Tlr and Pamp

download Tlr and Pamp

of 7

Transcript of Tlr and Pamp

  • 8/21/2019 Tlr and Pamp

    1/15

    Dong Hyun Song

     Jie-Oh LeeSensing of microbial molecularpatterns by Toll-like receptors

    Authors’ addresses

    Dong Hyun Song1, Jie-Oh Lee1,2

    1Department of Chemistry, KAIST, Daejon, Korea.2Graduate School of Nanoscience & Technology (WCU),

    KAIST, Daejon, Korea.

    Correspondence to:

     Jie-Oh Lee

    Department of Chemistry, Graduate School of Nanoscience

    & Technology (WCU), KAIST

    Daejon, Korea

    Tel.: +82 42 350 2839

    Fax: +82 42 350 2810

    e-mail: [email protected]

    Acknowledgement

    We gratefully acknowledge the support of the World Class

    University Program (R31 – 10071) of the Ministry of 

    Education, Science and Technology of Korea. The authors

    have no conflicts of interest to declare.

    This article is part of a series of reviews

    covering Insights from Structure appearing

    in Volume 250 of  Immunological Reviews.

    Summary:   Toll-like receptors (TLRs) sense structural patterns in micro-bial molecules and initiate immune defense mechanisms. The structuresof many extracellular and intracellular domains of TLRs have beenstudied in the last 10 years. These structures reveal the extraordinarydiversity of TLR-ligand interactions. Some TLRs use internal hydropho-bic pockets to bind bacterial ligands and others use solvent-exposed

    surfaces to bind hydrophilic ligands. The structures suggest a commonactivation mechanism for TLRs: ligand binding to extracellular domainsinduces dimerization of the intracellular domains and so activates intra-cellular signaling pathways. Recently, the structure of the death domaincomplex of one of the signaling adapters, myeloid differentiation factor88 (MyD88), has been determined. This structure shows how aggrega-tion of signaling adapters recruits downstream kinases. However, weare still far from a complete understanding of TLR activation. We needto study the structures of TLR7 – 10 in complex with their ligands. Wealso need to determine the structures of TLR-adapter aggregates tounderstand activation mechanisms and the specificity of the signalingpathways. Ultimately, we will have to study the structures of the com-plete TLR signaling complexes containing full-length receptors, ligands,signaling, and bridging adapters, and some of the downstream kinases

    to understand how TLRs sense microbial infections and activateimmune responses against them.

    Keywords:   Toll-like receptor, innate immunity, pattern recognition, MyD88

    Toll-like receptors

    Toll-like receptors (TLRs) provide efficient and immediate

    immune responses to bacterial, fungal, and viral infections

    by recognizing diverse molecules released from them.

    Because they recognize common patterns in microbial mole-

    cules, they are often referred to as ‘pattern recognitionreceptors’ (1, 2). Humans have 10 TLRs, and each of them

    binds a distinct family of microbial molecules (3, 4). TLR2

    forms a heterodimer with either TLR1 or TLR6 and binds

    bacterial lipoproteins. TLR4 in complex with MD-2 recog-

    nizes the lipopolysaccharide (LPS) that is a major compo-

    nent of the outer membrane of Gram-negative bacteria.

    TLR5 can be stimulated by a depolymerized form of flagel-

    lin. TLR3 and TLR7 – 9 respond to various forms of bacterial

    or viral nucleic acid. The engagement of TLRs with ligand

    Immunological Reviews 2012

    Vol. 250: 216–229

    Printed in Singapore. All rights reserved 

    © 2012 John Wiley & Sons A/SImmunological Reviews0105-2896

    ©

     2012 John Wiley & Sons A/S216   Immunological Reviews 250/2012

  • 8/21/2019 Tlr and Pamp

    2/15

    molecules induces dimerization of the receptors, which

    leads to recruitment of intracellular signaling adapters and

    kinases. The activation of TLR signaling pathways eventually

    results in the synthesis of pro-inflammatory cytokines and

    therefore prepares the entire immune system for infection.

    TLRs are important targets for drug design because excess

    activity of TLRs is involved in various inflammatory diseases(5). Antagonists of TLR4 are being developed as targets of 

    anti-sepsis drugs (6 – 9); agonists of TLR7 and TLR8 are

    being used as anti-viral and anti-cancer drugs, and many

    more are in clinical trials (5). A partial agonist of TLR4 has

    recently been approved as a general vaccine adjuvant, and

    agonists targeted to several TLRs are being developed for

    this purpose (10 – 12).

    Overview of the TLR activation mechanism

    Like other single transmembrane receptors, TLRs are acti-vated by ligand-induced multimerization (Fig. 1). The struc-

    tures of the key intermediates in these activation processes

    have been mainly studied by x-ray crystallography (13 – 21).

    The extracellular domain of each TLR has the characteristic

    horseshoe-like structure of the leucine-rich repeat (LRR)

    family of proteins (13, 14, 22, 23). The concave surface

    consists of a large central  b  sheet, and the convex surface is

    decorated by a combination of helices and irregular loops

    (Fig. 2). Binding of agonistic ligands to TLRs induces dimer-

    ization of their extracellular domains. We now have struc-

    tures for the extracellular domains of six of the 10 humanTLRs in the ligand-induced dimeric state. Although the

    ligand recognition mechanisms of the TLRs vary substan-

    tially, all the TLR dimers have a similar overall arrangement:

    the C-termini of the two extracellular domains converge in

    the middle, separated by distances ranging from 20 to 40

    angstrom. This structural observation has led to the proposal

    that ligand binding to the extracellular domains enhances

    dimerization of the intracellular domains because the latter

    are connected to the C-termini of the extracellular domains

    by transmembrane helices (15). The intracellular domains of 

    TLRs, as well as of several signaling adapters, share a

    domain named the Toll/interleukin-1 receptor (TIR)

    domain. Its structure is composed of a parallel  b  sheet sur-

    rounded by   a  helices (24 – 29). Ligand-induced aggregation

    of the receptors and adapters is mediated by homotypic and

    heterotypic interaction between these TIR domains.

    Although the structures of quite a few monomeric TIRs have

    already been reported, the structures of their homo- andheteromultimers remain unsolved.

    All these structural studies have been conducted with

    truncated extracellular or intracellular domains because the

    full-length receptors including the transmembrane helices

    are difficult to produce and crystallize. Therefore, the struc-

    tural information of the full-length TLRs under membrane

    embedded condition is not available. There is evidence that

    LPS

    Binding

    TLR4-MD-2

    Dimerization

     Adaptor-Kinase

    recruitment

    Receptor 

    clustering (?)

    TLR signaling complex

    MyD88 IRAK4 IRAK2

    MALTLR_TIR

    LPS

    Fig. 1.  Overall mechanism of TLR4 activation.  The plasma membrane is shown schematically by pink horizontal bars. The TLR4 extracellulardomains are shown as black arcs and the MD-2s as green balls.

    LRRCT

    LRRNT

    Convex

    Concave

    Lxx

    Lxxx

      L

    N

    N

    C

    Hydrophobiccore

    Intracellular 

    TIR domain

    Extracellular 

    LRR domain

       A  s  c  e  n   d

       i  n  g   l  a   t  e  r  a   l

       D  e  s  c  e  n   d   i  n  g   l  a   t  e  r  a   l

    Transmembranehelix

    Fig. 2.  Overall structure of a Toll-like receptor (TLR).  Theextracellular and intracellular domains of TLR4 are drawn as a bluesurface. A Ca trace of one of the LRR modules and the side chains of the conserved leucines and asparagine is shown. The LRRNT andLRRCT modules are colored pink and green, respectively.

    ©

     2012 John Wiley & Sons A/SImmunological Reviews 250/2012   217

    Song & Lee     Structure of Toll-like receptors

  • 8/21/2019 Tlr and Pamp

    3/15

    activated and dimerized TLRs form higher order clusters in

    the cell membrane (25, 30). However, structures of these

    proposed clusters have never been characterized in detail

    because of difficulties reconstituting the complete receptor

    signaling complexes that includes the full-length receptors,

    ligands, signaling, and bridging adapters and some of the

    downstream kinases. This is a common problem for allreceptors with single transmembrane helices, and future

    research in receptor biology should address this fundamental

    problem.

    Structures of the monomeric TLRs without bound

    ligands

    The extracellular domains

    The available x-ray structures of the extracellular domains of 

    TLRs in the monomeric state reveal the conformations of 

    inactive TLRs prior to binding ligand (14, 17, 31). In addi-tion to this, several structures of TLR2 and TLR4 bound to

    non-agonistic or antagonistic ligands have been determined

    (16, 17). The extracellular domains of TLRs belong to the

    LRR family. The amino acid sequences of LRR family pro-

    teins are composed of multiple copies of repeat modules

    named LRR modules (22, 23) (Fig. 2). Each module is 20 – 

    30 amino acids long and has a characteristic LxxLxLxxN

    sequence motif. The central LxL part of the module forms

    the core of a  b   strand. The two leucines point towards the

    interior of the protein, making up the hydrophobic core,

    whereas the variable   9  residues within the motif areexposed to solvent, and some are involved in interactions

    with ligands. The conserved asparagines in the motifs form

    a continuous hydrogen bonding network with backbone

    carbonyl oxygens in neighboring LRR modules throughout

    the protein. The overall structures of the LRR proteins

    resemble a horseshoe. The   b   strands provided by the each

    LRR module assemble into a large   b   sheet making up the

    entire concave surface of the horseshoe. Variable amino

    acids outside the conserved LRR motif of each module gen-

    erate the convex part of the structure. The LRR modules are

    protected by two special modules named LRRNT and LRRCT

    in the N- and C-termini of the proteins. These modules do

    not conform to the sequence conservation pattern of the

    LRR modules and often contain an anti-parallel   b   hairpin

    stabilized by disulfide bridges. The majority of the LRR

    family proteins are involved in protein – protein interactions,

    and their protein-binding sites are located in their concave

    surfaces. TLRs are atypical members of the LRR family: with

    the exception of TLR5, their ligands are not proteins and

    their ligand-binding sites are not located in their concave

    surfaces (4).

    The LRR family is divided into seven subfamilies, each of 

    which has a different module length and sequence conserva-

    tion pattern in the convex part of the structure (22, 23).

    Some of them have parallel   a   helices, and others have 310

    helices, polyproline II helices, or irregular loops. Theirsequence conservation patterns suggest that the TLRs belong

    to the ‘typical’ subfamily (17, 32, 33). However, they devi-

    ate substantially from the canonical pattern of the subfamily.

    The LRR modules of the typical subfamily contain 24 amino

    acids and have 310  helices in the convex area. On the other

    hand, the lengths of TLR modules vary between 20 and 28

    amino acids, and the convex parts of their structures are

    composed of mixtures of  a  helices, 310 helices, and irregular

    loops. Because of this, the convex surface of a TLR is rough

    rather than smooth and contains multiple grooves and shal-

    low pockets that can serve for ligand interaction. Further-

    more, the conserved asparagines are missing in some of the

    LRR modules in lipid-interacting TLRs, TLR1, TLR2, TLR4,

    and TLR6; this allows unusual distortions of the horseshoe-

    like structures, giving rise to sharp transitions in the torsion

    and twist angles of the central   b   sheet (4, 15, 17). Such

    structural distortion permits the formation of hydrophobic

    pockets in TLR1 and TLR2 and also has a role in the interac-

    tion of TLR4 with MD-2.

    Structures of the intracellular domainsIntracellular TIR domains are found not only in the TLRs

    but also in the adapters involved in TLR signaling pathways

    (34, 35). Five signaling and bridging adapters, MyD88,

    MyD88-adapter-like protein (MAL), TIR-domain-containing

    adapter-inducing interferon-b   (TRIF), translocating chain-

    associating membrane protein (TRAM), and sterile-a   and

    Armadillo motif-containing protein (SARM), have been

    found to be important in human TLR responses. MAL, TRIF,

    and TRAM are also called TIR-domain-containing adapter

    protein (TIRAP), TIR-containing adapter molecule-1 (TI-

    CAM-1), and TICAM-2, respectively. All TIR domains have a

    similar overall structure (Fig. 3), with a central five-stranded

    parallel   b-sheet surrounded by five   a   helices (24, 26 – 29).

    Sequence conservation in the TIR family is relatively low, in

    the 20 – 30% range, and domain lengths vary between 135

    and 160 residues. This is because of large deletions and

    insertions in the loop regions.

    The structures of monomeric TIR domains have been

    determined in the cases of TLR1, TLR2, and TLR10 (26,

    ©

     2012 John Wiley & Sons A/S218   Immunological Reviews 250/2012

    Song & Lee     Structure of Toll-like receptors

  • 8/21/2019 Tlr and Pamp

    4/15

    28). They share substantial structural homology with Ca

    rms differences of less than 2 Å. Mutagenesis studies have

    identified several regions that are important for signaling

    activity. The most prominent region identified is the BB

    loop that connects strand  bB and helix  aB (26, 28, 36 – 39).

    The BB loop extends away from the rest of the TIR domain,

    forming a protrusion from the surface. A P712H substitu-

    tion in TLR4 abolishes the host response to LPS. The corre-sponding residue in TLR2 is located in the BB loop and

    does not have any clear structural role, suggesting that it is

    functionally but not structurally important. The TLR10 TIR

    domain has been crystallized as a homodimer with an

    extensive dimer interface, although it exists as a monomer

    in solution (26). The dimerization interface is symmetrically

    shaped, involves the BB loop, and shows substantial

    sequence conservation. Because of this, the dimeric arrange-

    ment of TLR10 TIR has been used as the template for TIR

    homodimerization in several modeling studies (38). Some

    mutations in the DD loop that connects strand bD with helixaD also have significant effects on TLR activation, suggesting

    that the DD loop is part of the TIR – TIR interaction interface

    along with the BB loop (26, 28, 36 – 39).

    The structures of the TIR domains of the signaling adapt-

    ers MyD88 and MAL have also been reported (27, 29).

    Human MyD88 consists of 296 amino acids and is com-

    posed of two signaling domains. The C-terminal TIR

    domain mediates interaction with the receptor, and the N-

    terminal death domain is responsible for recruiting the

    IRAK4 and IRAK1/2 kinases that initiate downstream signal-

    ing. The structure of the MyD88 TIR domain has been

    determined by nuclear magnetic resonance (NMR) spectros-

    copy (27). Its overall structure is similar to those of the TLR

    TIR domains. The largest structural difference is in the BB

    and DD loop regions; the   aB helix becomes much shorter

    because some of the N-terminal residues of the helix are

    unwound to be part of the BB loop. The   aD helix is com-pletely changed to an irregularly shaped loop. Using a

    NMR-based method, they have performed binding studies

    between the MyD88 and MAL TIR domains. They found that

    these two TIR domains interact weakly, with a Kd value of 

    approximately 7 micromoles. Alanine substitution of R196

    of MyD88 reduced MAL binding twofold, and mutations of 

    the residues R288 and R217 also substantially reduced MAL

    binding. These observations suggest that the continuous

    interface containing the BB loop and these arginine residues

    mediates the TIR – TIR interaction with MAL.

    MAL is required for MyD88-dependent signaling by TLR2and TLR4 (34, 35). Other TLRs do not need MAL for

    MyD88-dependent signaling. Human MAL has 221 amino

    acid residues that make up a phosphatidyl inositol-binding

    motif and a TIR domain. Although the TIR domain has a

    similar overall fold composed of a central   b   sheet sur-

    rounded by   a   helices, its structure is rather different from

    other TLR TIR domains (29). Its sequence identity is in the

    low twenties and the Ca   rms deviation is greater than 2 Å.

    Furthermore, the  aB and  aD helices are changed to irregular

    Fig. 3.  TIR domain structures. The PdTIR, AtTIR, and L6 TIR domains are from the proteins of  Paracoccus denitrificans, Arabidopsis thaliana, and Linumusitatissimum, respectively. The figure was drawn using coordinate files with PDB codes, 1FYV (TLR1), 1FYW (TLR2), 2J67 (TLR10), 2Z5V

    (MyD88), 2Y92 (MAL), 1T3G (IL-1R), 3H16 (PdTIR), 3JRN (AtTIR), and 3OZI (L6). The  a  helices are green and the central  b  sheet is blue. TheBB and DD loops important in TLR signaling are shown in red. The AtTIR and L6 TIR domains have additional helices in the  aD region.

    ©

     2012 John Wiley & Sons A/SImmunological Reviews 250/2012   219

    Song & Lee     Structure of Toll-like receptors

  • 8/21/2019 Tlr and Pamp

    5/15

    loops. Cysteines play important roles in the structure and

    function of MAL, forming two disulfide bridges, C89 – C134

    and C142 – C174. Disulfide bridges are rare in intracellular

    proteins, although not unprecedented. The two other cyste-

    ines, C91 and C157, are not involved in forming disulfide

    bridges with protein residues. However, they appear to be

    chemically reactive and form simultaneous disulfide bridgeswith a single DTT molecule in the buffer. Mutagenesis

    experiments have shown that these cysteines are crucial for

    MyD88 binding. Interestingly, the C157 position is highly

    conserved in the TLR family and the corresponding residue

    in TLR4 is the target site of TAK-242 (40). This is an exper-

    imental drug that blocks TLR4 signaling by inhibiting its

    interaction with MAL and TRAM (9).

    TIR domains are found not only in the mammalian pro-

    teins but also in several bacterial and plant proteins. The

    bacterial TIR-containing proteins are thought to interfere

    with the host immune system by hijacking MyD88 through

    direct TIR – TIR interaction (41 – 44). Plants contain a large

    number of related proteins which are involved in resistance

    to infection (45, 46). These proteins contain single TIR

    domains that are thought to be involved in multimerization.

    The structures of the bacterial TIR domain of   Paracoccus deni-

    trificans  and the plant TIR domains of   Linum usitatissimum   and

     Arabidopsis thaliana  have been determined (36, 47, 48). These

    non-animal TIR domains have the similar overall fold

    expected from the sequence conservation, although the two

    plant TIRs have additional helices in the   aD helix region

    (Fig. 3).

    Dimerization and activation of TLRs

    Agonistic ligands induce homo- or heterodimerization of 

    TLR extracellular domains. The structures of these TLR-

    ligand complexes have been extensively studied by x-raycrystallography (15, 16, 18, 20, 21) (Fig. 4). Their interac-

    tions with ligands are extraordinarily diverse (Fig. 5A). TLR2

    forms heterodimers with TLR1 and TLR6 after binding lipo-

    peptides (15, 16). The lipid chains of the ligands are

    inserted into hydrophobic chambers within TLR2 and TLR1.

    For binding to the lipid chains of LPS, TLR4 use a hydro-

    phobic chamber not present in the TLR itself but in the

    accessory protein MD-2 (17, 19). MD-2 interacts with the

    N-terminal and central domains of TLR4, and LPS induces

    dimerization of the TLR4-MD-2 heterodimers, forming het-

    erotetrameric complexes of two TLRs and two MD-2s via asecond TLR4 interaction site in MD-2 (20). Hydrophilic

    interactions play major roles in ligand recognition by TLR3

    and TLR5. TLR3 has two ligand-binding sites located near

    the N-terminus and C-terminus, respectively, of its extracel-

    lular domain, and flagellin binds to a concave surface

    formed by the N-terminal nine LRR modules of TLR5 (18,

    21). In each of these cases, the ligands play direct roles in

    TLR dimerization by interacting simultaneously with two

    TLRs. All the resulting TLR structures have strikingly similar

    TLR2 TLR1 TLR2 TLR6   TLR3

    TLR4   TLR5

    Triacyl l ipopeptide Diacyl lipopeptide dsRNA

    FlagellinLPS

    Fig. 4.  Structures of Toll-like receptor (TLR)-ligand complexes.  The transmembrane helices and intracellular domains are not included in thecrystal structures. Synthetic lipopeptides that contain the di- or triacylated cysteine and several peptide residues were used in the TLR2-TLR1 andTLR2-TLR6 crystallographic studies. The 11 C-terminal modules of TLR5 are deleted in the crystal structure.

    ©

     2012 John Wiley & Sons A/S220   Immunological Reviews 250/2012

    Song & Lee     Structure of Toll-like receptors

  • 8/21/2019 Tlr and Pamp

    6/15

    shapes resembling the letter ‘m’. In these dimeric structures,

    the two C-termini of the extracellular domains converge in

    the center while the N-termini stretch outwards (15, 16,

    18, 20, 21). This finding suggests the hypothesis that

    dimerization of the extracellular domains leads to juxtaposi-

    tion of the intracellular domains, and the signaling adapters

    are believed to be recruited into such dimerized forms of 

    the TLR intracellular domains (15). The structural changes

    induced by ligands are minimal, and only minor adaptations

    of flexible loops or amino acid side chains are involved in

    ligand binding.

    Dimerization of TLR4-MD-2-LPS complexes

    The principal ligand of TLR4 is lipopolysaccharide (LPS),

    which is the main component of the outer membrane of 

    TLR2 TLR6

    Triacyl lipopeptide

    Diacyl lipopeptide

    TLR3

    dsRNA

    TLR4

    TLR5

    Flagellin

    D1

    D2

    D3D3

    D2

    D3D3

    MD-2 MD-2

    TLR2 TLR1LRRNT

    LRRCT

    N-terminal

    Central

       C  -   t  e  r  m   i  n  a   l   Dimerization

    interfacePrimary

    interface

    NN

    CC

    Triacyl lipoprotein Diacyl lipoprotein

    Cys   Cys

    Protein   Protein

    NHO

    NH3+

    S

    O

    O

    O

    O

    NHO

    NH

    OS

    O

    O

    O

    O

    O specific

    chain

    LPS

    Core

    Lipid A

    O

    O

    O

    O

    P-O

    O-

    O

    O   O

    HO   HO

    P

    O

    O-

    O-

    OO

    OHN

    O

    HO

    O   NH

    O

    OO

    O

    O

    A

    B

    LPS

    Fig. 5.  Ligand induced homo- and heterodimerization of the extracellular domains of Toll-like receptors (TLRs).  (A) Diversity of the TLR-ligand interaction. The extracellular domains of TLRs are shown as arcs. The C-terminal fragment of TLR5 deleted in the crystal structure is shownby dashed lines. The D3 domain of flagellin is not visible in the electron density map, presumably because it is very flexible. The boundaries of the LRRNT and LRRCT modules are drawn with solid lines. Some TLRs, namely TLR1, TLR2, TLR4, and TLR6, have structural transitions thatdivide them into three subdomains, N-terminal, central, and C-terminal. The boundaries of these subdomains are shown by dashed lines. (B)

    Chemical structures of LPS and lipoproteins.

    ©

     2012 John Wiley & Sons A/SImmunological Reviews 250/2012   221

    Song & Lee     Structure of Toll-like receptors

  • 8/21/2019 Tlr and Pamp

    7/15

    Gram-negative bacteria (49 – 51). LPS is a macromolecular

    glycolipid composed of a hydrophobic lipid A region

    attached to a long and branched carbohydrate chain

    (Fig. 5B). The lipid A has two glucosamine sugars to which

    typically four to seven lipid chains are attached (52 – 54).

    Lipid A is negatively charged because two phosphate groups

    are attached to the glucosamine backbone. Significant struc-tural diversity has been reported in the structure of lipid A.

    For example, the number of lipid chains can vary, and the

    phosphate groups can be deleted, or modified by several

    other chemical groups (52, 54). The carbohydrate chain of 

    LPS can be divided into two areas, the core and the O-spe-

    cific chain (50). The relatively conserved core oligosaccha-

    ride chain contains multiple copies of the unusual sugars, 3-

    deoxy-D-manno-oct-2-ulosonic acid (Kdo) and L-glycero-D-

    mannoheptose (Hep). The core region is frequently modified

    by negatively charged chemical groups such as phosphates.

    The O-specific chain is composed of many copies of repeat-

    ing units containing galactoses, galactosamines, and other

    sugars. The structure and composition of the repeating units

    differ between bacterial species and even between different

    growth conditions in the same species. Complete removal of 

    the core and O-specific chain has only a minor effect on the

    immunological activity of LPS. Therefore, the lipid A region

    is believed responsible for most of its inflammatory activity.

    Crystallographic investigations have confirmed this and

    have shown that hydrophobic interaction of lipid A with a

    pocket in MD-2 makes the main contribution to LPS binding(20).

    LPS is recognized by TLR4-MD-2 heterodimers expressed

    on the surface of immune cells (55 – 57). MD-2 is approxi-

    mately 18 kDa glycoprotein that functions as the LPS-bind-

    ing subunit of the TLR4-MD-2 heterodimer (57 – 60). It has

    two anti-parallel   b  sheets sandwiched together, like immu-

    noglobulin fold proteins. But the disulfide bond connecting

    the two sheets in the latter is absent from MD-2 (17, 19).

    Because of this, the two   b   sheets in the sandwiched struc-

    ture of MD-2 can be separated to generate a large internal

    pocket that is ideally shaped for binding flat hydrophobicligands like LPS. MD-2 has two TLR4 binding sites, named

    the primary interface and dimerization interface, respectively

    (17, 20) (Fig. 5A). The primary interaction interface is

    formed between one edge of MD-2 and part of the concave

    surface provided by the N-terminal and central domains of 

    TLR4. This well-conserved interaction interface is mostly

    composed of ionic and hydrogen bonds and does not

    require binding of LPS. The dimerization interface of MD-2

    is located opposite the primary interface and interacts with a

    convex surface provided by a small hydrophobic patch in

    the C-terminal domain of TLR4. One partially exposed lipid

    chain of LPS is located in the core of the dimerization inter-

    face. This lipid chain and surrounding hydrophobic amino

    acid residues of MD-2 interact with the hydrophobic patch

    of TLR4. Several hydrophilic interactions between protein

    residues of MD-2 and TLR4 support this core hydrophobicinterface. The two phosphate groups of lipid A form multi-

    ple charge and hydrogen bond interactions with both MD-2

    and the two TLR4s and therefore contribute to the stability

    of the dimerization interface.

    The structure-activity relationships of LPS have been stud-

    ied for decades using natural and chemically modified LPS

    (50 – 52, 61). The results show that the total number of lipid

    chains and the presence of the two phosphate groups in

    lipid A are the most important factors affecting its inflam-

    matory activity. The crystal structure provides an explanation

    of this observation (20). The six lipid chains are optimal

    because partial exposure of one of the lipid chains is critical

    for forming the dimerization interface. When the LPS has

    less than six lipid chains, part of the MD-2 binding pocket

    becomes empty, and this will destabilize the overall interac-

    tion (62). When the LPS has more than six lipid chains, the

    additional lipid chain may interfere with the interaction

    between TLR4 and MD-2 because it is likely to prevent the

    approaching TLR4 from achieving the optimal interaction

    distance. The phosphate groups are essential because they

    provide the charge and hydrogen bonding network. Dele-tion of either of these phosphates reduces inflammatory

    activity more than a 100-fold (54, 61, 63). A lipid A deriv-

    ative with one of the phosphate groups deleted retains more

    or less, all of its immune stimulatory activity, but loses most

    of its inflammatory toxicity. Because of this, it has recently

    been approved as a general vaccine adjuvant (10).

    Formation of TLR2-TLR1 and TLR2-TLR6-lipoprotein

    complexes

    TLR2 forms heterodimers with TLR1 or TLR6 and interacts

    with bacterial lipoproteins (15, 16, 64, 65). Lipoproteinsare a family of proteins that are anchored to the bacterial

    membrane by lipid chains covalently attached to conserved

    N-terminal cysteines (66 – 70) (Fig. 5B). The protein parts of 

    the lipoproteins have little sequential and functional homol-

    ogy. The lipoproteins found in the Gram-negative bacteria

    have three lipid chains; two of them are connected to the

    glycerol group, which is in turn attached to the sulfur atom

    of the N-terminal cysteine. These lipid chains are referred to

    as ‘ester-bound’ lipid chains. The third lipid chain is

    ©

     2012 John Wiley & Sons A/S222   Immunological Reviews 250/2012

    Song & Lee     Structure of Toll-like receptors

  • 8/21/2019 Tlr and Pamp

    8/15

    connected to the N-terminal NH2 via an amide bond. There-

    fore, it is referred to as an ‘amide-bound’ lipid chain. The

    lipoproteins and lipopeptides of Gram-positive bacteria and

    mycoplasmas were thought to have only two lipid chains

    because they lack the enzyme responsible for attaching the

    amide-bound lipid chain (71 – 73). However, this view has

    recently been challenged because several lipoproteins puri-fied from   Staphylococcus aureus  were shown to have three lipid

    chains (74, 75). In addition to lipoproteins, many bacterial

    and fungal molecules have been proposed to function as

    agonistic ligands for TLR2, and many, but not all, of these

    candidate ligands also contain diacylglycerol groups (3, 4).

    TLR2-TLR1 heterodimers are the principal receptors for

    triacylated lipoproteins (65). The three lipid chains of the

    ligand bridge the TLRs by interacting simultaneously with

    TLR2 and TLR1 via the three lipid chains; two lipid chains

    are inserted into the large hydrophobic pocket in TLR2, and

    the third, the amide-bound chain, is inserted into the

    narrow hydrophobic channel in TLR1 (15) (Fig. 5A). The

    entry of the TLR2 pocket lies near the boundary between

    the central and C-terminal domains and extends into an

    internal hydrophobic pocket formed by LRR modules 9 – 12.

    The glycerol and the peptide backbone atoms of the two

    N-terminal residues of the lipoprotein form extensive

    hydrogen bonds with amino acids in both TLR2 and TLR1.

    Direct TLR1 – TLR2 interaction also contributes to the stabil-

    ity of the heterodimer. The core of the interface is formed

    by several hydrophobic residues. It is surrounded by hydro-philic residues forming ionic and hydrogen bonds. The

    Pro315 SNP of TLR1 frequently found in Caucasian popula-

    tions appears to block TLR1 signaling by perturbing this

    heterodimerization interface (15, 76).

    Diacylated lipoproteins from Gram-positive bacteria or

    mycoplasmas mainly activate TLR2-TLR6 heterodimers (64).

    The two ester-bound lipid chains are inserted into the same

    TLR2 pocket as TLR2-TLR1 triacyl lipopeptide complexes,

    and the hydrophilic glycerols and peptide backbones of the

    lipoproteins form hydrogen bonds with amino acid residues

    of both TLR2 and TLR6 (16) (Fig. 5A). The TLR2-TLR6complex cannot bind to the triacylated lipoproteins because

    the hydrophobic channel responsible for interaction with

    the amide-bound lipid chain in TLR1 is blocked by two

    bulky phenylalanines in TLR6. Mutation of these residues

    into the ones found in TLR1 renders TLR6 capable of 

    responding to both diacylated and triacylated lipoproteins.

    In the TLR2-TLR1-triacylate lipopeptide complex, the

    amide-bound lipid chain plays an important role in TLR

    heterodimer formation. Without the amide-bound lipid

    chain in the diacylated lipopeptide, TLR2 and TLR6 can

    still form stable heterodimers because the size of the hydro-

    phobic core of the protein – protein interface is increased,

    and this can compensate for the missing lipid channel

    interaction.

    Homodimerization of the TLR3-dsRNA complex

    The principal ligand of TLR3 is the double-stranded RNA

    that can be generated during virus replication (3). The

    receptor is localized mainly in endocytic organelles and its

    activation requires TRIF instead of MyD88 (Fig. 6A). The

    crystal structure of the complex between the TLR3 extracel-

    lular domain and a 46 bp double-stranded RNA has been

    determined by x-ray crystallography (18). TLR3 has two

    RNA binding sites located in its N-terminal and C-terminal

    regions, respectively (Fig. 5A). The distance between the two

    sites is approximately 70 A˚

    , which corresponds to roughlytwo helical turns of RNA. The backbone phosphates and

    sugars of dsRNA make major contributions to TLR3 binding

    at both sites. Ligand binding and TLR3 dimerization require

    an acidic environment below pH 6.0 (77). The TLR3 struc-

    ture can explain this strong pH dependence because several

    histidines make crucial bonds with the phosphate backbones

    of the RNA, and their protonation states appear essential for

    stability of the interaction. The nucleic acid bases make min-

    imal contact with the TLR, which accounts for the

    sequence-independent binding of RNA. Although approxi-

    mately 40 bp of RNA appears to be the minimum requiredto induce TLR3 dimerization, a robust immune response

    requires RNA of more than approximately 100 bp. Further-

    more, binding of RNA by TLR3 shows strong positive coo-

    perativity (77). These observations suggest that clustering of 

    more than one TLR3 dimer may be necessary for an ade-

    quate immune response. A recent study supports this model

    (30). They produced antibodies that can block activity of 

    TLR3. Interestingly, two of these antibodies did not directly

    interfere with neither of the two ligand-binding sites. Based

    on crystallographic and molecular modeling study, they pro-

    posed that these antibodies interrupt TLR3 activation by

    inhibiting lateral clustering of the TLR3 dimers.

    Homodimerization of the TLR5-flagellin complex

    Flagellae are large protein complexes that confer motility on

    bacteria. They are made up of multiple copies of approxi-

    mately 20 different protein subunits (78, 79). Electron

    microscopy shows that they consist of three substructures, a

    basal body generating torque, a hook connecting the basal

    ©

     2012 John Wiley & Sons A/SImmunological Reviews 250/2012   223

    Song & Lee     Structure of Toll-like receptors

  • 8/21/2019 Tlr and Pamp

    9/15

    body to a helical fiber, and the long flagellin fiber. The latter

    is assembled from a large number of fliC flagellins. A high-

    resolution structure of a fliC fragment has been determined byx-ray crystallography (80). To obtain monomeric proteins

    suitable for crystallization, the D0 domain formed by the N-

    terminal and C-terminal fragments was removed by proteoly-

    sis and the resulting approximately 45 kD fragment contain-

    ing the remaining middle domains, D1, D2, and D3, was

    used for structure analysis. The highly conserved D1 domain

    plays the main role in polymerization of the fliC subunits into

    a helical fiber. It is composed of three long  a  helices and a  b

    hairpin. Two of the helices are formed by the N-terminal part

    of the protein, and the third helix by the C-terminal region.

    The helices and the  b  hairpin are arranged in an anti-parallel

    bundle-like structure. The D2 domain is composed of severalb  hairpin structures and connects the D1 and D3 domains.

    The D3 domain is formed by the middle part of fliC, and its

    sequence is highly diverse in different bacterial species. It is

    exposed to the surface of the flagellin fiber and mediates the

    reversible lateral interactions between flagellin fibers. Because

    neither vertebrates nor plants contain flagellae, the presence

    of these structures serves as an early sign of bacterial infection

    in both classes of organism (81). TLR5 is the major innate

    immune receptor for flagellin in vertebrates.

    Plasma membrane

    dsRNATriacyl

    lipoprotein

    Diacyl

    lipoprotein   LPSFlagellin

    ssRNAUnmethylated

    CpG DNA

    MAL

         T     L     R     1

         T     L     R     2

    MAL

         T     L     R     2

         T     L     R     6

         T     L     R     5

         T     L     R     5

    MAL

         T     L     R     4

         T     L     R     4

         M     D   -     2

         M     D   -     2

         T     L     R     3

         T     L     R     3

    TIR

    TIRTIR

    TIR

    TIRTIR

    TIRTIR

    TLR4-MD-2TLR4-MD-2LPS

    MyD88

    IRAK4 death domain

    IRAK2 death domain

    MAL

    Myddosome

    B

    MAL

    A

    TRIF

    Endosome

    Fig. 6.  Toll-like receptor (TLR) signaling complexes.  (A) A variety of signaling and bridging adapters are recruited to active TLRs. (B)Proposed structure of the TLR4-MD-2-LPS-Myddosome complex. The structures of the multimerized TIR complexes between receptors andadapters are not known and are shown schematically as boxes.

    ©

     2012 John Wiley & Sons A/S224   Immunological Reviews 250/2012

    Song & Lee     Structure of Toll-like receptors

  • 8/21/2019 Tlr and Pamp

    10/15

    Recently, the crystal structure of zebrafish TLR5 has been

    determined in a complex with a truncated fragment of 

    Salmonella   fliC (21). As with TLR2 and TLR4, several hybrid

    proteins consisting of TLR fragments fused with VLR

    proteins from jawless fishes were used for crystallization

    (15 – 17). The TLR5 of fish and humans have substantial

    sequence homology, suggesting that the two proteins havesimilar structures and ligand recognition mechanisms. The

    structure obtained shows that the flagellin D1 domain plays

    the dominant role in binding and dimerization of TLR5

    (Fig. 5A). One face of the D1 domain interacts with an

    extended surface area encompassing the LRR modules from

    LRRNT to LRR9 of TLR5. At the same time, the D1 domain

    forms a bond with the second TLR5 molecule in the dimer

    and therefore bridges the two TLRs. The overall shape and

    curvature of the horseshoe-like structure formed differ in

    only minor ways from those of TLR5 without bound ligand.

    This observation applies to all TLR structures, thus underlin-

    ing the high structural rigidity of the LRR fold. The spatial

    arrangement of the two TLR5 molecules in the dimer also

    resembles closely that of the other TLR dimers. The two C-

    terminal regions join in the center and the two N-termini

    splay outwards. Because the TLR5 fragment that they used

    for crystallization does not contain native TLR sequences

    beyond LRR14, it is not clear from the structure alone if the

    C-terminal part of the receptor contributes to dimerization

    or to ligand interaction. However, mutagenesis experiments

    suggest that the deleted C-terminal region probably has onlya minor role (21).

    Dimerization of intracellular domains and recruitment of 

    signaling adapters

    Binding of agonistic ligands initiates intracellular signaling

    by recruiting specific adapters to the intracellular TIR

    domains of TLRs (Fig. 6A). In humans, TLR activation

    involves five proteins, MyD88, MAL, TRIF, TRAM, and

    SARM (34, 35). MyD88 is required for signaling by all TLRs

    except TLR3. The latter uses TRIF instead of MyD88. The

    bridging adapter, MAL, is necessary for MyD88-dependentsignaling by TLR2 and TLR4. The MyD88-MAL-dependent

    signaling by TLR4 is activated mainly on the plasma mem-

    brane, whereas TRIF-TRAM-dependent signaling by TLR4 is

    activated on endocytic organelles (82). TRAM is the bridg-

    ing adapter for TRIF. It is necessary for signaling by TLR4

    but not by TLR3. SARM negatively regulates TLR signaling

    by interacting with TRIF and blocking its function (83). All

    these adapters contain TIR domains that are essential for sig-

    nal activation. Mutagenesis experiments have shown that

    recruitment of adapters to receptor TIRs is mediated by TIR

     – TIR interactions. In addition to TIR domains, MAL contains

    an N-terminal phosphatidyl inositol-binding motif and

    TRAM has a myristoylation site (84, 85). These additional

    motifs are necessary for membrane localization of these

    bridging adapters.

    An important aim of research has been to determine thestructures of TIR homo- and heterocomplexes. However, the

    TLR TIR domains produced as truncated forms by genetic

    manipulation interact only weakly or not at all with the TIR

    domains of signaling adapters (28). This makes the direct

    structural study of the TIR complexes very difficult. Hence,

    identification of interaction interfaces has been attempted by

    site-directed mutagenesis. Ronni   et al.   (86) introduced

    approximately 70 alanine substitution and short deletion

    mutations into the human TLR4 TIR domain and measured

    the effect on the IL-7 response in transfected HEK293 cells.

    To differentiate the signal of the introduced TLR4 from the

    endogenous TLR4, they used a constitutively active chimeric

    receptor with the extracellular domain of CD40 and the

    intracellular domain of TLR4. Using this chimeric receptor,

    they found that mutations clustered in two regions, the BB

    and DD loops, resulted in substantial attenuation of the

    TLR4 signaling. However, it is not clear whether these

    regions are involved in the homotypic aggregation of the

    TLR TIRs or the heterotypic interactions with adapter TIR

    domains.

    Recently, Bovijn  et al.  (87) advanced research on TIR – 

    TIRinteraction by identifying mutations that specifically affect

    homotypic interactions of TLR4 TIR domains using a tech-

    nique called MAPPIT, which detects protein interactions in

    mammalian cells using chimeric receptors. They discovered

    several interesting aspects of TIR interaction. First, the

    MyD88-TLR4 interaction requires as bridging adapter, the

    MAL TIR domain. Second, the TIR domains of the bridging

    adapters MAL and TRAM can interact directly with the TLR4

    TIR domain without the help of signaling adapters, MyD88

    or TRIF. Third, they found that the regions surrounding the

    BB loops are crucial for TLR4 TIR homomultimerization.Finally, they also found that all mutations that affect the

    MAL or TRAM interaction block TLR4 TIR multimerization,

    suggesting that TLR4 TIR multimerization is a prerequisite

    for the MAL/TRAM interaction.

    In addition to the TIR domain, MyD88 contains a death

    domain that is required for activation of downstream effec-

    tors. The downstream kinases that are directly recruited to

    the TLR-MyD88 complex are IRAK4 and IRAK1/IRAK2. Lin

    et al.   (25) reported a crystal structure of the death domain

    ©

     2012 John Wiley & Sons A/SImmunological Reviews 250/2012   225

    Song & Lee     Structure of Toll-like receptors

  • 8/21/2019 Tlr and Pamp

    11/15

    complex of MyD88-IRAK4-IRAK2 (Fig. 6B). The TIR domain

    of MyD88 and kinase domains of IRAKs were removed for

    the crystallization. These workers found that the death

    domains form a large helical complex that they have named

    as ‘Myddosome’. In this complex, six MyD88, four IRAK4,

    and four IRAK2 death domains are arranged as a left-handed

    helix composed of four distinct layers. The top two layershave six MyD88s, the middle layer, four IRAK4s, and the

    bottom layer, four IRAK2s. Some deletion or substitution

    mutations that cause life threatening bacterial infections

    mapped in the binding interface, demonstrating an indis-

    pensable role for the complex in TLR signaling. Based on

    this structure, it was proposed that formation of the Myddo-

    some brings the kinase domains of the IRAKs into proximity

    so that they can be phosphorylated and activated.

    Because the Myddosome complex has six MyD88 death

    domains, it appears that the complete TLR signaling com-

    plex contains more than one set of TLR dimers. If we

    assume that the TIR domain of MyD88 interacts 1:1 with

    the TLR TIR, then six MyD88 molecules in the Myddosome

    are capable of binding three TLR dimers (Fig. 6B). This kind

    of higher order clustering of receptor multimers has been

    proposed for quite a few membrane receptors (30, 88 – 92).

    However, characterization of high-resolution structures of 

    the clusters has not been possible, mostly due to difficulties

    in making full-length receptors and reconstituting complete

    signaling complexes in a membrane environment.

    Structures of functionally and structurally related

    proteins, LBP, CD14, RP105, and MD-1

    LPS and lipoproteins are amphipathic macromolecules, and

    they form aggregated micelle structures in water solution.

    Because of this, efficient extraction of these molecules from

    bacterial membrane requires accessary proteins, lipopolysac-

    charide-binding protein (LBP), and CD14. LBP is an acutely

    induced plasma protein that binds avidly to LPS and delivers

    it to CD14 (93). Addition of purified LBP enhances the

    sensitivity of macrophages to LPS 100- to 1000-fold (94).

    LBP belongs to the lipid transfer/lipopolysaccharide-bindingprotein (LT/LBP) family (95). Other members of the family

    include bactericidal/permeability-increasing protein (BPI),

    cholesterol ester transfer protein (CETP), phospholipid

    transfer protein (PLTP), and a few poorly characterized pro-

    teins. LBP and BPI share 48% sequence identity and bind to

    and regulate the biological effects of LPS. BPI is a boomer-

    ang-shaped protein formed by two domains of similar size

    connected by a proline-rich linker (96) (Fig. 7A). Each

    domain has a deep hydrophobic pocket that can bind

    phospholipids on the concave surface of the boomerang. As

    the sequences of LBP and BPI can be aligned over their

    entire lengths, LBP is expected to share the general structural

    features of BPI: its elongated shape and pseudosymmetry as

    well as the three structural elements, the two barrels, central

    sheet, and apolar phospholipid-binding pockets. However,

    the two proteins have clear functional difference: CD14transfers LPS to TLR4-MD-2, but BPI cannot. This difference

    need to be explained by future structural study of LBP.

    CD14 is an LRR family protein that is involved in the

    transfer of LPS from LBP to the TLR4-MD-2 complex. It is

    expressed on the surface of myelomonocytic cells as a gly-

    cosylphosphatidylinositol (GPI)-linked glycoprotein or in

    soluble form in serum (97). In addition to the LPS of 

    Gram-negative bacteria, CD14 can bind other amphipathic

    microbial products, such as peptidoglycan, lipoteichoic acid,

    lipoarabinomannan, and lipoproteins (98, 99). The mono-

    meric subunit of CD14 contains 11 LRR modules and a sin-

    gle LRRNT module (100) (Fig. 7B). CD14 does not contain

    an LRRCT module, which is required to cover the hydro-

    phobic core of the protein from exposure to solvents.

    Instead, CD14 uses the exposed hydrophobic core to form

    a homodimer. Each CD14 monomer contains an N-terminal

    hydrophobic pocket that is formed between the LRRNT and

    the first LRR modules. The convex surface between these

    BPI CD14

    RP105-MD-1

    N N

    C   C

    RP105   RP105

    A

    C

    B

    Fig. 7.  Structures of proteins involved in Toll-like receptor

    activation. Structures of BPI (A), CD14 (B), and the RP105-MD-1complex (C). CD14 forms a homodimer. The monomeric subunits of CD14 are colored in green and blue, respectively. The MD-1 moleculesbound to RP-105s are schematically drawn in grey.

    ©

     2012 John Wiley & Sons A/S226   Immunological Reviews 250/2012

    Song & Lee     Structure of Toll-like receptors

  • 8/21/2019 Tlr and Pamp

    12/15

    two modules is cracked open, exposing the hydrophobic

    core residues inside. The size of the pocket is large enough

    to accommodate the lipid chains of a single LPS molecule.

    Therefore, the function of CD14 appears to be monomer-

    ization of the aggregated ligands for efficient binding to

    TLR4.

    RP105 and MD-1 are structurally and sequentially homol-ogous to TLR4 and MD-2, respectively (101 – 103). How-

    ever, unlike TLR4, RP105 does not contain a sizable

    intracellular domain, suggesting it cannot activate signaling

    pathways by itself. The exact function and the physiological

    ligands for the RP105-MD-1 complex have not been clearly

    identified yet, although it has been proposed that this pro-

    tein complex regulates TLR2 and TLR4 signaling by directly

    interacting with them (104 – 107). Several structures of the

    RP105-MD-1 complex have recently been determined (108 – 

    111) (Fig. 7C). The individual structures of RP105 and MD-

    1 closely resemble those of TLR4 and MD-2, respectively,

    and the primary interaction interface between RP105 and

    MD-1 shows a homology with that of TLR4 and MD-2.

    However, unlike TLR4-MD-2, which requires LPS for the

    formation of a heterotetramer, the RP105-MD-1 can form a

    stable heterotetramer without bound ligands. Furthermore,

    the overall arrangement of the two RP105 in the complex

    shows a clear difference from the arrangement of TLR4 in

    the TLR4-MD-2-LPS complex: the N-terminal LRR modules,

    instead of the C-terminal modules, of the two RP105 chains

    converge in the center of the complex (Fig. 7C). This head-

    to-head arrangement of RP105 contrasts with the tail-to-tail

    mode, which is highly conserved in the ligand-activated

    TLR homo- and heterodimers (Fig. 5A). The structure of the

    MD-1 pocket is homologous with that of MD-2, suggesting

    that MD-1 may be able to interact with flat hydrophobic

    ligands like LPS.

    Conclusion

    Over the last 10 years, important progress has been made in

    structural understanding of the ligand recognition and acti-

    vation mechanisms of TLRs. The structures of the extracellu-

    lar domains of TLRs 1 – 6 have been determined in complex

    with their cognate ligands. The structures of the intracellular

    TIR domains of several TLRs and adapters have also been

    determined. These structures show us how TLRs recognize

    their ligands and how ligand binding induces structuralrearrangements of the receptor. For the next 10 years, the

    most important goal should be determining structures of the

    complete signaling complexes comprising the full-length

    receptors, ligands, signaling adapters, bridging adapters, and

    downstream kinases. To achieve this ambitious goal, we

    must first find ways to produce enough amounts of structur-

    ally and functionally intact receptors and to reconstitute the

    complete receptor-ligand adapter-kinase complexes from the

    purified components.

    References

    1. Janeway CA, Jr. Approaching the asymptote?

    Evolution and revolution in

    immunology. Cold Spring Harb Symp Quant

    Biol 1989;54:1 – 13.

    2. Medzhitov R. Toll-like receptors and

    innate immunity. Nat Rev Immunol

    2001;1:135 – 145.

    3. Gay NJ, Gangloff M. Structure and function of 

    Toll receptors and their ligands. Annu Rev

    Biochem 2007;76:141 – 165.

    4. Kang JY, Lee JO. Structural biology of the Toll-

    like receptor family. Annu Rev Biochem

    2011;80:917 – 

    941.5. Hennessy EJ, Parker AE, O’Neill LA. Targeting

    Toll-like receptors: emerging therapeutics? Nat

    Rev Drug Discov 2010;9:293 – 307.

    6. Bennett-Guerrero E, et al. A phase II, double-

    blind, placebo-controlled, ascending-dose study

    of Eritoran (E5564), a lipid A antagonist, in

    patients undergoing cardiac surgery with

    cardiopulmonary bypass. Anesth Analg

    2007;104:378 – 383.

    7. Tidswell M, et al. Phase 2 trial of eritoran

    tetrasodium (E5564), a toll-like receptor 4

    antagonist, in patients with severe sepsis. Crit

    Care Med 2010;38:72 – 83.

    8. Ii M, et al. A novel cyclohexene derivative, ethyl

    (6R)-6-[N-(2-Chloro-4-fluorophenyl)sulfamoyl]

    cyclohex-1-ene-1-carboxylate (TAK-242),

    selectively inhibits toll-like receptor 4-mediated

    cytokine production through suppression of 

    intracellular signaling. Mol Pharmacol

    2006;69:1288 – 1295.

    9. Kawamoto T, Ii M, Kitazaki T, Iizawa Y,

    Kimura H. TAK-242 selectively suppresses Toll-

    like receptor 4-signaling mediated by the

    intracellular domain. Eur J Pharmacol

    2008;584:40 – 48.

    10. Casella CR, Mitchell TC. Putting endotoxin towork for us: monophosphoryl lipid A as a safe

    and effective vaccine adjuvant. Cell Mol Life Sci

    2008;65:3231 – 3240.

    11. Stevceva L. Toll-like receptor agonists as

    adjuvants for HIV vaccines. Curr Med Chem

    2011;18:5079 – 5082.

    12. Mizel SB, Bates JT. Flagellin as an adjuvant:

    cellular mechanisms and potential. J Immunol

    2010;185:5677 – 5682.

    13. Bell JK, Askins J, Hall PR, Davies DR, Segal DM.

    The dsRNA binding site of human Toll-like

    receptor 3. Proc Natl Acad Sci USA

    2006;103:8792 – 8797.

    14. Choe J, Kelker MS, Wilson IA. Crystal structure

    of human toll-like receptor 3 (TLR3)

    ectodomain. Science 2005;309:581 – 585.

    15. Jin MS, et al. Crystal structure of the TLR1-TLR2

    heterodimer induced by binding of a tri-acylated

    lipopeptide. Cell 2007;130:1071 – 1082.

    16. Kang JY, et al. Recognition of lipopeptide

    patterns by Toll-like receptor 2-Toll-like receptor

    6 heterodimer. Immunity 2009;31:873 – 884.

    17. Kim HM, et al. Crystal structure of the TLR4-

    MD-2 complex with bound endotoxin antagonist

    Eritoran. Cell 2007;130:906 – 917.

    18. Liu L, et al. Structural basis of toll-like receptor 3signaling with double-stranded RNA. Science

    2008;320:379 – 381.

    19. Ohto U, Fukase K, Miyake K, Satow Y. Crystal

    structures of human MD-2 and its complex with

    antiendotoxic lipid IVa. Science 2007;316:1632 – 

    1634.

    20. Park BS, Song DH, Kim HM, Choi BS, Lee H, Lee

     JO. The structural basis of lipopolysaccharide

    recognition by the TLR4-MD-2 complex. Nature

    2009;458:1191 – 1195.

    21. Yoon SI, et al. Structural basis of TLR5-flagellin

    recognition and signaling. Science

    2012;335:859 – 864.

    ©

     2012 John Wiley & Sons A/SImmunological Reviews 250/2012   227

    Song & Lee     Structure of Toll-like receptors

  • 8/21/2019 Tlr and Pamp

    13/15

    22. Bella J, Hindle KL, McEwan PA, Lovell SC. The

    leucine-rich repeat structure. Cell Mol Life Sci

    2008;65:2307 – 2333.

    23. Kobe B, Kajava AV. The leucine-rich repeat as a

    protein recognition motif. Curr Opin Struct Biol

    2001;11:725 – 732.

    24. Khan JA, Brint EK, O’Neill LA, Tong L. Crystal

    structure of the Toll/interleukin-1 receptor

    domain of human IL-1RAPL. J Biol Chem2004;279:31664 – 31670.

    25. Lin SC, Lo YC, Wu H. Helical assembly in the

    MyD88-IRAK4-IRAK2 complex in TLR/IL-1R

    signalling. Nature 2010;465:885 – 890.

    26. Nyman T, Stenmark P, Flodin S, Johansson I,

    Hammarstrom M, Nordlund P. The crystal

    structure of the human toll-like receptor 10

    cytoplasmic domain reveals a

    putative signaling dimer. J Biol Chem

    2008;283:11861 – 11865.

    27. Ohnishi H, et al. Structural basis for the multiple

    interactions of the MyD88 TIR domain in TLR4

    signaling. Proc Natl Acad Sci USA

    2009;106:10260 – 10265.

    28. Xu Y, et al. Structural basis for signaltransduction by the Toll/interleukin-1 receptor

    domains. Nature 2000;408:111 – 115.

    29. Valkov E, et al. Crystal structure of Toll-like

    receptor adaptor MAL/TIRAP reveals the

    molecular basis for signal transduction and

    disease protection. Proc Natl Acad Sci USA

    2011;108:14879 – 14884.

    30. Luo J, et al. Lateral clustering of TLR3:dsRNA

    signaling units revealed by TLR3ecd:3Fabs

    quaternary structure. J Mol Biol 2012;421:112 – 

    124.

    31. Bell JK, et al. The molecular structure of the Toll-

    like receptor 3 ligand-binding domain. Proc Natl

    Acad Sci USA 2005;102:10976 – 10980.

    32. Jin MS, Lee JO. Structures of the toll-like receptorfamily and its ligand complexes. Immunity

    2008;29:182 – 191.

    33. Jin MS, Lee JO. Structures of TLR-ligand

    complexes. Curr Opin Immunol 2008;20:414 – 

    419.

    34. Kenny EF, O’Neill LA. Signalling adaptors used

    by Toll-like receptors: an update. Cytokine

    2008;43:342 – 349.

    35. Watters TM, Kenny EF, O’Neill LA. Structure,

    function and regulation of the Toll/IL-1 receptor

    adaptor proteins. Immunol Cell Biol

    2007;85:411 – 419.

    36. Chan SL, et al. Molecular mimicry in innate

    immunity: crystal structure of a bacterial

    TIR domain. J Biol Chem 2009;284:21386 – 

    21392.

    37. Gautam JK, Ashish, Comeau LD, Krueger JK,

    Smith MF, Jr. Structural and functional evidence

    for the role of the TLR2 DD loop in TLR1/TLR2

    heterodimerization and signaling. J Biol Chem

    2006;281:30132 – 30142.

    38. Nunez Miguel R, et al. A dimer of the Toll-like

    receptor 4 cytoplasmic domain provides a

    specific scaffold for the recruitment of signalling

    adaptor proteins. PLoS ONE 2007;2:e788.

    39. Underhill DM, et al. The Toll-like receptor 2 is

    recruited to macrophage phagosomes and

    discriminates between pathogens. Nature

    1999;401:811 – 815.

    40. Takashima K, et al. Analysis of binding site for

    the novel small-molecule TLR4 signal

    transduction inhibitor TAK-242 and its

    therapeutic effect on mouse sepsis model. Br J

    Pharmacol 2009;157:1250 – 1262.

    41. Newman RM, Salunkhe P, Godzik A, Reed JC.

    Identification and characterization of a novelbacterial virulence factor that shares homology

    with mammalian Toll/interleukin-1 receptor

    family proteins. Infect Immun 2006;74:594 – 

    601.

    42. Salcedo SP, et al. Brucella control of dendritic cell

    maturation is dependent on the TIR-containing

    protein Btp1. PLoS Pathog 2008;4:e21.

    43. Cirl C, et al. Subversion of Toll-like receptor

    signaling by a unique family of bacterial Toll/

    interleukin-1 receptor domain-containing

    proteins. Nat Med 2008;14:399 – 406.

    44. Spear AM, Loman NJ, Atkins HS, Pallen MJ.

    Microbial TIR domains: not necessarily

    agents of subversion? Trends Microbiol

    2009;17:393 – 

    398.45. Chisholm ST, Coaker G, Day B, Staskawicz BJ.

    Host-microbe interactions: shaping the evolution

    of the plant immune response. Cell

    2006;124:803 – 814.

    46. Rafiqi M, Bernoux M, Ellis JG, Dodds PN. In the

    trenches of plant pathogen recognition: role of 

    NB-LRR proteins. Semin Cell Dev Biol

    2009;20:1017 – 1024.

    47. Chan SL, Mukasa T, Santelli E, Low LY, Pascual J.

    The crystal structure of a TIR domain from

    Arabidopsis thaliana reveals a conserved helical

    region unique to plants. Protein Sci

    2010;19:155 – 161.

    48. Bernoux M, et al. Structural and functional

    analysis of a plant resistance protein TIR domainreveals interfaces for self-association, signaling,

    and autoregulation. Cell Host Microbe

    2011;9:200 – 211.

    49. Beutler B, Rietschel ET. Innate immune sensing

    and its roots: the story of endotoxin. Nat Rev

    Immunol 2003;3:169 – 176.

    50. Raetz CR. Biochemistry of endotoxins. Annu Rev

    Biochem 1990;59:129 – 170.

    51. Raetz CR, Whitfield C. Lipopolysaccharide

    endotoxins. Annu Rev Biochem 2002;71:635 – 

    700.

    52. Erridge C, Bennett-Guerrero E, Poxton IR.

    Structure and function of lipopolysaccharides.

    Microbes Infect 2002;4:837 – 851.

    53. Galanos C, et al. Synthetic and natural Escherichiacoli free lipid A express identical endotoxic

    activities. Eur J Biochem 1985;148:1 – 5.

    54. Rietschel ET, et al. Bacterial endotoxin: molecular

    relationships of structure to activity and function.

    FASEB J 1994;8:217 – 225.

    55. Medzhitov R, Preston-Hurlburt P, Janeway CA,

     Jr. A human homologue of the Drosophila Toll

    protein signals activation of adaptive immunity.

    Nature 1997;388:394 – 397.

    56. Poltorak A, et al. Defective LPS signaling in C3H/

    HeJ and C57BL/10ScCr mice: mutations in Tlr4

    gene. Science 1998;282:2085 – 2088.

    57. Shimazu R, et al. MD-2, a molecule that confers

    lipopolysaccharide responsiveness on Toll-like

    receptor 4. J Exp Med 1999;189:1777 – 1782.

    58. Nagai Y, et al. Essential role of MD-2 in LPS

    responsiveness and TLR4 distribution. Nat

    Immunol 2002;3:667 – 672.

    59. Schromm AB, et al. Molecular genetic analysis of 

    an endotoxin nonresponder mutant cell line: a

    point mutation in a conserved region of MD-2abolishes endotoxin-induced signaling. J Exp

    Med 2001;194:79 – 88.

    60. Viriyakosol S, Tobias PS, Kitchens RL, Kirkland

    TN. MD-2 binds to bacterial lipopolysaccharide. J

    Biol Chem 2001;276:38044 – 38051.

    61. Rietschel ET, et al. The chemical structure of 

    bacterial endotoxin in relation to bioactivity.

    Immunobiology 1993;187:169 – 190.

    62. Ohto U, Fukase K, Miyake K, Shimizu T.

    Structural basis of species-specific endotoxin

    sensing by innate immune receptor TLR4/

    MD-2. Proc Natl Acad Sci USA 2012;109:

    7421 – 7426.

    63. Qureshi N, Takayama K, Ribi E. Purification and

    structural determination of nontoxic lipid Aobtained from the lipopolysaccharide of 

    Salmonella typhimurium. J Biol Chem

    1982;257:11808 – 11815.

    64. Takeuchi O, et al. Discrimination of bacterial

    lipoproteins by Toll-like receptor 6. Int Immunol

    2001;13:933 – 940.

    65. Takeuchi O, et al. Cutting edge: role of Toll-like

    receptor 1 in mediating immune response to

    microbial lipoproteins. J Immunol 2002;169:10 – 

    14.

    66. Hantke K, Braun V. Covalent binding of lipid to

    protein Diglyceride and amide-linked fatty acid at

    the N-terminal end of the murein-lipoprotein of 

    the Escherichia coli outer membrane. Eur J

    Biochem 1973;34:284 – 

    296.67. Belisle JT, Brandt ME, Radolf JD, Norgard MV.

    Fatty acids of Treponema pallidum and Borrelia

    burgdorferi lipoproteins. J Bacteriol

    1994;176:2151 – 2157.

    68. Braun V. Covalent lipoprotein from the outer

    membrane of Escherichia coli. Biochim Biophys

    Acta 1975;415:335 – 377.

    69. Mizuno T. A novel peptidoglycan-associated

    lipoprotein found in the cell envelope of 

    Pseudomonas aeruginosa and Escherichia coli. J

    Biochem 1979;86:991 – 1000.

    70. Zlotnick GW, Sanfilippo VT, Mattler JA, Kirkley

    DH, Boykins RA, Seid RC, Jr. Purification and

    characterization of a peptidoglycan-associated

    lipoprotein from Haemophilus influenzae. J BiolChem 1988;263:9790 – 9794.

    71. Hutchings MI, Palmer T, Harrington DJ, Sutcliffe

    IC. Lipoprotein biogenesis in Gram-positive

    bacteria: knowing when to hold ‘em, knowing

    when to fold ‘em. Trends Microbiol

    2009;17:13 – 21.

    72. Muhlradt PF, Kiess M, Meyer H, Sussmuth R,

     Jung G. Isolation, structure elucidation, and

    synthesis of a macrophage stimulatory

    lipopeptide from Mycoplasma fermentans acting

    at picomolar concentration. J Exp Med

    1997;185:1951 – 1958.

    ©

     2012 John Wiley & Sons A/S228   Immunological Reviews 250/2012

    Song & Lee     Structure of Toll-like receptors

  • 8/21/2019 Tlr and Pamp

    14/15

    73. Shibata K, Hasebe A, Into T, Yamada M,

    Watanabe T. The N-terminal lipopeptide of a 44-

    kDa membrane-bound lipoprotein of 

    Mycoplasma salivarium is responsible for the

    expression of intercellular adhesion molecule-1

    on the cell surface of normal

    human gingival fibroblasts. J Immunol

    2000;165:6538 – 6544.

    74. Kurokawa K, et al. The triacylated ATP bindingcluster transporter substrate-binding lipoprotein

    of Staphylococcus aureus functions as a native

    ligand for Toll-like receptor 2. J Biol Chem

    2009;284:8406 – 8411.

    75. Tawaratsumida K, et al. Characterization of N-

    terminal structure of TLR2-activating lipoprotein

    in Staphylococcus aureus. J Biol Chem

    2009;284:9147 – 9152.

    76. Omueti KO, Mazur DJ, Thompson KS, Lyle EA,

    Tapping RI. The polymorphism P315L of human

    Toll-like receptor 1 impairs innate immune

    sensing of microbial cell wall components. J

    Immunol 2007;178:6387 – 6394.

    77. Leonard JN, et al. The TLR3 signaling complex

    forms by cooperative receptor dimerization. ProcNatl Acad Sci USA 2008;105:258 – 263.

    78. Berg HC. The rotary motor of bacterial flagella.

    Annu Rev Biochem 2003;72:19 – 54.

    79. Kojima S, Blair DF. The bacterial flagellar motor:

    structure and function of a complex molecular

    machine. Int Rev Cytol 2004;233:93 – 134.

    80. Samatey FA, et al. Structure of the bacterial

    flagellar protofilament and implications for a

    switch for supercoiling. Nature 2001;410:331 – 

    337.

    81. Zipfel C. Pattern-recognition receptors in plant

    innate immunity. Curr Opin Immunol

    2008;20:10 – 16.

    82. Gangloff M. Different dimerisation mode for

    TLR4 upon endosomal acidification? TrendsBiochem Sci 2012;37:92 – 98.

    83. Carty M, Goodbody R, Schroder M, Stack J,

    Moynagh PN, Bowie AG. The human adaptor

    SARM negatively regulates adaptor protein TRIF-

    dependent Toll-like receptor signaling. Nat

    Immunol 2006;7:1074 – 1081.

    84. Kagan JC, Medzhitov R. Phosphoinositide-

    mediated adaptor recruitment controls

    Toll-like receptor signaling. Cell 2006;125:943 – 

    955.

    85. Rowe DC, et al. The myristoylation of TRIF-

    related adaptor molecule is essential for Toll-like

    receptor 4 signal transduction. Proc Natl Acad Sci

    USA 2006;103:6299 – 6304.

    86. Ronni T, Agarwal V, Haykinson M, Haberland

    ME, Cheng G, Smale ST. Common interaction

    surfaces of the toll-like receptor 4 cytoplasmic

    domain stimulate multiple nuclear targets. Mol

    Cell Biol 2003;23:2543 – 2555.

    87. Bovijn C, et al. Identification of interaction sites

    for dimerization and adapter recruitment in Toll/

    interleukin-1 receptor (TIR) domain of Toll-like

    receptor 4. J Biol Chem 2012;287

    :4088 – 

    4098.88. Abulrob A, et al. Nanoscale imaging of 

    epidermal growth factor receptor clustering:

    effects of inhibitors. J Biol Chem

    2010;285:3145 – 3156.

    89. Germain RN. T-cell signaling: the importance of 

    receptor clustering. Curr Biol 1997;7:R640 – 

    R644.

    90. Himanen JP, et al. Architecture of Eph receptor

    clusters. Proc Natl Acad Sci USA 2010;107:

    10860 – 10865.

    91. Stock J. Receptor signaling: dimerization and

    beyond. Curr Biol 1996;6:825 – 827.

    92. Han SH, Kim JH, Martin M, Michalek SM, Nahm

    MH. Pneumococcal lipoteichoic acid (LTA) is not

    as potent as staphylococcal LTA in stimulatingToll-like receptor 2. Infect Immun

    2003;71:5541 – 5548.

    93. Weiss J. Bactericidal/permeability-increasing

    protein (BPI) and lipopolysaccharide-binding

    protein (LBP): structure, function and regulation

    in host defence against Gram-negative bacteria.

    Biochem Soc Trans 2003;31:785 – 790.

    94. Martin TR, et al. Lipopolysaccharide binding

    protein enhances the responsiveness of alveolar

    macrophages to bacterial lipopolysaccharide

    Implications for cytokine production in normal

    and injured lungs. J Clin Invest 1992;90:2209 – 

    2219.

    95. Mulero JJ, et al. Three new human members of 

    the lipid transfer/lipopolysaccharide bindingprotein family (LT/LBP). Immunogenetics

    2002;54:293 – 300.

    96. Beamer LJ, Carroll SF, Eisenberg D. Crystal

    structure of human BPI and two bound

    phospholipids at 2.4 angstrom resolution. Science

    1997;276:1861 – 1864.

    97. Tobias PS, Soldau K, Gegner JA, Mintz D,

    Ulevitch RJ. Lipopolysaccharide binding protein-

    mediated complexation of lipopolysaccharide

    with soluble CD14. J Biol Chem

    1995;270:10482 – 10488.

    98. Dziarski R. Recognition of bacterial

    peptidoglycan by the innate immune system. Cell

    Mol Life Sci 2003;60:1793 – 1804.

    99. Gregory CD, Devitt A. CD14 and apoptosis.

    Apoptosis 1999;4:11 – 20.

    100. Kim JI, et al. Crystal structure of CD14 and its

    implications for lipopolysaccharide signaling.

     J Biol Chem 2005;280:11347 – 11351.

    101. Miyake K, et al. Mouse MD-1, a molecule that is

    physically associated with RP105 and positively

    regulates its expression. J Immunol

    1998;161

    :1348 – 

    1353.102. Fugier-Vivier I, et al. Molecular cloning of 

    human RP105. Eur J Immunol 1997;27:1824 – 

    1827.

    103. Miyake K, Yamashita Y, Ogata M, Sudo T,

    Kimoto M. RP105, a novel B cell surface

    molecule implicated in B cell activation,

    is a member of the leucine-rich

    repeat protein family. J Immunol

    1995;154:3333 – 3340.

    104. Divanovic S, et al. Negative regulation of 

    Toll-like receptor 4 signaling by the Toll-like

    receptor homolog RP105. Nat Immunol

    2005;6:571 – 578.

    105. Divanovic S, et al. Regulation of TLR4 signaling

    and the host interface with pathogens anddanger: the role of RP105. J Leukoc Biol

    2007;82:265 – 271.

    106. Ogata H, et al. The toll-like receptor protein

    RP105 regulates lipopolysaccharide signaling in B

    cells. J Exp Med 2000;192:23 – 29.

    107. Nagai Y, et al. Requirement for MD-1 in cell

    surface expression of RP105/CD180 and B-cell

    responsiveness to lipopolysaccharide. Blood

    2002;99:1699 – 1705.

    108. Harada H, Ohto U, Satow Y. Crystal structure of 

    mouse MD-1 with endogenous phospholipid

    bound in its cavity. J Mol Biol 2010; 400:838 – 

    846.

    109. Yoon SI, Hong M, Wilson IA. An unusual

    dimeric structure and assembly for TLR4regulator RP105-MD-1. Nat Struct Mol Biol

    2011;18:1028 – 1035.

    110. Yoon SI, Hong M, Han GW, Wilson IA. Crystal

    structure of soluble MD-1 and its interaction

    with lipid IVa. Proc Natl Acad Sci USA

    2010;107:10990 – 10995.

    111. Ohto U, Miyake K, Shimizu T. Crystal

    structures of mouse and human RP105/MD-1

    complexes reveal unique dimer organization of 

    the toll-like receptor family. J Mol Biol

    2011;413:815 – 825.

    ©

     2012 John Wiley & Sons A/SImmunological Reviews 250/2012   229

    Song & Lee     Structure of Toll-like receptors

  • 8/21/2019 Tlr and Pamp

    15/15

    This document is a scanned copy of a printed document. No warranty is given about the accuracy of the copy.

    Users should refer to the original published version of the material.