Pattern recognition receptor: Difference between revisions

From Wikipedia, the free encyclopedia
Content deleted Content added
No edit summary
I rewrote the CLR paragraph, and classification based on citations and information available to me, added citations, were they were missing or inadequate, added more information to the text overall and wrote the paragraph use in human medicine. Stylistics
Line 1: Line 1:
'''Pattern recognition receptors''' ('''PRRs''')<ref>{{cite journal|last1=Mahla|first1=R.S.|title=Sweeten PAMPs: Role of Sugar Complexed PAMPs in Innate Immunity and Vaccine Biology|journal=Front Immunol|doi=10.3389/fimmu.2013.00248|pmid=24032031|volume=4|pmc=3759294|year=2013|pages=248}}</ref> are a part of the [[innate immune system]]. They are [[protein]]s expressed by cells of the [[innate immune system]] to identify two classes of molecules: [[pathogen-associated molecular patterns]] (PAMPs), which are associated with microbial [[pathogens]], and [[damage-associated molecular patterns]] (DAMPs), which are associated with cell components that are released during cell damage or death. They are also called '''primitive pattern recognition receptors''' because they evolved before other parts of the immune system, particularly before [[adaptive immunity]].
'''Pattern recognition receptors''' ('''PRRs''')<ref>{{cite journal|last1=Mahla|first1=R.S.|title=Sweeten PAMPs: Role of Sugar Complexed PAMPs in Innate Immunity and Vaccine Biology|journal=Front Immunol|doi=10.3389/fimmu.2013.00248|pmid=24032031|volume=4|pmc=3759294|year=2013|pages=248}}</ref> play a crucial role in the proper function of the [[innate immune system]]. PRRs are germline-encoded host sensors, which detect molecules typical for the pathogens<ref name=":0">{{Cite journal|last=Kumar|first=Himanshu|last2=Kawai|first2=Taro|last3=Akira|first3=Shizuo|date=2011-01-01|title=Pathogen Recognition by the Innate Immune System|url=http://dx.doi.org/10.3109/08830185.2010.529976|journal=International Reviews of Immunology|volume=30|issue=1|pages=16–34|doi=10.3109/08830185.2010.529976|issn=0883-0185|pmid=21235323}}</ref>. They are [[protein]]s expressed by cells of the [[innate immune system]], such as dendritic cells, macrophages, monocytes, neutrophils and epithelial cells<ref>{{Cite journal|last=Alberts|first=Bruce|last2=Johnson|first2=Alexander|last3=Lewis|first3=Julian|last4=Raff|first4=Martin|last5=Roberts|first5=Keith|last6=Walter|first6=Peter|date=2002|title=Innate Immunity|url=https://www.ncbi.nlm.nih.gov/books/NBK26846/|language=en}}</ref><ref>{{Cite journal|last=Schroder|first=Kate|last2=Tschopp|first2=Jurg|title=The Inflammasomes|url=https://doi.org/10.1016/j.cell.2010.01.040|journal=Cell|volume=140|issue=6|pages=821–832|doi=10.1016/j.cell.2010.01.040}}</ref>, to identify two classes of molecules: [[pathogen-associated molecular patterns]] (PAMPs), which are associated with microbial [[pathogens]], and [[damage-associated molecular patterns]] (DAMPs), which are associated with components of host's cells that are released during cell damage or death. They are also called '''primitive pattern recognition receptors''' because they evolved before other parts of the immune system, particularly before [[adaptive immunity]]. PRRs also mediate the initiation of antigen-specific adaptive immune response and release of inflammatory cytokines<ref name=":0" /><ref name=":1">{{Cite web|url=http://www.annualreviews.org/doi/10.1146/annurev.immunol.21.120601.141126|title=Toll-Like Receptors|last=Takeda|first=Kiyoshi|last2=Kaisho|first2=Tsuneyasu|date=2003-11-28|website=https://doi.org/10.1146/annurev.immunol.21.120601.141126|language=en|doi=10.1146/annurev.immunol.21.120601.141126|access-date=2017-07-27|last3=Akira|first3=Shizuo}}</ref>


==Molecules recognized==
==Molecules recognized==
The microbe-specific molecules that are recognized by a given PRR are called [[pathogen-associated molecular pattern]]s (PAMPs) and include bacterial carbohydrates (such as [[lipopolysaccharide]] or LPS, [[mannose]]), nucleic acids (such as bacterial or viral DNA or RNA), bacterial peptides (flagellin, microtubule elongation factors), [[peptidoglycan]]s and [[lipoteichoic acid]]s (from Gram-positive bacteria), [[N-Formylmethionine|''N''-formylmethionine]], [[lipoprotein]]s and fungal [[glucan]]s and chitin.
The microbe-specific molecules that are recognized by a given PRR are called [[pathogen-associated molecular pattern]]s (PAMPs) and include bacterial carbohydrates (such as [[lipopolysaccharide]] or LPS, [[mannose]]), nucleic acids (such as bacterial or viral DNA or RNA), bacterial peptides (flagellin, microtubule elongation factors), [[peptidoglycan]]s and [[lipoteichoic acid]]s (from Gram-positive bacteria), [[N-Formylmethionine|''N''-formylmethionine]], [[lipoprotein]]s and fungal [[glucan]]s and chitin.


Endogenous stress signals are called [[damage-associated molecular pattern]]s (DAMPs) and include [[uric acid]] and extracellular [[adenosine triphosphate|ATP]], among many other compounds.
Endogenous stress signals are called [[damage-associated molecular pattern]]s (DAMPs) and include [[uric acid]] and extracellular [[adenosine triphosphate|ATP]], among many other compounds<ref name=":0" />.


==Classification==
==Classification==
PRRs are classified according to their [[ligand]] specificity, function, localization and/or evolutionary relationships. On the basis of function, PRRs may be divided into endocytic PRRs and signaling PRRs.
There are several subgroups of PRRs. They are classified according to their [[ligand]] specificity, function, localization and/or evolutionary relationships. Based on their localization, PRRs may be divided into membrane-bound PRRs and cytoplasmic PRRs.


* ''Signaling PRRs'' include the large families of membrane-bound [[Toll-like receptor]]s and cytoplasmic [[NOD-like receptor]]s.
* ''Membrane-bound PRRs'' include [[Toll like receptors (TLRs)]] and [[C-type lectin receptors (CLRs)]].
* ''Cytoplasmic PRRs'' include [[NOD-like receptors (NLRs)]] and [[RIG-I-like receptors (RLRs)]].
* ''Endocytic PRRs'' promote the attachment, engulfment and destruction of microorganisms by phagocytes, without relaying an intracellular signal. These PRRs recognize carbohydrates and include mannose receptors of [[macrophage]]s, [[glucan]] receptors present on all [[phagocytes]] and [[scavenger receptor (immunology)|scavenger receptor]]s that recognize charged ligands, are found on all phagocytes and mediate removal of apoptotic cells.


==PRR types and localization==
==PRR types and localization==
Line 19: Line 19:
PRRs were first discovered in plants.<ref name="pmid8525370">{{cite journal |vauthors=Song WY, Wang GL, Chen LL, Kim HS, Pi LY, Holsten T, Gardner J, Wang B, Zhai WX, Zhu LH, Fauquet C, Ronald P | title = A receptor kinase-like protein encoded by the rice disease resistance gene, Xa21 | journal = Science | volume = 270 | issue = 5243 | pages = 1804–6 |date=December 1995 | pmid = 8525370 | doi = 10.1126/science.270.5243.1804| url = | issn = |bibcode = 1995Sci...270.1804S }}</ref> Since that time many plant PRRs have been predicted by genomic analysis (370 in rice; 47 in Arabidopsis). Unlike animal PRRs, which associated with intracellular kinases via adaptor proteins (see non-RD kinases below), plant PRRs are composed of an extracellular domain, transmembrane domain, juxtamembrane domain and intracellular kinase domain as part of a single protein.
PRRs were first discovered in plants.<ref name="pmid8525370">{{cite journal |vauthors=Song WY, Wang GL, Chen LL, Kim HS, Pi LY, Holsten T, Gardner J, Wang B, Zhai WX, Zhu LH, Fauquet C, Ronald P | title = A receptor kinase-like protein encoded by the rice disease resistance gene, Xa21 | journal = Science | volume = 270 | issue = 5243 | pages = 1804–6 |date=December 1995 | pmid = 8525370 | doi = 10.1126/science.270.5243.1804| url = | issn = |bibcode = 1995Sci...270.1804S }}</ref> Since that time many plant PRRs have been predicted by genomic analysis (370 in rice; 47 in Arabidopsis). Unlike animal PRRs, which associated with intracellular kinases via adaptor proteins (see non-RD kinases below), plant PRRs are composed of an extracellular domain, transmembrane domain, juxtamembrane domain and intracellular kinase domain as part of a single protein.


====Toll-like receptors (TLR)====
===Toll-like receptors (TLR)===
Recognition of extracellular or endosomal pathogen-associated molecular patterns is mediated by transmembrane proteins known as [[toll-like receptor]]s (TLRs).<ref name="pmid16551253">{{cite journal |vauthors=Beutler B, Jiang Z, Georgel P, Crozat K, Croker B, Rutschmann S, Du X, Hoebe K | title = Genetic analysis of host resistance: Toll-like receptor signaling and immunity at large | journal = Annu. Rev. Immunol. | volume = 24 | issue = | pages = 353–389 | year = 2006 | pmid = 16551253 | doi = 10.1146/annurev.immunol.24.021605.090552 | url = | issn = }}</ref> Toll-like receptors were first discovered in ''[[Drosophila melanogaster|Drosophila]]'' and trigger the synthesis and secretion of [[cytokine]]s and activation of other host defense programs that are necessary for innate or adaptive immune responses. TLRs have been found in many species. In mammals, these receptors have been assigned numbers 1 to 11 (TLR1-TLR11). Interaction of TLRs with their specific PAMP induces [[NF-κB]] signaling and the [[Mitogen-activated protein kinase|MAP kinase]] pathway and therefore the secretion of pro-inflammatory [[cytokine]]s and [[co-stimulation|co-stimulatory molecule]]s. Molecules released following TLR activation signal to other cells of the immune system making TLRs key elements of [[innate immunity]] and [[adaptive immunity]].<ref name="pmid16930560">{{cite journal |vauthors=Doyle SL, O'Neill LA | title = Toll-like receptors: from the discovery of NFkappaB to new insights into transcriptional regulations in innate immunity | journal = Biochem. Pharmacol. | volume = 72 | issue = 9 | pages = 1102–1113 |date=October 2006 | pmid = 16930560 | doi = 10.1016/j.bcp.2006.07.010 | url = | issn = }}</ref>
Recognition of extracellular or endosomal pathogen-associated molecular patterns is mediated by transmembrane proteins known as [[toll-like receptor]]s (TLRs).<ref name="pmid16551253">{{cite journal|year=2006|title=Genetic analysis of host resistance: Toll-like receptor signaling and immunity at large|url=|journal=Annu. Rev. Immunol.|volume=24|issue=|pages=353–389|doi=10.1146/annurev.immunol.24.021605.090552|issn=|pmid=16551253|vauthors=Beutler B, Jiang Z, Georgel P, Crozat K, Croker B, Rutschmann S, Du X, Hoebe K}}</ref> TLRs share a typical structural motif, the [[leucine rich repeats (LRR)]], which give them their specific appearance and are also responsible for TLR functionality<ref>{{Cite journal|last=Botos|first=Istvan|last2=Segal|first2=David M.|last3=Davies|first3=David R.|title=The Structural Biology of Toll-like Receptors|url=http://linkinghub.elsevier.com/retrieve/pii/S0969212611000724|journal=Structure|volume=19|issue=4|pages=447–459|doi=10.1016/j.str.2011.02.004|pmc=PMC3075535|pmid=21481769}}</ref>. Toll-like receptors were first discovered in ''[[Drosophila melanogaster|Drosophila]]'' and trigger the synthesis and secretion of [[cytokine]]s and activation of other host defense programs that are necessary for both innate or adaptive immune responses. 10 functional members of the TLR family have been described in humans so far<ref>{{Cite web|url=http://www.annualreviews.org/doi/10.1146/annurev.immunol.21.120601.141126|title=Toll-Like Receptors|last=Takeda|first=Kiyoshi|last2=Kaisho|first2=Tsuneyasu|date=2003-11-28|website=https://doi.org/10.1146/annurev.immunol.21.120601.141126|language=en|doi=10.1146/annurev.immunol.21.120601.141126|access-date=2017-07-27|last3=Akira|first3=Shizuo}}</ref>. Studies have been conducted on [[TLR11]] as well, and it has been shown that it recognizes flagellin and profilin-like proteins in mice<ref>{{Cite journal|last=Hatai|first=Hirotsugu|last2=Lepelley|first2=Alice|last3=Zeng|first3=Wangyong|last4=Hayden|first4=Matthew S.|last5=Ghosh|first5=Sankar|date=2016-02-09|title=Toll-Like Receptor 11 (TLR11) Interacts with Flagellin and Profilin through Disparate Mechanisms|url=http://journals.plos.org/plosone/article?id=10.1371/journal.pone.0148987|journal=PLOS ONE|volume=11|issue=2|pages=e0148987|doi=10.1371/journal.pone.0148987|issn=1932-6203|pmc=PMC4747465|pmid=26859749}}</ref>. Nonetheless, TLR11 is only a [[pseudogene]] in humans without direct function or functional protein expression. Each of the TLR has been shown to interact with a specific PAMP<ref name=":1" /><ref name=":2">{{Cite journal|last=Ozinsky|first=Adrian|last2=Underhill|first2=David M.|last3=Fontenot|first3=Jason D.|last4=Hajjar|first4=Adeline M.|last5=Smith|first5=Kelly D.|last6=Wilson|first6=Christopher B.|last7=Schroeder|first7=Lea|last8=Aderem|first8=Alan|date=2000-12-05|title=The repertoire for pattern recognition of pathogens by the innate immune system is defined by cooperation between Toll-like receptors|url=http://www.pnas.org/content/97/25/13766|journal=Proceedings of the National Academy of Sciences|language=en|volume=97|issue=25|pages=13766–13771|doi=10.1073/pnas.250476497|issn=0027-8424|pmc=PMC17650|pmid=11095740}}</ref><ref>{{Cite journal|last=Lien|first=E.|last2=Sellati|first2=T. J.|last3=Yoshimura|first3=A.|last4=Flo|first4=T. H.|last5=Rawadi|first5=G.|last6=Finberg|first6=R. W.|last7=Carroll|first7=J. D.|last8=Espevik|first8=T.|last9=Ingalls|first9=R. R.|date=1999-11-19|title=Toll-like receptor 2 functions as a pattern recognition receptor for diverse bacterial products|url=https://www.ncbi.nlm.nih.gov/pubmed/10559223|journal=The Journal of Biological Chemistry|volume=274|issue=47|pages=33419–33425|issn=0021-9258|pmid=10559223}}</ref>.


====C-type lectin receptors (CLR)====
===== The TLR signaling mechanism =====
TLRs tend to dimerize, [[TLR4]] forms [[homodimers]], and [[TLR6]] can dimerize with either [[TLR 1|TLR1]] or [[TLR2]]. <ref name=":2" /> Interaction of TLRs with their specific PAMP is mediated through either [[MYD88|MyD88]]- dependent pathway and triggers the signaling through [[NF-κB]] and the [[Mitogen-activated protein kinase|MAP kinase]] pathway and therefore the secretion of pro-inflammatory [[cytokine]]s and co-stimulatory molecules or [[TRIF]] - dependent signaling pathway<ref name=":0" /><ref name=":1" /><ref name=":2" />. MyD88 - dependent pathway is induced by various PAMPs stimulating the TLRs on macrophages and dendritic cells. MyD88 attracts the IRAK4 molecule, IRAK4 recruits IRAK1 and IRAK2 to form a signaling complex. The signaling complex reacts with TRAF6 which leads to TAK1 activation and consequently the induction of inflammatory cytokines.The TRIF-dependent pathway is induced by macrophages and DCs after [[TLR3]] and TLR4 stimulation.<ref name=":0" /> Molecules released following TLR activation signal to other cells of the immune system making TLRs key elements of [[innate immunity]] and [[adaptive immunity]].<ref name=":0" /><ref>{{Cite journal|last=Akira|first=Shizuo|last2=Takeda|first2=Kiyoshi|title=Toll-like receptor signalling|url=http://www.nature.com/doifinder/10.1038/nri1391|journal=Nature Reviews Immunology|volume=4|issue=7|pages=499–511|doi=10.1038/nri1391}}</ref><ref name="pmid16930560">{{cite journal |vauthors=Doyle SL, O'Neill LA | title = Toll-like receptors: from the discovery of NFkappaB to new insights into transcriptional regulations in innate immunity | journal = Biochem. Pharmacol. | volume = 72 | issue = 9 | pages = 1102–1113 |date=October 2006 | pmid = 16930560 | doi = 10.1016/j.bcp.2006.07.010 | url = | issn = }}</ref>
Many different cells involved in the innate immune system, starting from cells as complex as immature dendritic cells to cells as small as platelets, express a myriad of CLRs which shape innate immunity by virtue of their pattern recognition ability.<ref>Nat Rev Immunol. 2009 Jul;9(7):465-79. Signalling through C-type lectin receptors: shaping immune responses. Geijtenbeek TB, Gringhuis SI. http://www.mh-hannover.de/fileadmin/mhh/bilder/international/hbrs_mdphd/ZIB/Vorlesungen/Paper_09-10/Rev_IM_Geijtenbeek.pdf</ref> Most classes of human pathogens are covered by CLRs: e.g. among external or infectious pathogens mannose is the recognition motif for many viruses, fungi and mycobacteria; similarly fucose for certain bacteria and helminths; and glucans are present on mycobacteria and fungi. In addition, many of acquired nonself surfaces e.g. carcinoembryonic/oncofetal type neoantigens carrying "internal danger source"/"self turned nonself" type pathogen pattern are also identified and destroyed (e.g. by complement fixation or other cytotoxic attacks) or sequestered (phagocytosed or ensheathed) by the immune system by virtue of the CLRs. The name lectin is also a bit misleading because the family includes proteins with at least one C-type lectin domain (CTLD) which is a specific type of carbohydrate recognition domain. CTLD is a ligand binding motif found in more than 1000 known proteins (more than 100 in humans) and the ligands are often not sugars.<ref>Cummings RD, McEver RP. C-type Lectins. In: Varki A, Cummings RD, Esko JD, et al., editors. Essentials of Glycobiology. 2nd edition. Cold Spring Harbor (NY): Cold Spring Harbor Laboratory Press; 2009. http://www.ncbi.nlm.nih.gov/books/NBK1943/</ref> If and when the ligand is sugar they need Ca2+ – hence the name "C-type", but many of them don't even have a known sugar ligand thus despite carrying a lectin type fold structure, some of them are technically not "lectin" in function.


===C-type lectin receptors (CLR)===
CLRs are often endocytic receptors but may be also signaling as well, and depending on the ligand their signaling and trafficking fate might be different. They may have direct signaling including [[MAP kinase]] and [[NFkB]] activation or may act indirectly by modulating TLRs. CLRs are a very big category and includes soluble PRRs as well (see at the [[Pattern recognition receptor#Secreted PRRs|end]] of this article) that serve many important immune functions.
Many different cells of the innate immune system express a myriad of CLRs which shape innate immunity by virtue of their pattern recognition ability.<ref>Nat Rev Immunol. 2009 Jul;9(7):465-79. Signalling through C-type lectin receptors: shaping immune responses. Geijtenbeek TB, Gringhuis SI. http://www.mh-hannover.de/fileadmin/mhh/bilder/international/hbrs_mdphd/ZIB/Vorlesungen/Paper_09-10/Rev_IM_Geijtenbeek.pdf</ref> Even though, most classes of human pathogens are covered by CLRs, CLRs are a major receptor for recognition of funghi<ref name=":3">{{Cite journal|last=Hoving|first=J. Claire|last2=Wilson|first2=Gillian J.|last3=Brown|first3=Gordon D.|date=2014-02-01|title=Signalling C-Type lectin receptors, microbial recognition and immunity|url=http://onlinelibrary.wiley.com/doi/10.1111/cmi.12249/abstract|journal=Cellular Microbiology|language=en|volume=16|issue=2|pages=185–194|doi=10.1111/cmi.12249|issn=1462-5822|pmc=PMC4016756|pmid=24330199}}</ref><ref>{{Cite journal|last=Hardison|first=Sarah E|last2=Brown|first2=Gordon D|title=C-type lectin receptors orchestrate antifungal immunity|url=http://www.nature.com/doifinder/10.1038/ni.2369|journal=Nature Immunology|volume=13|issue=9|pages=817–822|doi=10.1038/ni.2369|pmc=PMC3432564|pmid=22910394}}</ref>: nonetheless, other PAMPs have been identifies in studies as targets of CLRs as well e.g. mannose is the recognition motif for many viruses, fungi and mycobacteria; similarly fucose presents the same for certain bacteria and helminths; and glucans are present on mycobacteria and fungi. In addition, many of acquired nonself surfaces e.g. carcinoembryonic/oncofetal type neoantigens carrying "internal danger source"/"self turned nonself" type pathogen pattern are also identified and destroyed (e.g. by complement fixation or other cytotoxic attacks) or sequestered (phagocytosed or ensheathed) by the immune system by virtue of the CLRs. The name lectin is a bit misleading because the family includes proteins with at least one C-type lectin domain (CTLD) which is a specific type of carbohydrate recognition domain. CTLD is a ligand binding motif found in more than 1000 known proteins (more than 100 in humans) and the ligands are often not sugars.<ref>Cummings RD, McEver RP. C-type Lectins. In: Varki A, Cummings RD, Esko JD, et al., editors. Essentials of Glycobiology. 2nd edition. Cold Spring Harbor (NY): Cold Spring Harbor Laboratory Press; 2009. http://www.ncbi.nlm.nih.gov/books/NBK1943/</ref> If and when the ligand is sugar they need Ca2+ – hence the name "C-type", but many of them don't even have a known sugar ligand thus despite carrying a lectin type fold structure, some of them are technically not "lectin" in function.


There are several types of signaling involved in CLRs induced immune response, major connection has been identified between TLR and CLR signaling, therefore we differentiate between TLR-dependent and TLR-independent signaling. DC-SIGN leading to RAF1-MEK-ERK cascade, BDCA2 signaling via ITAM and signaling through ITIM belong among the TLR-dependent signaling. TLR-independent signaling such as Dectin 1, and Dectin 2 - mincle signaling lead to [[MAP kinase]] and [[NFkB]] activation<ref name=":4">{{Cite journal|last=Geijtenbeek|first=Teunis B. H.|last2=Gringhuis|first2=Sonja I.|title=Signalling through C-type lectin receptors: shaping immune responses|url=http://www.nature.com/doifinder/10.1038/nri2569|journal=Nature Reviews Immunology|volume=9|issue=7|pages=465–479|doi=10.1038/nri2569}}</ref><ref>{{Cite journal|last=Hoving|first=J. Claire|last2=Wilson|first2=Gillian J.|last3=Brown|first3=Gordon D.|date=2014-02-01|title=Signalling C-Type lectin receptors, microbial recognition and immunity|url=http://onlinelibrary.wiley.com/doi/10.1111/cmi.12249/abstract|journal=Cellular Microbiology|language=en|volume=16|issue=2|pages=185–194|doi=10.1111/cmi.12249|issn=1462-5822|pmc=PMC4016756|pmid=24330199}}</ref>.
Membrane receptor CLRs are broadly grouped into:


Membrane receptor CLRs have been divided into 17 groups based on structure and phylogenetic origin.<ref>{{Cite journal|last=Zelensky|first=Alex N|last2=Gready|first2=Jill E|date=2005-12-01|title=The C-type lectin-like domain superfamily|url=http://onlinelibrary.wiley.com/doi/10.1111/j.1742-4658.2005.05031.x/abstract|journal=FEBS Journal|language=en|volume=272|issue=24|pages=6179–6217|doi=10.1111/j.1742-4658.2005.05031.x|issn=1742-4658}}</ref> Generally there is a large group, which recognizes and binds carbohydrates, so called carbohydrate recognition domains (CRDs) and the previously mentioned CTLDs.
=====Group I CLRs: The mannose receptors=====

The [[mannose receptor]] (MR) is a PRR primarily present on the surface of [[macrophage]]s and [[dendritic cell]]s. The MR belongs to the multilectin receptor protein group and, like the [[Toll-like receptor|TLRs]], provides a link between innate and adaptive immunity.<ref name="pmid11899091">{{cite journal |vauthors=Apostolopoulos V, McKenzie IF | title = Role of the mannose receptor in the immune response | journal = Curr. Mol. Med. | volume = 1 | issue = 4 | pages = 469–74 |date=September 2001 | pmid = 11899091 | doi = 10.2174/1566524013363645| url = | issn = }}</ref> It recognizes and binds to repeated mannose units on the surfaces of infectious agents and its activation triggers endocytosis and phagocytosis of the microbe via the complement system. Specifically, mannose binding triggers recruitment of MBL-associated serine proteases (MASPs). The serine proteases activate themselves in a cascade, amplifying the immune response: MBL interacts with C4, binding the C4b subunit and releasing C4a into the bloodstream; similarly, binding of C2 causes release of C2b. Together, MBL, C4b and C2a are known as the C3 convertase. C3 is cleaved into its a and b subunits, and C3b binds the convertase. These together are called the C5 convertase. Similarly again, C5b is bound and C5a is released. C5b recruits C6, C7, C8 and multiple C9s. C5, C6, C7, C8 and C9 form the membrane attack complex (MAC).
Another potential characterization of the CLRs can be into mannose receptors and asialoglycoprotein receptors<ref name=":4" />.

===== Group I CLRs: The mannose receptors<ref>{{Cite journal|last=East|first=L|title=The mannose receptor family|url=https://doi.org/10.1016/S0304-4165(02)00319-7|journal=Biochimica et Biophysica Acta (BBA) - General Subjects|volume=1572|issue=2-3|pages=364–386|doi=10.1016/s0304-4165(02)00319-7}}</ref> =====
The [[mannose receptor]] (MR) is a PRR primarily present on the surface of [[macrophage]]s and [[dendritic cell]]s. It belongs into the calcium-dependent multiple CRD group.<ref name=":3" /> The MR belongs to the multilectin receptor protein group and, like the [[Toll-like receptor|TLRs]], provides a link between innate and adaptive immunity.<ref name="pmid11899091">{{cite journal |vauthors=Apostolopoulos V, McKenzie IF | title = Role of the mannose receptor in the immune response | journal = Curr. Mol. Med. | volume = 1 | issue = 4 | pages = 469–74 |date=September 2001 | pmid = 11899091 | doi = 10.2174/1566524013363645| url = | issn = }}</ref><ref>{{Cite journal|last=Vukman|first=Krisztina V.|last2=Ravidà|first2=Alessandra|last3=Aldridge|first3=Allison M.|last4=O'Neill|first4=Sandra M.|date=2013-09-01|title=Mannose receptor and macrophage galactose-type lectin are involved in Bordetella pertussis mast cell interaction|url=http://www.jleukbio.org/content/94/3/439|journal=Journal of Leukocyte Biology|language=en|volume=94|issue=3|pages=439–448|doi=10.1189/jlb.0313130|issn=0741-5400|pmid=23794711}}</ref> It recognizes and binds to repeated mannose units on the surfaces of infectious agents and its activation triggers endocytosis and phagocytosis of the microbe via the complement system. Specifically, mannose binding triggers recruitment of MBL-associated serine proteases (MASPs). The serine proteases activate themselves in a cascade, amplifying the immune response: MBL interacts with C4, binding the C4b subunit and releasing C4a into the bloodstream; similarly, binding of C2 causes release of C2b. Together, MBL, C4b and C2a are known as the C3 convertase. C3 is cleaved into its a and b subunits, and C3b binds the convertase. These together are called the C5 convertase. Similarly again, C5b is bound and C5a is released. C5b recruits C6, C7, C8 and multiple C9s. C5, C6, C7, C8 and C9 form the membrane attack complex (MAC).


=====Group II CLRs: [[asialoglycoprotein]] receptor family=====
=====Group II CLRs: [[asialoglycoprotein]] receptor family=====
This is another large superfamily of CLRs that includes molecules like
This is another large superfamily of CLRs that includes
#The classic asialoglycoprotein receptor [[macrophage galactose-type lectin (MGL)]]
#The classic asialoglycoprotein receptor [[macrophage galactose-type lectin (MGL)]]
#[[DC-SIGN]] (CLEC4L)
#[[DC-SIGN]] (CLEC4L)
Line 54: Line 59:
===Cytoplasmic PRRs===
===Cytoplasmic PRRs===


====NOD-like receptors (NLR)====
===NOD-like receptors (NLR)===
''For more details, see [[NOD-like receptor]].''
''For more details, see [[NOD-like receptor]].''


The NOD-like receptors (NLRs) are cytoplasmic proteins, which recognize bacterial peptidoglycans and mount proinflammatory and antimicrobial immune response<ref name=":5">{{Cite journal|last=Caruso|first=Roberta|last2=Warner|first2=Neil|last3=Inohara|first3=Naohiro|last4=Núñez|first4=Gabriel|title=NOD1 and NOD2: Signaling, Host Defense, and Inflammatory Disease|url=http://dx.doi.org/10.1016/j.immuni.2014.12.010|journal=Immunity|volume=41|issue=6|pages=898–908|doi=10.1016/j.immuni.2014.12.010|pmc=PMC4272446|pmid=25526305}}</ref>. Approximately 20 of these proteins have been found in the mammalian genome and include nucleotide-binding oligomerization domain (NODs), which binds [[nucleoside triphosphate]]. Among other proteins the most important are: the [[MHC Class II]] transactivator ([[CIITA]]), IPAF, BIRC1 etc.<ref name="pmid15870018">{{cite journal |vauthors=Ting JP, Williams KL | title = The CATERPILLER family: an ancient family of immune/apoptotic proteins | journal = Clin. Immunol. | volume = 115 | issue = 1 | pages = 33–37 |date=April 2005 | pmid = 15870018 | doi = 10.1016/j.clim.2005.02.007 | url = | issn = }}</ref>
The NOD-like receptors (NLRs) are cytoplasmic proteins that regulate inflammatory and apoptotic responses. Approximately 20 of these proteins have been found in the mammalian genome and include two major subfamilies called NODs and NALPs, the [[MHC Class II]] transactivator ([[CIITA]]), IPAF, BIRC1, and other molecules.<ref name="pmid15870018">{{cite journal |vauthors=Ting JP, Williams KL | title = The CATERPILLER family: an ancient family of immune/apoptotic proteins | journal = Clin. Immunol. | volume = 115 | issue = 1 | pages = 33–37 |date=April 2005 | pmid = 15870018 | doi = 10.1016/j.clim.2005.02.007 | url = | issn = }}</ref> This family of proteins is greatly expanded in plants, and constitutes a core component of [[Plant disease resistance|plant immune systems]].<ref>{{cite journal |vauthors=Jones DG, Dangl JL | title=The plant immune system | journal=Nature | year=2006 | pmid=17108957 | doi=10.1038/nature05286 | volume=444 | issue=7117 |pages=323–329 | bibcode=2006Natur.444..323J}}</ref> Some of these proteins recognize endogenous or microbial molecules or stress responses and form oligomers that, in animals, activate inflammatory caspases (e.g. [[caspase 1]]) causing cleavage and activation of important inflammatory [[cytokine]]s such as [[Interleukin 1|IL-1]], and/or activate the [[NF-κB]] signaling pathway to induce production of inflammatory molecules. The NLR family is known under several different names, including the CATERPILLER (or CLR) or NOD-LRR family.<ref name="pmid15870018" /><ref name="pmid15952891">{{cite journal | author = Inohara, Inohara, McDonald C, Nuñez G | title = NOD-LRR proteins: role in host-microbial interactions and inflammatory disease | journal = Annu. Rev. Biochem. | volume = 74 | issue = | pages = 355–383 | year = 2005 | pmid = 15952891 | doi = 10.1146/annurev.biochem.74.082803.133347 | url = | issn = }}</ref>

Some of these proteins recognize endogenous or microbial molecules or stress responses and form oligomers that, in animals, activate inflammatory caspases (e.g. [[caspase 1]]) causing cleavage and activation of important inflammatory [[cytokine]]s such as [[Interleukin 1|IL-1]], and/or activate the [[NF-κB]] signaling pathway to induce production of inflammatory molecules. The NLR family is known under several different names, including the CATERPILLER (or CLR) or NOD-LRR family.<ref name="pmid15870018" /><ref name="pmid15952891">{{cite journal|year=2005|title=NOD-LRR proteins: role in host-microbial interactions and inflammatory disease|url=|journal=Annu. Rev. Biochem.|volume=74|issue=|pages=355–383|doi=10.1146/annurev.biochem.74.082803.133347|issn=|pmid=15952891|author=Inohara, Inohara, McDonald C, Nuñez G}}</ref> The most significant members of the NLRs are NOD1 and NOD2. They sense the conserved microbial peptidoglycans in the cytoplasm of the cell and therefore represent another level of immune response after membrane-bound receptors such as TLRs and CLRs.<ref name=":5" />

This family of proteins is greatly expanded in plants, and constitutes a core component of [[Plant disease resistance|plant immune systems]].<ref>{{cite journal |vauthors=Jones DG, Dangl JL | title=The plant immune system | journal=Nature | year=2006 | pmid=17108957 | doi=10.1038/nature05286 | volume=444 | issue=7117 |pages=323–329 | bibcode=2006Natur.444..323J}}</ref>


:;NODs
:;NODs
::The ligands are currently known for [[NOD1]] and [[NOD2]]. NOD1 recognizes a molecule called meso-DAP, which is a [[peptidoglycan]] constituent only of [[Gram negative]] bacteria. NOD2 proteins recognize intracellular MDP (muramyl dipeptide), which is a peptidoglycan constituent of both Gram positive and Gram negative bacteria. NODs transduce signals in the pathway of [[NF-κB]] and [[MAP kinase]]s via the [[serine-threonine kinase]] called RIP2. NOD proteins are so named because they contain a [[nucleotide-binding oligomerization domain]] which binds [[nucleoside triphosphate]]. NODs signal via N-terminal [[CARD domain]]s to activate downstream gene induction events, and interact with microbial molecules by means of a [[C-terminal]] [[leucine-rich repeat]] (LRR) region.<ref name="pmid16493424">{{cite journal |vauthors=Strober W, Murray PJ, Kitani A, Watanabe T | title = Signalling pathways and molecular interactions of NOD1 and NOD2 | journal = Nat. Rev. Immunol. | volume = 6 | issue = 1 | pages = 9–20 |date=January 2006 | pmid = 16493424 | doi = 10.1038/nri1747 | url = | issn = }}</ref>
::The ligands are currently known for [[NOD1]] and [[NOD2]]. NOD1 recognizes a molecule called meso-DAP, which is a [[peptidoglycan]] constituent only of [[Gram negative]] bacteria. NOD2 proteins recognize intracellular MDP (muramyl dipeptide), which is a peptidoglycan constituent of both Gram positive and Gram negative bacteria. When inactive, NODs are in the cytosol in a monomeric state and they undergo conformational change only after ligand recognition, which leads to their activation.<ref name=":5" /> NODs transduce signals in the pathway of [[NF-κB]] and [[MAP kinase]]s via the [[serine-threonine kinase]] called RIP2. NODs signal via N-terminal [[CARD domain]]s to activate downstream gene induction events, and interact with microbial molecules by means of a [[C-terminal]] [[leucine-rich repeat]] (LRR) region.<ref name="pmid16493424">{{cite journal |vauthors=Strober W, Murray PJ, Kitani A, Watanabe T | title = Signalling pathways and molecular interactions of NOD1 and NOD2 | journal = Nat. Rev. Immunol. | volume = 6 | issue = 1 | pages = 9–20 |date=January 2006 | pmid = 16493424 | doi = 10.1038/nri1747 | url = | issn = }}</ref>
::The interaction and cooperation among different types of receptors typical for the innate immune system has been established. An interesting cooperation has been discovered between TLRs and NLRs, particularly between TLR4 and NOD1 in response to ''Escherichia coli'' infection.<ref>{{Cite journal|last=Burberry|first=Aaron|last2=Zeng|first2=Melody Y.|last3=Ding|first3=Lei|last4=Wicks|first4=Ian|last5=Inohara|first5=Naohiro|last6=Morrison|first6=Sean J.|last7=Núñez|first7=Gabriel|title=Infection Mobilizes Hematopoietic Stem Cells through Cooperative NOD-like Receptor and Toll-like Receptor Signaling|url=http://linkinghub.elsevier.com/retrieve/pii/S1931312814001528|journal=Cell Host & Microbe|volume=15|issue=6|pages=779–791|doi=10.1016/j.chom.2014.05.004|pmc=PMC4085166|pmid=24882704}}</ref> Another proof of the cooperation and integration of the entire immune system has been shown in vivo, when TLR signaling was inhibited or disabled, NOD receptors took over role of TLRs.<ref>{{Cite journal|last=Kim|first=Yun-Gi|last2=Park|first2=Jong-Hwan|last3=Shaw|first3=Michael H.|last4=Franchi|first4=Luigi|last5=Inohara|first5=Naohiro|last6=Núñez|first6=Gabriel|title=The Cytosolic Sensors Nod1 and Nod2 Are Critical for Bacterial Recognition and Host Defense after Exposure to Toll-like Receptor Ligands|url=http://linkinghub.elsevier.com/retrieve/pii/S1074761308000307|journal=Immunity|volume=28|issue=2|pages=246–257|doi=10.1016/j.immuni.2007.12.012}}</ref>


===== NLRPs =====
:;NALPs
::Like NODs, these proteins contain C-terminal LRRs, which appear to act as a regulatory domain and may be involved in the recognition of microbial pathogens. Also like NODs, these proteins also contain a nucleotide binding site (NBS) for nucleoside triphosphates. Interaction with other proteins (e.g. the adaptor molecule [[PYCARD|ASC]]) is mediated via N-terminal pyrin (PYD) domain. There are 14 members of this subfamily in humans (called NALP1 to NALP14). Mutations in NALP3 are responsible for the autoinflammatory diseases [[familial cold autoinflammatory syndrome]], [[Muckle–Wells syndrome]] and [[neonatal onset multisystem inflammatory disease]]. Activators of NALP3 include [[muramyl dipeptide]], bacterial DNA, [[Adenosine triphosphate|ATP]], toxins, double stranded [[RNA]], [[paramyxovirus]]es and [[uric acid]] crystals. Although these specific molecules have been shown to activate NALP3, it remains unclear whether this is due to direct binding or due to cellular stress induced by these agents.
::Like NODs, these proteins contain C-terminal LRRs, which appear to act as a regulatory domain and may be involved in the recognition of microbial pathogens. Also like NODs, these proteins contain a nucleotide binding site (NBS) for nucleoside triphosphates. Interaction with other proteins (e.g. the adaptor molecule [[PYCARD|ASC]]) is mediated via N-terminal pyrin (PYD) domain. There are 14 members of this protein subfamily in humans (called NLRP1 to NLRP14). NLRP3 and NLRP4 are responsible for the [[inflammasome]] activation.<ref>{{Cite journal|last=Ip|first=W. K. Eddie|last2=Medzhitov|first2=Ruslan|date=2015-05-11|title=Macrophages monitor tissue osmolarity and induce inflammatory response through NLRP3 and NLRC4 inflammasome activation|url=https://www.nature.com/articles/ncomms7931|journal=Nature Communications|language=en|volume=6|pages=ncomms7931|doi=10.1038/ncomms7931|pmc=PMC4430126|pmid=25959047}}</ref> NLRP3 can be activated adn give rise to NLRP3 inflammasome by ATP, bacterial pore-forming toxins, alum and crystals. Alongside the listed molecules, which lead to activation of NLRP3 inflammasome, the assembly and activation can also be induced by K<sup>+</sup> efflux, Ca<sup>2+</sup> influx, disruption of lysosomes and ROS originating from mitochondria.<ref>{{Cite journal|last=Ip|first=W. K. Eddie|last2=Medzhitov|first2=Ruslan|date=2015-05-11|title=Macrophages monitor tissue osmolarity and induce inflammatory response through NLRP3 and NLRC4 inflammasome activation|url=https://www.nature.com/articles/ncomms7931|journal=Nature Communications|language=en|volume=6|pages=ncomms7931|doi=10.1038/ncomms7931|pmc=PMC4430126|pmid=25959047}}</ref> The NLRP3 inflammasome is essential for induction of effective immune response. The NLRP3 inflammasome can be induced by a wide range of stimuli in contrast to the NLRP4 inflammasome, which binds more limited number and variety of ligands and works in a complex with NAIP protein.<ref>{{Cite journal|last=Guo|first=Haitao|last2=Callaway|first2=Justin B|last3=Ting|first3=Jenny P-Y|title=Inflammasomes: mechanism of action, role in disease, and therapeutics|url=http://www.nature.com/doifinder/10.1038/nm.3893|journal=Nature Medicine|volume=21|issue=7|pages=677–687|doi=10.1038/nm.3893|pmc=PMC4519035|pmid=26121197}}</ref>


:;Other NLRs
:;Other NLRs
::Other NLRs such as IPAF and NAIP5/Birc1e have also been shown to activate caspase-1 in response to [[Salmonella]] and [[Legionella]].
::Other NLRs such as IPAF and NAIP5/Birc1e have also been shown to activate caspase-1 in response to [[Salmonella]] and [[Legionella]].


====RIG-I-like receptors (RLR)====
===RIG-I-like receptors (RLR)===
{{further|RIG-I-like receptor}}
{{further|RIG-I-like receptor}}
Three RLR helicases have so far been described: [[RIG-I]] and [[MDA5]] (recognizing 5'triphosphate-RNA and dsRNA, respectively), which activate antiviral signaling, and [[LGP2]], which appears to act as a dominant-negative inhibitor. RLRs initiate the release of inflammatory cytokines and type I interferon (IFN I)<ref name=":0" />. RLRs, are [[Rna helicases|RNA helicases]], which have been shown to participate in intracellular recognition of viral double-stranded (ds) and single stranded [[RNA]] which recruit factors via twin N-terminal [[CARD domain]]s to activate antiviral gene programs, which may be exploited in therapy of viral infections.<ref>{{Cite journal|last=Pattabhi|first=Sowmya|last2=Wilkins|first2=Courtney R.|last3=Dong|first3=Ran|last4=Knoll|first4=Megan L.|last5=Posakony|first5=Jeffrey|last6=Kaiser|first6=Shari|last7=Mire|first7=Chad E.|last8=Wang|first8=Myra L.|last9=Ireton|first9=Renee C.|date=2016-03-01|title=Targeting Innate Immunity for Antiviral Therapy through Small Molecule Agonists of the RLR Pathway|url=http://jvi.asm.org/content/90/5/2372|journal=Journal of Virology|language=en|volume=90|issue=5|pages=2372–2387|doi=10.1128/jvi.02202-15|issn=0022-538X|pmid=26676770}}</ref><ref>{{Cite journal|last=Eisenächer|first=Katharina|last2=Krug|first2=Anne|title=Regulation of RLR-mediated innate immune signaling – It is all about keeping the balance|url=http://linkinghub.elsevier.com/retrieve/pii/S0171933511000288|journal=European Journal of Cell Biology|volume=91|issue=1|pages=36–47|doi=10.1016/j.ejcb.2011.01.011}}</ref> It has been suggested that the main antiviral program induced by RLR is based on [[ATPase]] activity<ref>{{Cite journal|last=Satoh|first=Takashi|last2=Kato|first2=Hiroki|last3=Kumagai|first3=Yutaro|last4=Yoneyama|first4=Mitsutoshi|last5=Sato|first5=Shintaro|last6=Matsushita|first6=Kazufumi|last7=Tsujimura|first7=Tohru|last8=Fujita|first8=Takashi|last9=Akira|first9=Shizuo|date=2010-01-26|title=LGP2 is a positive regulator of RIG-I– and MDA5-mediated antiviral responses|url=http://www.pnas.org/content/107/4/1512|journal=Proceedings of the National Academy of Sciences|language=en|volume=107|issue=4|pages=1512–1517|doi=10.1073/pnas.0912986107|issn=0027-8424|pmc=PMC2824407|pmid=20080593}}</ref>. RLRs often interact and create cross-talk with the TLRs in the innate immune response and in regulation of adaptive immune response.<ref>{{Cite journal|last=Loo|first=Yueh-Ming|last2=Gale|first2=Michael|title=Immune Signaling by RIG-I-like Receptors|url=http://linkinghub.elsevier.com/retrieve/pii/S1074761311001877|journal=Immunity|volume=34|issue=5|pages=680–692|doi=10.1016/j.immuni.2011.05.003|pmc=PMC3177755|pmid=21616437}}</ref>
Intracellular recognition of viral double-stranded (ds) and single stranded [[RNA]] has been shown to be mediated by a group of [[RNA Helicase]]s which in turn recruit factors via twin N-terminal [[CARD domain]]s to activate antiviral gene programs. Three such helicases have been described in mammals: [[RIG-I]] and [[MDA5]] (recognizing 5'triphosphate-RNA and dsRNA, respectively), which activate antiviral signaling, and [[LGP2]], which appears to act as a [[dominant-negative]] inhibitor.
<gallery>
<gallery>
Image:Simon (RIG-I and Mda5).jpg|RIG-I and Mda5-mediated signalling pathway.
Image:Simon (RIG-I and Mda5).jpg|RIG-I and Mda5-mediated signalling pathway.
</gallery>
</gallery>


====Plant PRRs====
==Plant PRRs==
Plants contain a significant number of PRRs that share remarkable structural and functional similarity with drosophila TOLL and mammalian TLRs. The first PRR identified in plants or animals was the Xa21 protein, conferring resistance to the Gram-negative bacterial pathogen Xanthomonas oryzae pv. oryzae.<ref name="pmid8525370"/><ref name="pmid24482761">{{cite journal |vauthors=Bahar O, Pruitt R, Luu DD, Schwessinger B, Daudi A, Liu F, Ruan R, Fontaine-Bodin L, Koebnik R, Ronald P | title = The Xanthomonas Ax21 protein is processed by the general secretory system and is secreted in association with outer membrane vesicles | journal = PeerJ | volume = 2 | pages = e242| year = 2014 | pmid = 24482761 | doi = 10.7717/peerj.242 | url = https://peerj.com/articles/242/ | issn = | pmc=3897388}}</ref> Since that time two other plants PRRs, Arabidopsis FLS2 (flagellin) and EFR (elongation factor Tu receptor)have been isolated.<ref name="pmid19400727"/> The corresponding PAMPs for FLS2 and EFR have been identified.<ref name="pmid19400727">{{cite journal |vauthors=Boller T, Felix G | title = A renaissance of elicitors: perception of microbe-associated molecular patterns and danger signals by pattern-recognition receptors | journal = Annu Rev Plant Biol | volume = 60 | issue = | pages = 379–406 | year = 2009 | pmid = 19400727 | doi = 10.1146/annurev.arplant.57.032905.105346 | url = | issn = }}</ref> Upon ligand recognition, the plant PRRs transduce "PAMP-triggered immunity" (PTI).<ref name = "pmid16497589">{{cite journal |vauthors=Chisholm ST, Coaker G, Day B, Staskawicz BJ | title = Host-microbe interactions: shaping the evolution of the plant immune response | journal = Cell | year = 2006 | volume = 124 | issue = 4 | pages = 803–814 | pmid = 16497589 | doi = 10.1016/j.cell.2006.02.008 }}</ref> [[Plant disease resistance|Plant immune systems]] also encode resistance proteins that resemble NOD-like receptors (see above), that feature NBS and [[LRR domain]]s and can also carry other conserved interaction domains such as the TIR cytoplasmic domain found in Toll and Interleukin Receptors.<ref name = "pmid16677430">{{cite journal |vauthors=McHale L, Tan X, Koehl P, Michelmore RW | title = Plant NBS-LRR proteins: adaptable guards | journal = Genome Biol | year = 2006 | volume = 7 | issue = 4 | pages = 212 | pmid = 16677430 | doi = 10.1186/gb-2006-7-4-212 | pmc = 1557992 }}</ref> The NBS-LRR proteins are required for effector triggered immunity (ETI).
Plants contain a significant number of PRRs that share remarkable structural and functional similarity with drosophila TOLL and mammalian TLRs. The first PRR identified in plants or animals was the Xa21 protein, conferring resistance to the Gram-negative bacterial pathogen Xanthomonas oryzae pv. oryzae.<ref name="pmid8525370" /><ref name="pmid24482761">{{cite journal |vauthors=Bahar O, Pruitt R, Luu DD, Schwessinger B, Daudi A, Liu F, Ruan R, Fontaine-Bodin L, Koebnik R, Ronald P | title = The Xanthomonas Ax21 protein is processed by the general secretory system and is secreted in association with outer membrane vesicles | journal = PeerJ | volume = 2 | pages = e242| year = 2014 | pmid = 24482761 | doi = 10.7717/peerj.242 | url = https://peerj.com/articles/242/ | issn = | pmc=3897388}}</ref> Since that time two other plants PRRs, Arabidopsis FLS2 (flagellin) and EFR (elongation factor Tu receptor)have been isolated.<ref name="pmid19400727" /> The corresponding PAMPs for FLS2 and EFR have been identified.<ref name="pmid19400727">{{cite journal |vauthors=Boller T, Felix G | title = A renaissance of elicitors: perception of microbe-associated molecular patterns and danger signals by pattern-recognition receptors | journal = Annu Rev Plant Biol | volume = 60 | issue = | pages = 379–406 | year = 2009 | pmid = 19400727 | doi = 10.1146/annurev.arplant.57.032905.105346 | url = | issn = }}</ref> Upon ligand recognition, the plant PRRs transduce "PAMP-triggered immunity" (PTI).<ref name="pmid16497589">{{cite journal |vauthors=Chisholm ST, Coaker G, Day B, Staskawicz BJ | title = Host-microbe interactions: shaping the evolution of the plant immune response | journal = Cell | year = 2006 | volume = 124 | issue = 4 | pages = 803–814 | pmid = 16497589 | doi = 10.1016/j.cell.2006.02.008 }}</ref> [[Plant disease resistance|Plant immune systems]] also encode resistance proteins that resemble NOD-like receptors (see above), that feature NBS and [[LRR domain]]s and can also carry other conserved interaction domains such as the TIR cytoplasmic domain found in Toll and Interleukin Receptors.<ref name="pmid16677430">{{cite journal |vauthors=McHale L, Tan X, Koehl P, Michelmore RW | title = Plant NBS-LRR proteins: adaptable guards | journal = Genome Biol | year = 2006 | volume = 7 | issue = 4 | pages = 212 | pmid = 16677430 | doi = 10.1186/gb-2006-7-4-212 | pmc = 1557992 }}</ref> The NBS-LRR proteins are required for effector triggered immunity (ETI).

=== NonRD kinases ===

PRRs commonly associate with or contain members of a monophyletic group of kinases called the interleukin-1 receptor-associated kinase (IRAK) family that include Drosophila Pelle, human IRAKs, rice XA21 and Arabidopsis FLS2. In mammals, PRRs can also associate with members of the receptor-interacting protein (RIP) kinase family, distant relatives to the IRAK family. Some IRAK and RIP family kinases fall into a small functional class of kinases termed non-RD, many of which do not autophosphorylate the activation loop. A survey of the yeast, fly, worm, human, Arabidopsis, and rice kinomes (3,723 kinases) revealed that despite the small number of non-RD kinases in these genomes (9%–29%), 12 of 15 kinases known or predicted to function in PRR signaling fall into the non-RD class. In plants, all PRRs characterized to date belong to the non-RD class. These data indicate that kinases associated with PRRs can largely be predicted by the lack of a single conserved residue and reveal new potential plant PRR subfamilies.<ref>{{Cite journal|last=Dardick|first=Chris|last2=Schwessinger|first2=Benjamin|last3=Ronald|first3=Pamela|title=Non-arginine-aspartate (non-RD) kinases are associated with innate immune receptors that recognize conserved microbial signatures|url=https://doi.org/10.1016/j.pbi.2012.05.002|journal=Current Opinion in Plant Biology|volume=15|issue=4|pages=358–366|doi=10.1016/j.pbi.2012.05.002}}</ref><ref name="pmid16424920">{{cite journal|date=January 2006|title=Plant and animal pathogen recognition receptors signal through non-RD kinases|url=|journal=PLoS Pathog.|volume=2|issue=1|pages=e2|doi=10.1371/journal.ppat.0020002|issn=|pmc=1331981|pmid=16424920|vauthors=Dardick C, Ronald P}}</ref>


=== Secreted PRRs ===
=== Secreted PRRs ===
A number of PRRs do not remain associated with the cell that produces them. [[Complement (biology)|Complement receptor]]s, [[collectin]]s, [[ficolin]]s, [[pentraxin]]s such as serum [[amyloid]] and [[C-reactive protein]], [[lipid transferase]]s, [[peptidoglycan recognition protein]]s (PGRs) and the LRR, XA21D<ref name="pmid9596635">{{cite journal |vauthors=Wang GL, Ruan DL, Song WY, Sideris S, Chen L, Pi LY, Zhang S, Zhang Z, Fauquet C, Gaut BS, Whalen MC, Ronald PC | title = Xa21D encodes a receptor-like molecule with a leucine-rich repeat domain that determines race-specific recognition and is subject to adaptive evolution | journal = Plant Cell | volume = 10 | issue = 5 | pages = 765–79 |date=May 1998 | pmid = 9596635 | pmc = 144027 | doi = 10.2307/3870663| url = | issn = }}</ref> are all secreted proteins. One very important collectin is [[mannan-binding lectin]] (MBL), a major PRR of the innate immune system that binds to a wide range of bacteria, viruses, fungi and protozoa. MBL predominantly recognizes certain sugar groups on the surface of microorganisms but also binds [[phospholipid]]s, [[nucleic acid]]s and non-[[glycosylation|glycosylated]] proteins.<ref name="pmid16948640">{{cite journal |vauthors=Dommett RM, Klein N, Turner MW | title = Mannose-binding lectin in innate immunity: past, present and future | journal = Tissue Antigens | volume = 68 | issue = 3 | pages = 193–209 |date=September 2006 | pmid = 16948640 | doi = 10.1111/j.1399-0039.2006.00649.x | url = | issn = }}</ref> Once bound to the ligands MBL and Ficolin oligomers recruit [[MASP1 (protein)|MASP1]] and [[MASP2 (protein)|MASP2]] and initiate the [[Mannan-binding lectin pathway|lectin pathway]] of complement activation which is somewhat similar to the [[classical complement pathway]].
A number of PRRs do not remain associated with the cell that produces them. [[Complement (biology)|Complement receptor]]s, [[collectin]]s, [[ficolin]]s, [[pentraxin]]s such as serum [[amyloid]] and [[C-reactive protein]], [[lipid transferase]]s, [[peptidoglycan recognition protein]]s (PGRs) and the LRR, XA21D<ref name="pmid9596635">{{cite journal |vauthors=Wang GL, Ruan DL, Song WY, Sideris S, Chen L, Pi LY, Zhang S, Zhang Z, Fauquet C, Gaut BS, Whalen MC, Ronald PC | title = Xa21D encodes a receptor-like molecule with a leucine-rich repeat domain that determines race-specific recognition and is subject to adaptive evolution | journal = Plant Cell | volume = 10 | issue = 5 | pages = 765–79 |date=May 1998 | pmid = 9596635 | pmc = 144027 | doi = 10.2307/3870663| url = | issn = }}</ref> are all secreted proteins. One very important collectin is [[mannan-binding lectin]] (MBL), a major PRR of the innate immune system that binds to a wide range of bacteria, viruses, fungi and protozoa. MBL predominantly recognizes certain sugar groups on the surface of microorganisms but also binds [[phospholipid]]s, [[nucleic acid]]s and non-[[glycosylation|glycosylated]] proteins.<ref name="pmid16948640">{{cite journal |vauthors=Dommett RM, Klein N, Turner MW | title = Mannose-binding lectin in innate immunity: past, present and future | journal = Tissue Antigens | volume = 68 | issue = 3 | pages = 193–209 |date=September 2006 | pmid = 16948640 | doi = 10.1111/j.1399-0039.2006.00649.x | url = | issn = }}</ref> Once bound to the ligands MBL and Ficolin oligomers recruit [[MASP1 (protein)|MASP1]] and [[MASP2 (protein)|MASP2]] and initiate the [[Mannan-binding lectin pathway|lectin pathway]] of complement activation which is somewhat similar to the [[classical complement pathway]].


== Signaling via PRR ==
== PRRs in human medicine ==
Research groups have recently conducted extensive research into the involvement and potential use of patient's immune system in the therapy of various diseases, the so called immunotherapy, including monoclonal antibodies, non-specific immunotherapies, oncolytic virus therapy, T-cell therapy and cancer vaccines.<ref>{{Cite news|url=http://www.cancer.net/navigating-cancer-care/how-cancer-treated/immunotherapy-and-vaccines/understanding-immunotherapy|title=Understanding Immunotherapy|date=2013-03-25|work=Cancer.Net|access-date=2017-07-31|language=en}}</ref> NOD2 has been associated through a loss- and gain- of function with development of Crohn's disease and early-onset sarcoidosis<ref name=":5" /><ref>{{Cite journal|last=Chen|first=Edward S.|title=Innate immunity in sarcoidosis pathobiology|url=http://content.wkhealth.com/linkback/openurl?sid=WKPTLP:landingpage&an=00063198-201609000-00013|journal=Current Opinion in Pulmonary Medicine|volume=22|issue=5|pages=469–475|doi=10.1097/mcp.0000000000000305}}</ref>. Mutations in NOD2 in cooperation with enviromental factors lead to development of chronic inflammation in the intestine. <ref name=":5" /><ref>{{Cite journal|last=Philpott|first=Dana J.|last2=Sorbara|first2=Matthew T.|last3=Robertson|first3=Susan J.|last4=Croitoru|first4=Kenneth|last5=Girardin|first5=Stephen E.|title=NOD proteins: regulators of inflammation in health and disease|url=http://www.nature.com/doifinder/10.1038/nri3565|journal=Nature Reviews Immunology|volume=14|issue=1|pages=9–23|doi=10.1038/nri3565}}</ref> Therefore, it has been suggested to treat the disease by inhibiting the small molecules, which are able to modulate the NOD2 signaling, particularly RIP2. Two therapeutics have been approved by FDA so far inhibiting the phosphorylation on RIP2, which is necessary for proper NOD2 functioning, gefitinib and erlotinib. <ref>{{Cite journal|last=Jun|first=Janice C.|last2=Cominelli|first2=Fabio|last3=Abbott|first3=Derek W.|date=2013-11-01|title=RIP2 activity in inflammatory disease and implications for novel therapeutics|url=http://www.jleukbio.org/content/94/5/927|journal=Journal of Leukocyte Biology|language=en|volume=94|issue=5|pages=927–932|doi=10.1189/jlb.0213109|issn=0741-5400|pmc=PMC3800061|pmid=23794710}}</ref><ref>{{Cite journal|last=Tigno-Aranjuez|first=Justine T.|last2=Benderitter|first2=Pascal|last3=Rombouts|first3=Frederik|last4=Deroose|first4=Frederik|last5=Bai|first5=XiaoDong|last6=Mattioli|first6=Benedetta|last7=Cominelli|first7=Fabio|last8=Pizarro|first8=Theresa T.|last9=Hoflack|first9=Jan|date=2014-10-24|title=In Vivo Inhibition of RIPK2 Kinase Alleviates Inflammatory Disease|url=http://www.jbc.org/content/289/43/29651|journal=Journal of Biological Chemistry|language=en|volume=289|issue=43|pages=29651–29664|doi=10.1074/jbc.m114.591388|issn=0021-9258|pmid=25213858}}</ref> Additionaly, research has been conducted on GSK583, a highly specific RIP2 inhibitor, which seems highly promising in inhibiting NOD1 and NOD2 signaling and therefore, limiting inflammation caused by NOD1, NOD2 signaling pathways.<ref>{{Cite journal|last=Haile|first=Pamela A.|last2=Votta|first2=Bartholomew J.|last3=Marquis|first3=Robert W.|last4=Bury|first4=Michael J.|last5=Mehlmann|first5=John F.|last6=Singhaus|first6=Robert|last7=Charnley|first7=Adam K.|last8=Lakdawala|first8=Ami S.|last9=Convery|first9=Máire A.|date=2016-05-26|title=The Identification and Pharmacological Characterization of 6-(tert-Butylsulfonyl)-N-(5-fluoro-1H-indazol-3-yl)quinolin-4-amine (GSK583), a Highly Potent and Selective Inhibitor of RIP2 Kinase|url=http://dx.doi.org/10.1021/acs.jmedchem.6b00211|journal=Journal of Medicinal Chemistry|volume=59|issue=10|pages=4867–4880|doi=10.1021/acs.jmedchem.6b00211|issn=0022-2623}}</ref> Another possibility is to remove the sensor for NOD2, which has been proved efficient in murine models in the effort to suppress the symptoms of Crohn's disease.<ref>{{Cite journal|last=Corridoni|first=D|last2=Rodriguez-Palacios|first2=A|last3=Stefano|first3=G Di|last4=Martino|first4=L Di|last5=Antonopoulos|first5=D A|last6=Chang|first6=E B|last7=Arseneau|first7=K O|last8=Pizarro|first8=T T|last9=Cominelli|first9=F|title=Genetic deletion of the bacterial sensor NOD2 improves murine Crohn’s disease-like ileitis independent of functional dysbiosis|url=http://www.nature.com/doifinder/10.1038/mi.2016.98|journal=Mucosal Immunology|volume=10|issue=4|pages=971–982|doi=10.1038/mi.2016.98|pmc=PMC5433921|pmid=27848951}}</ref> Type II kinase inhibitors, which are highly specific, have shown promising results in blocking the TNF arising from NOD-dependent pathways, which is shows a high potential in treatment of inflammation associated tumors.<ref>{{Cite journal|last=Canning|first=Peter|last2=Ruan|first2=Qui|last3=Schwerd|first3=Tobias|last4=Hrdinka|first4=Matous|last5=Maki|first5=Jenny L.|last6=Saleh|first6=Danish|last7=Suebsuwong|first7=Chalada|last8=Ray|first8=Soumya|last9=Brennan|first9=Paul E.|title=Inflammatory Signaling by NOD-RIPK2 Is Inhibited by Clinically Relevant Type II Kinase Inhibitors|url=http://dx.doi.org/10.1016/j.chembiol.2015.07.017|journal=Chemistry & Biology|volume=22|issue=9|pages=1174–1184|doi=10.1016/j.chembiol.2015.07.017|pmc=PMC4579271|pmid=26320862}}</ref>


Another possible exploitation of PRRs in human medicine is also related to tumor malignancies of the intestines. ''Helicobacter pylori'' has been shown by studies to significantly correlate with the development of a gastrointestinal tumors. In a healthy individual ''Helicobacter pylori'' infection is targeted by the combination of PRRs, namely TLRs, NLRs, RLRs and CLR DC-SIGN. In case of their malfunction, these receptors have also been connected to carcinogenesis. When the ''Helicobacter pylori'' infection is left to progress in the intestine it develops into chronic inflammation, atrophy and eventually dysplasia leading to development of cancer. Since all types of PRRs play a role in the identification and eradication of the infection, their specific agonists mount a strong immune response to cancers and other PRR-related diseases. The inhibition of TLR2 has been shown to significantly correlate with improved state of the patient and suppression of the gastric adenocarcinoma.<ref>{{Cite journal|last=Castaño-Rodríguez|first=Natalia|last2=Kaakoush|first2=Nadeem O.|last3=Mitchell|first3=Hazel M.|date=2014|title=Pattern-Recognition Receptors and Gastric Cancer|url=http://journal.frontiersin.org/article/10.3389/fimmu.2014.00336/abstract|journal=Frontiers in Immunology|language=English|volume=5|doi=10.3389/fimmu.2014.00336|issn=1664-3224|pmc=PMC4105827|pmid=25101079}}</ref>
===NonRD kinases===


The PRRs are also tightly connected to the proper function of neuronal networks and tissues, especially because of their involvement in the processes of inflammation, which are essential for proper function but may cause irreparable damage if not under control. The TLRs are expressed on most cells of the central nervous system (CNS) and they play a crucial role in sterile inflammation. After an injury, they lead to impairment of axonal growth and slow down or even halt the recovery altogether. Another important structure involved in and potentially exploitable in therapy after injury is the inflammasome. Through its induction of proinflammatory cytokines, IL-1β and IL-18 it has been proposed, that inhibition of inflammasome may also serve as an efficient therapeutic method.<ref>{{Cite journal|last=Kigerl|first=Kristina A.|last2=Vaccari|first2=Juan Pablo de Rivero|last3=Dietrich|first3=W. Dalton|last4=Popovich|first4=Phillip G.|last5=Keane|first5=Robert W.|title=Pattern recognition receptors and central nervous system repair|url=http://dx.doi.org/10.1016/j.expneurol.2014.01.001|journal=Experimental Neurology|volume=258|pages=5–16|doi=10.1016/j.expneurol.2014.01.001|pmc=PMC4974939|pmid=25017883}}</ref> The involvement of inflammasome has also been researched in several other diseases including experimental autoimmune encephalomyelitis (EAE), Alzheimer's and Parkinson's diseases and in atherosclerosis connected with type II diabetes in patients. The suggested therapies include degradation of NLRP3 or inhibit the proinflammatory cytokines.<ref>{{Cite journal|last=Kigerl|first=Kristina A.|last2=Vaccari|first2=Juan Pablo de Rivero|last3=Dietrich|first3=W. Dalton|last4=Popovich|first4=Phillip G.|last5=Keane|first5=Robert W.|title=Pattern recognition receptors and central nervous system repair|url=http://dx.doi.org/10.1016/j.expneurol.2014.01.001|journal=Experimental Neurology|volume=258|pages=5–16|doi=10.1016/j.expneurol.2014.01.001|pmc=PMC4974939|pmid=25017883}}</ref>
PRRs commonly associate with or contain members of a monophyletic group of kinases called the interleukin-1 receptor-associated kinase (IRAK) family that include Drosophila Pelle, human IRAKs, rice XA21 and Arabidopsis FLS2. In mammals, PRRs can also associate with members of the receptor-interacting protein (RIP) kinase family, distant relatives to the IRAK family. Some IRAK and RIP family kinases fall into a small functional class of kinases termed non-RD, many of which do not autophosphorylate the activation loop. A survey of the yeast, fly, worm, human, Arabidopsis, and rice kinomes (3,723 kinases) revealed that despite the small number of non-RD kinases in these genomes (9%–29%), 12 of 15 kinases known or predicted to function in PRR signaling fall into the non-RD class.<ref name="pmid16424920">{{cite journal |vauthors=Dardick C, Ronald P | title = Plant and animal pathogen recognition receptors signal through non-RD kinases | journal = PLoS Pathog. | volume = 2 | issue = 1 | pages = e2 |date=January 2006 | pmid = 16424920 | pmc = 1331981 | doi = 10.1371/journal.ppat.0020002 | url = | issn = }}</ref> In plants, all PRRs characterized to date belong to the non-RD class. These data indicate that kinases associated with PRRs can largely be predicted by the lack of a single conserved residue and reveal new potential plant PRR subfamilies.


==References==
==References==

Revision as of 09:07, 31 July 2017

Pattern recognition receptors (PRRs)[1] play a crucial role in the proper function of the innate immune system. PRRs are germline-encoded host sensors, which detect molecules typical for the pathogens[2]. They are proteins expressed by cells of the innate immune system, such as dendritic cells, macrophages, monocytes, neutrophils and epithelial cells[3][4], to identify two classes of molecules: pathogen-associated molecular patterns (PAMPs), which are associated with microbial pathogens, and damage-associated molecular patterns (DAMPs), which are associated with components of host's cells that are released during cell damage or death. They are also called primitive pattern recognition receptors because they evolved before other parts of the immune system, particularly before adaptive immunity. PRRs also mediate the initiation of antigen-specific adaptive immune response and release of inflammatory cytokines[2][5]

Molecules recognized

The microbe-specific molecules that are recognized by a given PRR are called pathogen-associated molecular patterns (PAMPs) and include bacterial carbohydrates (such as lipopolysaccharide or LPS, mannose), nucleic acids (such as bacterial or viral DNA or RNA), bacterial peptides (flagellin, microtubule elongation factors), peptidoglycans and lipoteichoic acids (from Gram-positive bacteria), N-formylmethionine, lipoproteins and fungal glucans and chitin.

Endogenous stress signals are called damage-associated molecular patterns (DAMPs) and include uric acid and extracellular ATP, among many other compounds[2].

Classification

There are several subgroups of PRRs. They are classified according to their ligand specificity, function, localization and/or evolutionary relationships. Based on their localization, PRRs may be divided into membrane-bound PRRs and cytoplasmic PRRs.

PRR types and localization

Membrane-bound PRRs

Receptor kinases

PRRs were first discovered in plants.[6] Since that time many plant PRRs have been predicted by genomic analysis (370 in rice; 47 in Arabidopsis). Unlike animal PRRs, which associated with intracellular kinases via adaptor proteins (see non-RD kinases below), plant PRRs are composed of an extracellular domain, transmembrane domain, juxtamembrane domain and intracellular kinase domain as part of a single protein.

Toll-like receptors (TLR)

Recognition of extracellular or endosomal pathogen-associated molecular patterns is mediated by transmembrane proteins known as toll-like receptors (TLRs).[7] TLRs share a typical structural motif, the leucine rich repeats (LRR), which give them their specific appearance and are also responsible for TLR functionality[8]. Toll-like receptors were first discovered in Drosophila and trigger the synthesis and secretion of cytokines and activation of other host defense programs that are necessary for both innate or adaptive immune responses. 10 functional members of the TLR family have been described in humans so far[9]. Studies have been conducted on TLR11 as well, and it has been shown that it recognizes flagellin and profilin-like proteins in mice[10]. Nonetheless, TLR11 is only a pseudogene in humans without direct function or functional protein expression. Each of the TLR has been shown to interact with a specific PAMP[5][11][12].

The TLR signaling mechanism

TLRs tend to dimerize, TLR4 forms homodimers, and TLR6 can dimerize with either TLR1 or TLR2. [11] Interaction of TLRs with their specific PAMP is mediated through either MyD88- dependent pathway and triggers the signaling through NF-κB and the MAP kinase pathway and therefore the secretion of pro-inflammatory cytokines and co-stimulatory molecules or TRIF - dependent signaling pathway[2][5][11]. MyD88 - dependent pathway is induced by various PAMPs stimulating the TLRs on macrophages and dendritic cells. MyD88 attracts the IRAK4 molecule, IRAK4 recruits IRAK1 and IRAK2 to form a signaling complex. The signaling complex reacts with TRAF6 which leads to TAK1 activation and consequently the induction of inflammatory cytokines.The TRIF-dependent pathway is induced by macrophages and DCs after TLR3 and TLR4 stimulation.[2] Molecules released following TLR activation signal to other cells of the immune system making TLRs key elements of innate immunity and adaptive immunity.[2][13][14]

C-type lectin receptors (CLR)

Many different cells of the innate immune system express a myriad of CLRs which shape innate immunity by virtue of their pattern recognition ability.[15] Even though, most classes of human pathogens are covered by CLRs, CLRs are a major receptor for recognition of funghi[16][17]: nonetheless, other PAMPs have been identifies in studies as targets of CLRs as well e.g. mannose is the recognition motif for many viruses, fungi and mycobacteria; similarly fucose presents the same for certain bacteria and helminths; and glucans are present on mycobacteria and fungi. In addition, many of acquired nonself surfaces e.g. carcinoembryonic/oncofetal type neoantigens carrying "internal danger source"/"self turned nonself" type pathogen pattern are also identified and destroyed (e.g. by complement fixation or other cytotoxic attacks) or sequestered (phagocytosed or ensheathed) by the immune system by virtue of the CLRs. The name lectin is a bit misleading because the family includes proteins with at least one C-type lectin domain (CTLD) which is a specific type of carbohydrate recognition domain. CTLD is a ligand binding motif found in more than 1000 known proteins (more than 100 in humans) and the ligands are often not sugars.[18] If and when the ligand is sugar they need Ca2+ – hence the name "C-type", but many of them don't even have a known sugar ligand thus despite carrying a lectin type fold structure, some of them are technically not "lectin" in function.

There are several types of signaling involved in CLRs induced immune response, major connection has been identified between TLR and CLR signaling, therefore we differentiate between TLR-dependent and TLR-independent signaling. DC-SIGN leading to RAF1-MEK-ERK cascade, BDCA2 signaling via ITAM and signaling through ITIM belong among the TLR-dependent signaling. TLR-independent signaling such as Dectin 1, and Dectin 2 - mincle signaling lead to MAP kinase and NFkB activation[19][20].

Membrane receptor CLRs have been divided into 17 groups based on structure and phylogenetic origin.[21] Generally there is a large group, which recognizes and binds carbohydrates, so called carbohydrate recognition domains (CRDs) and the previously mentioned CTLDs.

Another potential characterization of the CLRs can be into mannose receptors and asialoglycoprotein receptors[19].

Group I CLRs: The mannose receptors[22]

The mannose receptor (MR) is a PRR primarily present on the surface of macrophages and dendritic cells. It belongs into the calcium-dependent multiple CRD group.[16] The MR belongs to the multilectin receptor protein group and, like the TLRs, provides a link between innate and adaptive immunity.[23][24] It recognizes and binds to repeated mannose units on the surfaces of infectious agents and its activation triggers endocytosis and phagocytosis of the microbe via the complement system. Specifically, mannose binding triggers recruitment of MBL-associated serine proteases (MASPs). The serine proteases activate themselves in a cascade, amplifying the immune response: MBL interacts with C4, binding the C4b subunit and releasing C4a into the bloodstream; similarly, binding of C2 causes release of C2b. Together, MBL, C4b and C2a are known as the C3 convertase. C3 is cleaved into its a and b subunits, and C3b binds the convertase. These together are called the C5 convertase. Similarly again, C5b is bound and C5a is released. C5b recruits C6, C7, C8 and multiple C9s. C5, C6, C7, C8 and C9 form the membrane attack complex (MAC).

Group II CLRs: asialoglycoprotein receptor family

This is another large superfamily of CLRs that includes

  1. The classic asialoglycoprotein receptor macrophage galactose-type lectin (MGL)
  2. DC-SIGN (CLEC4L)
  3. Langerin (CLEC4K)
  4. Myeloid DAP12‑associating lectin (MDL)‑1 (CLEC5A)
  5. DC‑associated C‑type lectin 1 (Dectin1) subfamily which includes
    1. dectin 1/CLEC7A
    2. DNGR1/CLEC9A
    3. Myeloid C‑type lectin‑like receptor (MICL) (CLEC12A)
    4. CLEC2 (also called CLEC1B)- the platelet activation receptor for podoplanin on lymphatic endothelial cells and invading front of some carcinomas.
    5. CLEC12B
  6. DC immunoreceptor (DCIR) subfamily which includes:
    1. DCIR/CLEC4A
    2. Dectin 2/CLEC6A
    3. Blood DC antigen 2 (BDCA2) ( CLEC4C)
    4. Mincle i.e. macrophage‑inducible C‑type lectin (CLEC4E)

The nomenclature (mannose versus asialoglycoprotein) is a bit misleading as these the asialoglycoprotein receptors are not necessarily galactose (one of the commonest outer residues of asialo-glycoprotein) specific receptors and even many of this family members can also bind to mannose after which the other group is named.

Cytoplasmic PRRs

NOD-like receptors (NLR)

For more details, see NOD-like receptor.

The NOD-like receptors (NLRs) are cytoplasmic proteins, which recognize bacterial peptidoglycans and mount proinflammatory and antimicrobial immune response[25]. Approximately 20 of these proteins have been found in the mammalian genome and include nucleotide-binding oligomerization domain (NODs), which binds nucleoside triphosphate. Among other proteins the most important are: the MHC Class II transactivator (CIITA), IPAF, BIRC1 etc.[26]

Some of these proteins recognize endogenous or microbial molecules or stress responses and form oligomers that, in animals, activate inflammatory caspases (e.g. caspase 1) causing cleavage and activation of important inflammatory cytokines such as IL-1, and/or activate the NF-κB signaling pathway to induce production of inflammatory molecules. The NLR family is known under several different names, including the CATERPILLER (or CLR) or NOD-LRR family.[26][27] The most significant members of the NLRs are NOD1 and NOD2. They sense the conserved microbial peptidoglycans in the cytoplasm of the cell and therefore represent another level of immune response after membrane-bound receptors such as TLRs and CLRs.[25]

This family of proteins is greatly expanded in plants, and constitutes a core component of plant immune systems.[28]

NODs
The ligands are currently known for NOD1 and NOD2. NOD1 recognizes a molecule called meso-DAP, which is a peptidoglycan constituent only of Gram negative bacteria. NOD2 proteins recognize intracellular MDP (muramyl dipeptide), which is a peptidoglycan constituent of both Gram positive and Gram negative bacteria. When inactive, NODs are in the cytosol in a monomeric state and they undergo conformational change only after ligand recognition, which leads to their activation.[25] NODs transduce signals in the pathway of NF-κB and MAP kinases via the serine-threonine kinase called RIP2. NODs signal via N-terminal CARD domains to activate downstream gene induction events, and interact with microbial molecules by means of a C-terminal leucine-rich repeat (LRR) region.[29]
The interaction and cooperation among different types of receptors typical for the innate immune system has been established. An interesting cooperation has been discovered between TLRs and NLRs, particularly between TLR4 and NOD1 in response to Escherichia coli infection.[30] Another proof of the cooperation and integration of the entire immune system has been shown in vivo, when TLR signaling was inhibited or disabled, NOD receptors took over role of TLRs.[31]
NLRPs
Like NODs, these proteins contain C-terminal LRRs, which appear to act as a regulatory domain and may be involved in the recognition of microbial pathogens. Also like NODs, these proteins contain a nucleotide binding site (NBS) for nucleoside triphosphates. Interaction with other proteins (e.g. the adaptor molecule ASC) is mediated via N-terminal pyrin (PYD) domain. There are 14 members of this protein subfamily in humans (called NLRP1 to NLRP14). NLRP3 and NLRP4 are responsible for the inflammasome activation.[32] NLRP3 can be activated adn give rise to NLRP3 inflammasome by ATP, bacterial pore-forming toxins, alum and crystals. Alongside the listed molecules, which lead to activation of NLRP3 inflammasome, the assembly and activation can also be induced by K+ efflux, Ca2+ influx, disruption of lysosomes and ROS originating from mitochondria.[33] The NLRP3 inflammasome is essential for induction of effective immune response. The NLRP3 inflammasome can be induced by a wide range of stimuli in contrast to the NLRP4 inflammasome, which binds more limited number and variety of ligands and works in a complex with NAIP protein.[34]
Other NLRs
Other NLRs such as IPAF and NAIP5/Birc1e have also been shown to activate caspase-1 in response to Salmonella and Legionella.

RIG-I-like receptors (RLR)

Three RLR helicases have so far been described: RIG-I and MDA5 (recognizing 5'triphosphate-RNA and dsRNA, respectively), which activate antiviral signaling, and LGP2, which appears to act as a dominant-negative inhibitor. RLRs initiate the release of inflammatory cytokines and type I interferon (IFN I)[2]. RLRs, are RNA helicases, which have been shown to participate in intracellular recognition of viral double-stranded (ds) and single stranded RNA which recruit factors via twin N-terminal CARD domains to activate antiviral gene programs, which may be exploited in therapy of viral infections.[35][36] It has been suggested that the main antiviral program induced by RLR is based on ATPase activity[37]. RLRs often interact and create cross-talk with the TLRs in the innate immune response and in regulation of adaptive immune response.[38]

Plant PRRs

Plants contain a significant number of PRRs that share remarkable structural and functional similarity with drosophila TOLL and mammalian TLRs. The first PRR identified in plants or animals was the Xa21 protein, conferring resistance to the Gram-negative bacterial pathogen Xanthomonas oryzae pv. oryzae.[6][39] Since that time two other plants PRRs, Arabidopsis FLS2 (flagellin) and EFR (elongation factor Tu receptor)have been isolated.[40] The corresponding PAMPs for FLS2 and EFR have been identified.[40] Upon ligand recognition, the plant PRRs transduce "PAMP-triggered immunity" (PTI).[41] Plant immune systems also encode resistance proteins that resemble NOD-like receptors (see above), that feature NBS and LRR domains and can also carry other conserved interaction domains such as the TIR cytoplasmic domain found in Toll and Interleukin Receptors.[42] The NBS-LRR proteins are required for effector triggered immunity (ETI).

NonRD kinases

PRRs commonly associate with or contain members of a monophyletic group of kinases called the interleukin-1 receptor-associated kinase (IRAK) family that include Drosophila Pelle, human IRAKs, rice XA21 and Arabidopsis FLS2. In mammals, PRRs can also associate with members of the receptor-interacting protein (RIP) kinase family, distant relatives to the IRAK family. Some IRAK and RIP family kinases fall into a small functional class of kinases termed non-RD, many of which do not autophosphorylate the activation loop. A survey of the yeast, fly, worm, human, Arabidopsis, and rice kinomes (3,723 kinases) revealed that despite the small number of non-RD kinases in these genomes (9%–29%), 12 of 15 kinases known or predicted to function in PRR signaling fall into the non-RD class. In plants, all PRRs characterized to date belong to the non-RD class. These data indicate that kinases associated with PRRs can largely be predicted by the lack of a single conserved residue and reveal new potential plant PRR subfamilies.[43][44]

Secreted PRRs

A number of PRRs do not remain associated with the cell that produces them. Complement receptors, collectins, ficolins, pentraxins such as serum amyloid and C-reactive protein, lipid transferases, peptidoglycan recognition proteins (PGRs) and the LRR, XA21D[45] are all secreted proteins. One very important collectin is mannan-binding lectin (MBL), a major PRR of the innate immune system that binds to a wide range of bacteria, viruses, fungi and protozoa. MBL predominantly recognizes certain sugar groups on the surface of microorganisms but also binds phospholipids, nucleic acids and non-glycosylated proteins.[46] Once bound to the ligands MBL and Ficolin oligomers recruit MASP1 and MASP2 and initiate the lectin pathway of complement activation which is somewhat similar to the classical complement pathway.

PRRs in human medicine

Research groups have recently conducted extensive research into the involvement and potential use of patient's immune system in the therapy of various diseases, the so called immunotherapy, including monoclonal antibodies, non-specific immunotherapies, oncolytic virus therapy, T-cell therapy and cancer vaccines.[47] NOD2 has been associated through a loss- and gain- of function with development of Crohn's disease and early-onset sarcoidosis[25][48]. Mutations in NOD2 in cooperation with enviromental factors lead to development of chronic inflammation in the intestine. [25][49] Therefore, it has been suggested to treat the disease by inhibiting the small molecules, which are able to modulate the NOD2 signaling, particularly RIP2. Two therapeutics have been approved by FDA so far inhibiting the phosphorylation on RIP2, which is necessary for proper NOD2 functioning, gefitinib and erlotinib. [50][51] Additionaly, research has been conducted on GSK583, a highly specific RIP2 inhibitor, which seems highly promising in inhibiting NOD1 and NOD2 signaling and therefore, limiting inflammation caused by NOD1, NOD2 signaling pathways.[52] Another possibility is to remove the sensor for NOD2, which has been proved efficient in murine models in the effort to suppress the symptoms of Crohn's disease.[53] Type II kinase inhibitors, which are highly specific, have shown promising results in blocking the TNF arising from NOD-dependent pathways, which is shows a high potential in treatment of inflammation associated tumors.[54]

Another possible exploitation of PRRs in human medicine is also related to tumor malignancies of the intestines. Helicobacter pylori has been shown by studies to significantly correlate with the development of a gastrointestinal tumors. In a healthy individual Helicobacter pylori infection is targeted by the combination of PRRs, namely TLRs, NLRs, RLRs and CLR DC-SIGN. In case of their malfunction, these receptors have also been connected to carcinogenesis. When the Helicobacter pylori infection is left to progress in the intestine it develops into chronic inflammation, atrophy and eventually dysplasia leading to development of cancer. Since all types of PRRs play a role in the identification and eradication of the infection, their specific agonists mount a strong immune response to cancers and other PRR-related diseases. The inhibition of TLR2 has been shown to significantly correlate with improved state of the patient and suppression of the gastric adenocarcinoma.[55]

The PRRs are also tightly connected to the proper function of neuronal networks and tissues, especially because of their involvement in the processes of inflammation, which are essential for proper function but may cause irreparable damage if not under control. The TLRs are expressed on most cells of the central nervous system (CNS) and they play a crucial role in sterile inflammation. After an injury, they lead to impairment of axonal growth and slow down or even halt the recovery altogether. Another important structure involved in and potentially exploitable in therapy after injury is the inflammasome. Through its induction of proinflammatory cytokines, IL-1β and IL-18 it has been proposed, that inhibition of inflammasome may also serve as an efficient therapeutic method.[56] The involvement of inflammasome has also been researched in several other diseases including experimental autoimmune encephalomyelitis (EAE), Alzheimer's and Parkinson's diseases and in atherosclerosis connected with type II diabetes in patients. The suggested therapies include degradation of NLRP3 or inhibit the proinflammatory cytokines.[57]

References

  1. ^ Mahla, R.S. (2013). "Sweeten PAMPs: Role of Sugar Complexed PAMPs in Innate Immunity and Vaccine Biology". Front Immunol. 4: 248. doi:10.3389/fimmu.2013.00248. PMC 3759294. PMID 24032031.{{cite journal}}: CS1 maint: unflagged free DOI (link)
  2. ^ a b c d e f g Kumar, Himanshu; Kawai, Taro; Akira, Shizuo (2011-01-01). "Pathogen Recognition by the Innate Immune System". International Reviews of Immunology. 30 (1): 16–34. doi:10.3109/08830185.2010.529976. ISSN 0883-0185. PMID 21235323.
  3. ^ Alberts, Bruce; Johnson, Alexander; Lewis, Julian; Raff, Martin; Roberts, Keith; Walter, Peter (2002). "Innate Immunity". {{cite journal}}: Cite journal requires |journal= (help)
  4. ^ Schroder, Kate; Tschopp, Jurg. "The Inflammasomes". Cell. 140 (6): 821–832. doi:10.1016/j.cell.2010.01.040.
  5. ^ a b c Takeda, Kiyoshi; Kaisho, Tsuneyasu; Akira, Shizuo (2003-11-28). "Toll-Like Receptors". https://doi.org/10.1146/annurev.immunol.21.120601.141126. doi:10.1146/annurev.immunol.21.120601.141126. Retrieved 2017-07-27. {{cite web}}: External link in |website= (help)
  6. ^ a b Song WY, Wang GL, Chen LL, Kim HS, Pi LY, Holsten T, Gardner J, Wang B, Zhai WX, Zhu LH, Fauquet C, Ronald P (December 1995). "A receptor kinase-like protein encoded by the rice disease resistance gene, Xa21". Science. 270 (5243): 1804–6. Bibcode:1995Sci...270.1804S. doi:10.1126/science.270.5243.1804. PMID 8525370.
  7. ^ Beutler B, Jiang Z, Georgel P, Crozat K, Croker B, Rutschmann S, Du X, Hoebe K (2006). "Genetic analysis of host resistance: Toll-like receptor signaling and immunity at large". Annu. Rev. Immunol. 24: 353–389. doi:10.1146/annurev.immunol.24.021605.090552. PMID 16551253.
  8. ^ Botos, Istvan; Segal, David M.; Davies, David R. "The Structural Biology of Toll-like Receptors". Structure. 19 (4): 447–459. doi:10.1016/j.str.2011.02.004. PMC 3075535. PMID 21481769.{{cite journal}}: CS1 maint: PMC format (link)
  9. ^ Takeda, Kiyoshi; Kaisho, Tsuneyasu; Akira, Shizuo (2003-11-28). "Toll-Like Receptors". https://doi.org/10.1146/annurev.immunol.21.120601.141126. doi:10.1146/annurev.immunol.21.120601.141126. Retrieved 2017-07-27. {{cite web}}: External link in |website= (help)
  10. ^ Hatai, Hirotsugu; Lepelley, Alice; Zeng, Wangyong; Hayden, Matthew S.; Ghosh, Sankar (2016-02-09). "Toll-Like Receptor 11 (TLR11) Interacts with Flagellin and Profilin through Disparate Mechanisms". PLOS ONE. 11 (2): e0148987. doi:10.1371/journal.pone.0148987. ISSN 1932-6203. PMC 4747465. PMID 26859749.{{cite journal}}: CS1 maint: PMC format (link) CS1 maint: unflagged free DOI (link)
  11. ^ a b c Ozinsky, Adrian; Underhill, David M.; Fontenot, Jason D.; Hajjar, Adeline M.; Smith, Kelly D.; Wilson, Christopher B.; Schroeder, Lea; Aderem, Alan (2000-12-05). "The repertoire for pattern recognition of pathogens by the innate immune system is defined by cooperation between Toll-like receptors". Proceedings of the National Academy of Sciences. 97 (25): 13766–13771. doi:10.1073/pnas.250476497. ISSN 0027-8424. PMC 17650. PMID 11095740.{{cite journal}}: CS1 maint: PMC format (link)
  12. ^ Lien, E.; Sellati, T. J.; Yoshimura, A.; Flo, T. H.; Rawadi, G.; Finberg, R. W.; Carroll, J. D.; Espevik, T.; Ingalls, R. R. (1999-11-19). "Toll-like receptor 2 functions as a pattern recognition receptor for diverse bacterial products". The Journal of Biological Chemistry. 274 (47): 33419–33425. ISSN 0021-9258. PMID 10559223.
  13. ^ Akira, Shizuo; Takeda, Kiyoshi. "Toll-like receptor signalling". Nature Reviews Immunology. 4 (7): 499–511. doi:10.1038/nri1391.
  14. ^ Doyle SL, O'Neill LA (October 2006). "Toll-like receptors: from the discovery of NFkappaB to new insights into transcriptional regulations in innate immunity". Biochem. Pharmacol. 72 (9): 1102–1113. doi:10.1016/j.bcp.2006.07.010. PMID 16930560.
  15. ^ Nat Rev Immunol. 2009 Jul;9(7):465-79. Signalling through C-type lectin receptors: shaping immune responses. Geijtenbeek TB, Gringhuis SI. http://www.mh-hannover.de/fileadmin/mhh/bilder/international/hbrs_mdphd/ZIB/Vorlesungen/Paper_09-10/Rev_IM_Geijtenbeek.pdf
  16. ^ a b Hoving, J. Claire; Wilson, Gillian J.; Brown, Gordon D. (2014-02-01). "Signalling C-Type lectin receptors, microbial recognition and immunity". Cellular Microbiology. 16 (2): 185–194. doi:10.1111/cmi.12249. ISSN 1462-5822. PMC 4016756. PMID 24330199.{{cite journal}}: CS1 maint: PMC format (link)
  17. ^ Hardison, Sarah E; Brown, Gordon D. "C-type lectin receptors orchestrate antifungal immunity". Nature Immunology. 13 (9): 817–822. doi:10.1038/ni.2369. PMC 3432564. PMID 22910394.{{cite journal}}: CS1 maint: PMC format (link)
  18. ^ Cummings RD, McEver RP. C-type Lectins. In: Varki A, Cummings RD, Esko JD, et al., editors. Essentials of Glycobiology. 2nd edition. Cold Spring Harbor (NY): Cold Spring Harbor Laboratory Press; 2009. http://www.ncbi.nlm.nih.gov/books/NBK1943/
  19. ^ a b Geijtenbeek, Teunis B. H.; Gringhuis, Sonja I. "Signalling through C-type lectin receptors: shaping immune responses". Nature Reviews Immunology. 9 (7): 465–479. doi:10.1038/nri2569.
  20. ^ Hoving, J. Claire; Wilson, Gillian J.; Brown, Gordon D. (2014-02-01). "Signalling C-Type lectin receptors, microbial recognition and immunity". Cellular Microbiology. 16 (2): 185–194. doi:10.1111/cmi.12249. ISSN 1462-5822. PMC 4016756. PMID 24330199.{{cite journal}}: CS1 maint: PMC format (link)
  21. ^ Zelensky, Alex N; Gready, Jill E (2005-12-01). "The C-type lectin-like domain superfamily". FEBS Journal. 272 (24): 6179–6217. doi:10.1111/j.1742-4658.2005.05031.x. ISSN 1742-4658.
  22. ^ East, L. "The mannose receptor family". Biochimica et Biophysica Acta (BBA) - General Subjects. 1572 (2–3): 364–386. doi:10.1016/s0304-4165(02)00319-7.
  23. ^ Apostolopoulos V, McKenzie IF (September 2001). "Role of the mannose receptor in the immune response". Curr. Mol. Med. 1 (4): 469–74. doi:10.2174/1566524013363645. PMID 11899091.
  24. ^ Vukman, Krisztina V.; Ravidà, Alessandra; Aldridge, Allison M.; O'Neill, Sandra M. (2013-09-01). "Mannose receptor and macrophage galactose-type lectin are involved in Bordetella pertussis mast cell interaction". Journal of Leukocyte Biology. 94 (3): 439–448. doi:10.1189/jlb.0313130. ISSN 0741-5400. PMID 23794711.
  25. ^ a b c d e Caruso, Roberta; Warner, Neil; Inohara, Naohiro; Núñez, Gabriel. "NOD1 and NOD2: Signaling, Host Defense, and Inflammatory Disease". Immunity. 41 (6): 898–908. doi:10.1016/j.immuni.2014.12.010. PMC 4272446. PMID 25526305.{{cite journal}}: CS1 maint: PMC format (link)
  26. ^ a b Ting JP, Williams KL (April 2005). "The CATERPILLER family: an ancient family of immune/apoptotic proteins". Clin. Immunol. 115 (1): 33–37. doi:10.1016/j.clim.2005.02.007. PMID 15870018.
  27. ^ Inohara, Inohara, McDonald C, Nuñez G (2005). "NOD-LRR proteins: role in host-microbial interactions and inflammatory disease". Annu. Rev. Biochem. 74: 355–383. doi:10.1146/annurev.biochem.74.082803.133347. PMID 15952891.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  28. ^ Jones DG, Dangl JL (2006). "The plant immune system". Nature. 444 (7117): 323–329. Bibcode:2006Natur.444..323J. doi:10.1038/nature05286. PMID 17108957.
  29. ^ Strober W, Murray PJ, Kitani A, Watanabe T (January 2006). "Signalling pathways and molecular interactions of NOD1 and NOD2". Nat. Rev. Immunol. 6 (1): 9–20. doi:10.1038/nri1747. PMID 16493424.
  30. ^ Burberry, Aaron; Zeng, Melody Y.; Ding, Lei; Wicks, Ian; Inohara, Naohiro; Morrison, Sean J.; Núñez, Gabriel. "Infection Mobilizes Hematopoietic Stem Cells through Cooperative NOD-like Receptor and Toll-like Receptor Signaling". Cell Host & Microbe. 15 (6): 779–791. doi:10.1016/j.chom.2014.05.004. PMC 4085166. PMID 24882704.{{cite journal}}: CS1 maint: PMC format (link)
  31. ^ Kim, Yun-Gi; Park, Jong-Hwan; Shaw, Michael H.; Franchi, Luigi; Inohara, Naohiro; Núñez, Gabriel. "The Cytosolic Sensors Nod1 and Nod2 Are Critical for Bacterial Recognition and Host Defense after Exposure to Toll-like Receptor Ligands". Immunity. 28 (2): 246–257. doi:10.1016/j.immuni.2007.12.012.
  32. ^ Ip, W. K. Eddie; Medzhitov, Ruslan (2015-05-11). "Macrophages monitor tissue osmolarity and induce inflammatory response through NLRP3 and NLRC4 inflammasome activation". Nature Communications. 6: ncomms7931. doi:10.1038/ncomms7931. PMC 4430126. PMID 25959047.{{cite journal}}: CS1 maint: PMC format (link)
  33. ^ Ip, W. K. Eddie; Medzhitov, Ruslan (2015-05-11). "Macrophages monitor tissue osmolarity and induce inflammatory response through NLRP3 and NLRC4 inflammasome activation". Nature Communications. 6: ncomms7931. doi:10.1038/ncomms7931. PMC 4430126. PMID 25959047.{{cite journal}}: CS1 maint: PMC format (link)
  34. ^ Guo, Haitao; Callaway, Justin B; Ting, Jenny P-Y. "Inflammasomes: mechanism of action, role in disease, and therapeutics". Nature Medicine. 21 (7): 677–687. doi:10.1038/nm.3893. PMC 4519035. PMID 26121197.{{cite journal}}: CS1 maint: PMC format (link)
  35. ^ Pattabhi, Sowmya; Wilkins, Courtney R.; Dong, Ran; Knoll, Megan L.; Posakony, Jeffrey; Kaiser, Shari; Mire, Chad E.; Wang, Myra L.; Ireton, Renee C. (2016-03-01). "Targeting Innate Immunity for Antiviral Therapy through Small Molecule Agonists of the RLR Pathway". Journal of Virology. 90 (5): 2372–2387. doi:10.1128/jvi.02202-15. ISSN 0022-538X. PMID 26676770.
  36. ^ Eisenächer, Katharina; Krug, Anne. "Regulation of RLR-mediated innate immune signaling – It is all about keeping the balance". European Journal of Cell Biology. 91 (1): 36–47. doi:10.1016/j.ejcb.2011.01.011.
  37. ^ Satoh, Takashi; Kato, Hiroki; Kumagai, Yutaro; Yoneyama, Mitsutoshi; Sato, Shintaro; Matsushita, Kazufumi; Tsujimura, Tohru; Fujita, Takashi; Akira, Shizuo (2010-01-26). "LGP2 is a positive regulator of RIG-I– and MDA5-mediated antiviral responses". Proceedings of the National Academy of Sciences. 107 (4): 1512–1517. doi:10.1073/pnas.0912986107. ISSN 0027-8424. PMC 2824407. PMID 20080593.{{cite journal}}: CS1 maint: PMC format (link)
  38. ^ Loo, Yueh-Ming; Gale, Michael. "Immune Signaling by RIG-I-like Receptors". Immunity. 34 (5): 680–692. doi:10.1016/j.immuni.2011.05.003. PMC 3177755. PMID 21616437.{{cite journal}}: CS1 maint: PMC format (link)
  39. ^ Bahar O, Pruitt R, Luu DD, Schwessinger B, Daudi A, Liu F, Ruan R, Fontaine-Bodin L, Koebnik R, Ronald P (2014). "The Xanthomonas Ax21 protein is processed by the general secretory system and is secreted in association with outer membrane vesicles". PeerJ. 2: e242. doi:10.7717/peerj.242. PMC 3897388. PMID 24482761.{{cite journal}}: CS1 maint: unflagged free DOI (link)
  40. ^ a b Boller T, Felix G (2009). "A renaissance of elicitors: perception of microbe-associated molecular patterns and danger signals by pattern-recognition receptors". Annu Rev Plant Biol. 60: 379–406. doi:10.1146/annurev.arplant.57.032905.105346. PMID 19400727.
  41. ^ Chisholm ST, Coaker G, Day B, Staskawicz BJ (2006). "Host-microbe interactions: shaping the evolution of the plant immune response". Cell. 124 (4): 803–814. doi:10.1016/j.cell.2006.02.008. PMID 16497589.
  42. ^ McHale L, Tan X, Koehl P, Michelmore RW (2006). "Plant NBS-LRR proteins: adaptable guards". Genome Biol. 7 (4): 212. doi:10.1186/gb-2006-7-4-212. PMC 1557992. PMID 16677430.{{cite journal}}: CS1 maint: unflagged free DOI (link)
  43. ^ Dardick, Chris; Schwessinger, Benjamin; Ronald, Pamela. "Non-arginine-aspartate (non-RD) kinases are associated with innate immune receptors that recognize conserved microbial signatures". Current Opinion in Plant Biology. 15 (4): 358–366. doi:10.1016/j.pbi.2012.05.002.
  44. ^ Dardick C, Ronald P (January 2006). "Plant and animal pathogen recognition receptors signal through non-RD kinases". PLoS Pathog. 2 (1): e2. doi:10.1371/journal.ppat.0020002. PMC 1331981. PMID 16424920.{{cite journal}}: CS1 maint: unflagged free DOI (link)
  45. ^ Wang GL, Ruan DL, Song WY, Sideris S, Chen L, Pi LY, Zhang S, Zhang Z, Fauquet C, Gaut BS, Whalen MC, Ronald PC (May 1998). "Xa21D encodes a receptor-like molecule with a leucine-rich repeat domain that determines race-specific recognition and is subject to adaptive evolution". Plant Cell. 10 (5): 765–79. doi:10.2307/3870663. PMC 144027. PMID 9596635.
  46. ^ Dommett RM, Klein N, Turner MW (September 2006). "Mannose-binding lectin in innate immunity: past, present and future". Tissue Antigens. 68 (3): 193–209. doi:10.1111/j.1399-0039.2006.00649.x. PMID 16948640.
  47. ^ "Understanding Immunotherapy". Cancer.Net. 2013-03-25. Retrieved 2017-07-31.
  48. ^ Chen, Edward S. "Innate immunity in sarcoidosis pathobiology". Current Opinion in Pulmonary Medicine. 22 (5): 469–475. doi:10.1097/mcp.0000000000000305.
  49. ^ Philpott, Dana J.; Sorbara, Matthew T.; Robertson, Susan J.; Croitoru, Kenneth; Girardin, Stephen E. "NOD proteins: regulators of inflammation in health and disease". Nature Reviews Immunology. 14 (1): 9–23. doi:10.1038/nri3565.
  50. ^ Jun, Janice C.; Cominelli, Fabio; Abbott, Derek W. (2013-11-01). "RIP2 activity in inflammatory disease and implications for novel therapeutics". Journal of Leukocyte Biology. 94 (5): 927–932. doi:10.1189/jlb.0213109. ISSN 0741-5400. PMC 3800061. PMID 23794710.{{cite journal}}: CS1 maint: PMC format (link)
  51. ^ Tigno-Aranjuez, Justine T.; Benderitter, Pascal; Rombouts, Frederik; Deroose, Frederik; Bai, XiaoDong; Mattioli, Benedetta; Cominelli, Fabio; Pizarro, Theresa T.; Hoflack, Jan (2014-10-24). "In Vivo Inhibition of RIPK2 Kinase Alleviates Inflammatory Disease". Journal of Biological Chemistry. 289 (43): 29651–29664. doi:10.1074/jbc.m114.591388. ISSN 0021-9258. PMID 25213858.{{cite journal}}: CS1 maint: unflagged free DOI (link)
  52. ^ Haile, Pamela A.; Votta, Bartholomew J.; Marquis, Robert W.; Bury, Michael J.; Mehlmann, John F.; Singhaus, Robert; Charnley, Adam K.; Lakdawala, Ami S.; Convery, Máire A. (2016-05-26). "The Identification and Pharmacological Characterization of 6-(tert-Butylsulfonyl)-N-(5-fluoro-1H-indazol-3-yl)quinolin-4-amine (GSK583), a Highly Potent and Selective Inhibitor of RIP2 Kinase". Journal of Medicinal Chemistry. 59 (10): 4867–4880. doi:10.1021/acs.jmedchem.6b00211. ISSN 0022-2623.
  53. ^ Corridoni, D; Rodriguez-Palacios, A; Stefano, G Di; Martino, L Di; Antonopoulos, D A; Chang, E B; Arseneau, K O; Pizarro, T T; Cominelli, F. "Genetic deletion of the bacterial sensor NOD2 improves murine Crohn's disease-like ileitis independent of functional dysbiosis". Mucosal Immunology. 10 (4): 971–982. doi:10.1038/mi.2016.98. PMC 5433921. PMID 27848951.{{cite journal}}: CS1 maint: PMC format (link)
  54. ^ Canning, Peter; Ruan, Qui; Schwerd, Tobias; Hrdinka, Matous; Maki, Jenny L.; Saleh, Danish; Suebsuwong, Chalada; Ray, Soumya; Brennan, Paul E. "Inflammatory Signaling by NOD-RIPK2 Is Inhibited by Clinically Relevant Type II Kinase Inhibitors". Chemistry & Biology. 22 (9): 1174–1184. doi:10.1016/j.chembiol.2015.07.017. PMC 4579271. PMID 26320862.{{cite journal}}: CS1 maint: PMC format (link)
  55. ^ Castaño-Rodríguez, Natalia; Kaakoush, Nadeem O.; Mitchell, Hazel M. (2014). "Pattern-Recognition Receptors and Gastric Cancer". Frontiers in Immunology. 5. doi:10.3389/fimmu.2014.00336. ISSN 1664-3224. PMC 4105827. PMID 25101079.{{cite journal}}: CS1 maint: PMC format (link) CS1 maint: unflagged free DOI (link)
  56. ^ Kigerl, Kristina A.; Vaccari, Juan Pablo de Rivero; Dietrich, W. Dalton; Popovich, Phillip G.; Keane, Robert W. "Pattern recognition receptors and central nervous system repair". Experimental Neurology. 258: 5–16. doi:10.1016/j.expneurol.2014.01.001. PMC 4974939. PMID 25017883.{{cite journal}}: CS1 maint: PMC format (link)
  57. ^ Kigerl, Kristina A.; Vaccari, Juan Pablo de Rivero; Dietrich, W. Dalton; Popovich, Phillip G.; Keane, Robert W. "Pattern recognition receptors and central nervous system repair". Experimental Neurology. 258: 5–16. doi:10.1016/j.expneurol.2014.01.001. PMC 4974939. PMID 25017883.{{cite journal}}: CS1 maint: PMC format (link)

External links