.; Zeidler, H. (June 1996). "Characterization of a soluble form of CD58 in synovial fluid of patients with rheumatoid arthritis (RA)". Clinical and Experimental Immunology. 104 (3): 460–466. doi:10.1046/j.1365-2249.1996.41749.x. ISSN   0009-9104. PMC   2200442 . PMID   9099931.