Thromb Haemost 2019; 119(04): 553-566
DOI: 10.1055/s-0039-1677803
Theme Issue Article
Georg Thieme Verlag KG Stuttgart · New York

Macrophage Migration Inhibitory Factor (MIF)-Based Therapeutic Concepts in Atherosclerosis and Inflammation

Dzmitry Sinitski*
1   Department of Vascular Biology, Institute for Stroke and Dementia Research (ISD), Klinikum der Universität München (KUM), Ludwig-Maximilians-University (LMU), Munich, Germany
,
Christos Kontos*
2   Division of Peptide Biochemistry, Technische Universität München (TUM), Freising, Germany
,
Christine Krammer*
1   Department of Vascular Biology, Institute for Stroke and Dementia Research (ISD), Klinikum der Universität München (KUM), Ludwig-Maximilians-University (LMU), Munich, Germany
,
Yaw Asare
3   Department of Translational Medicine, Institute for Stroke and Dementia Research (ISD), Klinikum der Universität München (KUM), Ludwig-Maximilians-University (LMU), Munich, Germany
,
Aphrodite Kapurniotu**
2   Division of Peptide Biochemistry, Technische Universität München (TUM), Freising, Germany
,
Jürgen Bernhagen**
1   Department of Vascular Biology, Institute for Stroke and Dementia Research (ISD), Klinikum der Universität München (KUM), Ludwig-Maximilians-University (LMU), Munich, Germany
4   Munich Heart Alliance, Munich, Germany
5   Munich Cluster for Systems Neurology (SyNergy), Munich, Germany
› Author Affiliations
Funding This work was supported by Deutsche Forschungsgemeinschaft (DFG) grant SFB1123-A03 to J.B. and A.K., SFB1123-B03 to Y.A., by DFG within the framework of Munich Cluster for Systems Neurology (EXC 1010 SyNergy) and of LMUexc (LMU-Singapore strategic partnership) to J.B.
Further Information

Address for correspondence

Jürgen Bernhagen, PhD
Chair of Vascular Biology, Institute for Stroke and Dementia Research (ISD)
Klinikum der Universität München (KUM), Ludwig-Maximilians-University (LMU) Munich, Feodor-Lynen-Straße 17, 81377 Munich
Germany   
Aphrodite Kapurniotu, PhD
Division of Peptide Biochemistry, Technische Universität München (TUM)
Emil-Erlenmeyer-Forum 5, 85354 Freising
Germany   

Publication History

01 October 2018

21 December 2018

Publication Date:
04 February 2019 (online)

 

Abstract

Chemokines orchestrate leukocyte recruitment in atherosclerosis and their blockade is a promising anti-atherosclerotic strategy, but few chemokine-based approaches have advanced into clinical trials, in part owing to the complexity and redundancy of the chemokine network. Macrophage migration inhibitory factor (MIF) is a pivotal mediator of atherosclerotic lesion formation. It has been characterized as an inflammatory cytokine and atypical chemokine that promotes atherogenic leukocyte recruitment and lesional inflammation through interactions with the chemokine receptors CXCR2 and CXCR4, but also exhibits phase-specific CD74-mediated cardioprotective activity. The unique structural properties of MIF and its homologue MIF-2/D-DT offer intriguing therapeutic opportunities including small molecule-, antibody- and peptide-based approaches that may hold promise as inhibitors of atherosclerosis, while sparing tissue-protective classical chemokine pathways. In this review, we summarize the pros and cons of anti-MIF protein strategies and discuss their molecular characteristics and receptor specificities with a focus on cardiovascular disease.


#

Introduction

Atherosclerosis is a chronic inflammatory disease of our arteries that is characterized by the development of lipid-rich inflamed plaques in the vessel wall. Lesion progression and plaque rupture may result in detrimental cardiovascular events such as acute myocardial infarction and ischaemic stroke,[1] [2] the leading causes of death worldwide.[3] Influenced by genetic and environmental risk factors such as hyperlipidaemia, atherosclerosis is initiated by endothelial dysfunction, followed by an accumulation of oxidized low-density lipoproteins (oxLDLs) and an inflammatory cell infiltrate dominated by monocytes and T cells into the atherogenic vessel wall. Infiltrating monocytes differentiate into macrophages and lipid-laden foam cells. Lesion progression also involves vascular smooth muscle cell (VSMC) proliferation, necrotic core formation and wall remodelling that may eventually lead to plaque destabilization, rupture and thrombosis.[4]

These processes are mediated by inflammatory cytokines and chemokines at all stages. Some 50 classical chemokines interact with 18 G-protein-coupled receptor (GPCR)-type chemokine receptors. This network is characterized by a high degree of redundancy and promiscuity and chemokines are divided into CC-, CXC-, CX3C- and C-type sub-classes and correspondingly termed receptors.[5] [6]

Due to their causal role in atherogenesis, anti-cytokine/-chemokine approaches are pursued as therapeutic strategies to attenuate atherosclerosis.[7] Several chemokine-blocking antibodies and chemokine receptor-inhibiting small molecule drug (SMD) compounds are in advanced pre-clinical testing and (early) clinical trial phases.[7] [8] [9] [10] [11] Importantly, the promising results obtained with an interleukin-1β (IL-1β)-blocking antibody in the CANTOS trial have validated the inflammatory hypothesis in atherosclerosis and demonstrated the power of cytokine-based anti-inflammatory drugs in patients with established atherosclerotic disease.[12]

Macrophage migration inhibitory factor (MIF) is an inflammatory cytokine with chemokine-like characteristics and unique structural properties and is classified as a prototypical member of the emerging family of atypical chemokines (ACKs).[13] [14] [15] [16] ACKs lack the typical chemokine-fold and conserved N-terminal cysteines of classical chemokines,[6] but exhibit chemotactic activity and bind to classical chemokine receptors.[16] MIF is up-regulated in human atherosclerotic lesions[17] and its levels correlate with coronary artery disease (CAD).[18] [19] Mif gene deletion (Mif-KO) and antibody-based neutralization of MIF in experimental atherosclerosis suggest it is a major driver of atheroprogression during several stages of the disease.[14] [18]

Here, we discuss molecular strategies to inhibit MIF and its structural homologue D-dopachrome tautomerase (D-DT), also termed MIF-2, in atherosclerosis and other inflammatory diseases. We cover antibody-based strategies, small molecules directed at the unique MIF catalytic pocket around N-terminal proline-2 or at allosteric sites and emerging peptide-based approaches. The pros and cons of these strategies, potential side effects and envisaged receptor pathway specificities are compared.


#

MIF is a Chemokine-Like Inflammatory Mediator that Promotes Atherosclerosis

MIF is one of the first cytokines to be discovered. It was originally described in 1966 by John David as a soluble factor produced by human lymphocytes that was capable of inhibiting the random migration of macrophage-like cells out of capillary tubes, while earlier reports on myeloid cell migration even date back to 1932.[15] [20] [21] MIF is a 12.5 kD protein containing 114 amino acids that crystallizes as a trimer, but equilibria between monomers, dimers and trimers are observed under physiological solution conditions.[22] [23] Today, MIF is known as a pleiotropic inflammatory mediator that is structurally distinct from other cytokines, but shares structural homology with bacterial tautomerases/isomerases, suggesting evolutionary conservation.[15] [23] [24] It is broadly expressed, but regulated secretion that occurs from semi-constitutive cytosolic stores by a p115-dependent non-conventional mechanism is predominantly seen in cells of the immune system as well as endothelial and tumour cells.[15] [25] [26] MIF is the founding member of the MIF protein family that also comprises D-DT/MIF-2 and MIF-like orthologs in numerous species. MIF is an upstream regulator of the host innate and adaptive immune response, but—if dysregulated—it is a driver of inflammatory diseases as well as cardiovascular diseases including atherosclerosis. Contrary to its eponymous name, MIF has been classified as an ACK that, similar to arrest chemokines such as CXCL1/8, enhances atherogenic leukocyte chemotaxis and arrest. It has been suggested that inhibition of random macrophage migration as observed in the historic experiments, is likely to represent a desensitization effect as well-known for chemokines.[14] [16]

Serving as an inflammatory, chemokine-like cytokine and upstream regulator of innate immunity, it is not unexpected that MIF has a key role in numerous inflammatory and autoimmune conditions, including septic shock, rheumatoid arthritis (RA), systemic lupus erythematosus, Crohn's disease, obesity, glomerulonephritis and inflammatory and allergic lung conditions (reviewed in Refs.[15] [27] [28] [29] [30]). Owing to the close mechanistic links between chronic inflammation and cancer, MIF also has been identified as a pro-tumorigenic factor in several tumour entities, enhancing cancer cell proliferation, promoting tumour angiogenesis and modulating anti-tumour immunity.[15] [31] [32] [33] [34]

Its chemokine-like and inflammatory properties render MIF a potent regulator of the atherogenic process. MIF expression is up-regulated in human and murine atherosclerotic lesions with peak levels observed in advanced plaques.[17] [19] It is not only up-regulated in the atherogenic endothelium and infiltrating leukocytes, but also in VSMCs and platelets following inflammatory stimulation.[17] [35] [36] Antibody-mediated neutralization in Apoe/− mice resulted in reduced lesional immune cell content and lowered levels of inflammatory mediators associated with atherosclerosis.[37] Similarly, Mif-deficient Ldlr/− mice showed reduced atherosclerotic plaque areas compared with controls.[14] [38] Targeting MIF with neutralizing antibodies resulted in significant plaque regression.[14] The pro-atherogenic activity of MIF is predominantly mediated via non-cognate interaction with the chemokine receptors CXCR2 and CXCR4, leading to monocyte and T cell recruitment, respectively.[14] This is accompanied by an up-regulation of adhesion molecules like intercellular adhesion molecule 1 and release of atherogenic chemokines such as CCL-2.[39] [40] Moreover, MIF stimulates oxLDL uptake to promote foam cell formation. Foam cells undergo apoptosis and form a necrotic core surrounded by a fibrous cap.[41] MIF is associated with plaque instability as it induces matrix degradation through matrix metalloproteinases, followed by fibrous cap thinning resulting in plaque rupture.[42] MIF also promotes intra-plaque inflammation by stimulating macrophages to secrete inflammatory mediators such as tumour necrosis factor (TNF)-α or IL-1β.[43] The role of MIF-2 in chronic atherogenesis is subject to current investigations.

Pro-atherogenic effects of MIF are supported by observational clinical studies in CAD patients. For the G/C single-nucleotide polymorphism rs755622 at position –173, a higher susceptibility to develop CAD has been observed for C allele carriers.[44] [45] [46] Moreover, the MIF gene features a tetranucleotide CATT repeat polymorphism (‘the CATT5–8 microsatellite’) at position –794 that was initially identified in RA patients and controls gene expression from the MIF promoter.[47] CATT7 or CATTnon-5 MIF high expressers show an increased severity of coronary artery atherosclerosis and patients carrying the rs755622 C allele and CATT7/C haplotype are more prone to develop CAD.[48] This correlates with associations between plasma MIF and CAD, for example, in acute coronary syndrome (ACS).[18] [19] [49] [50] Moreover, MIF plasma levels were found to be elevated in a high proportion of ST-elevation myocardial infarction (STEMI) patients, were suggested to be an early marker of acute STEMI, and STEMI patients with high admission MIF level experienced a poorer recovery of cardiac function and worse long-term adverse outcomes.[19] [51] [52] [53] Moreover, the role of MIF in myocardial ischaemia/reperfusion injury (MI/RI) has become apparent from clinical studies in cardiac surgery patients and mouse models. The cardiac surgery procedure recapitulates the ischaemic and reperfusion stress seen in myocardial infarction patients, but in contrast to the endogenously occurring myocardial infarction pathology in STEMI patients subjected to percutaneous coronary intervention, the onset of inflammation and oxidative injury in cardiac surgical patients is predictable as cardiac surgery with the cardioplegia-induced myocardial arrest, assistance of cardiopulmonary bypass and the following myocardial reperfusion, reproducibly elicits an ischaemia–reperfusion sequelae. Intriguingly, the increase in peri-operative MIF levels in cardiac surgery patients, as well as ratios of MIF and its soluble receptor CD74 (sCD74), suggest a cardioprotective role of MIF in the ischaemic and early reperfusion phase after myocardial infarction.[54] [55] [56] In fact, cardioprotection by MIF in MI/RI is confirmed in numerous mouse models.[19] [57] [58] [59] [60] [61] [62] Along the same lines, the myocardium-specific conditional knockout of D-dt/Mif-2 exacerbates MI/RI, while MIF-2 levels positively correlated with worse outcome in cardiac surgery patients.[54] [63] Interestingly, experimental models addressing the later post-ischaemic phase indicated that the role of MIF in cardiac ischaemia is complex, with phase-dependent cardioprotective and exacerbating effects observed.[57] [58] [59] [61] [64] MIF is initially released by ischaemic cardiomyocytes or endothelial cells and triggers a cardioprotective autocrine/paracrine signalling response in cardiomyocytes. Here, MIF not only binds to the chemokine receptors CXCR2 and CXCR4, but also to CD74, the surface-expressed form of the major histocompatibility complex (MHC) class II invariant chain, which serves a secondary function as a high-affinity MIF receptor.[65] Cardiac-derived—first wave—MIF interacts with cardiomyocyte-expressed CD74 in the ischaemic and early reperfusion phase of MI/RI to trigger cardioprotective signalling through adenosine monophosphate (AMP) kinase metabolic reprogramming, an increase in glucose uptake via membrane translocation of GLUT4 and the AKT and extracellular-signal-regulated kinase survival pathways, while pro-apoptotic c-Jun N-terminal kinase signalling is attenuated.[58] [60] MIF-2 also mediates cardioprotection in this phase via CD74/AMP-activated protein kinase (AMPK) signalling.[63] MIF/CD74/AMPK-mediated ischaemic recovery is impaired in the senescent heart, suggesting that this protective mechanism could be dampened in aged CHD patients.[66] MIF's antioxidant capacity that is based on its redox-active CXXC motif and that it shares with thiol-protein oxidoreductases such as thioredoxin[67] also contributes to cardioprotection in the early phase of MI/RI stress.[57] [59] [62] In contrast, MIF's role in the later phase of MI/RI stress in the heart is an ‘inflammatory’ one that is mediated by CXCR2/CXCR4-dependent recruitment of monocytes and neutrophils. Second wave MIF is additionally and abundantly produced by infiltrating inflammatory cells to amplify the inflammatory response.[64] MIF's chemokine receptors serve a dual role in this phase with both protective (cardiomyocyte-expressed CXCR2/4) and pro-inflammatory (CXCR2/4 expressed on infiltrating myeloid cells) activity.[68] [69]


#

MIF Proteins and Their Receptors

MIF binds to CXCR2 and CXCR4, representing non-cognate interactions between an ACK and classical chemokine receptors. MIF also binds to CD74, the surface-expressed form of invariant chain. All three receptors have important roles in atherosclerosis and cardiac disease (see above). Recent evidence also suggests engagement of CXCR7-mediated pathways by MIF.[70] [71] [72]

Binding of MIF to CXCR2 drives atherogenic recruitment of monocytes and neutrophils.[14] [73] [74] Mechanistically, binding of MIF to CXCR2 is similar but not identical to that of the cognate ligand CXCL8 and requires an N-like loop and pseudo-Glu-Leu-Arg (ELR) motif.[16] [74] [75] [76] While data are not yet available, it has been speculated that MIF-2 does not activate CXCR2 as it lacks the pseudo-(E)LR motif of MIF. MIF/CXCR4 binding supports the recruitment of atherogenic T cells,[14] but has also been implicated in cancer metastasis and endothelial progenitor cell recruitment.[16] Recent evidence suggests an important role of the MIF/CXCR4 axis in B cell migration that may also contribute to the pro-atherogenic phenotype of MIF.[70] [77] [78] CXCR4 is one of the few GPCRs for which an X-ray structure has been elucidated,[79] [80] and recent structure-activity studies (SAR) revealed that the MIF/CXCR4 interface involves an extended N-like loop of MIF, an Arg-Leu-Arg (RLR) motif at position 87–89 and the N-terminal Pro-2.[81] [82]

Interestingly, both CXCR2 and CXCR4 are able to form receptor complexes with CD74, offering unexpected mechanistic options as to the fine-tuning of MIF-driven pathways in atherogenesis. In fact, MIF-mediated CXCR2/CD74 signalling has a role in atherogenic leukocyte recruitment,[14] [83] while CXCR4/CD74 complex formation and/or cross-talk is necessary for MIF-driven B cell migration responses and elicits downstream ZAP-70 signalling.[77] [84]

Intracellular CD74/invariant chain acts as an MHC II chaperone facilitating antigen loading to class II complexes in the endoplasmic reticulum.[85] However, CD74 may be also expressed in class II-negative cells, that is, upon inflammatory stimulation, and exhibits a major role as cytokine receptor for MIF and MIF-2.[65] [86] MIF and MIF-2 bind to the extracellular domain of CD74, but as CD74 exhibits a short cytoplasmic domain, signal transduction necessitates accessory molecules such as CD44 or CXCR2/4.[14] [16] [65] [87] A soluble form of CD74 (sCD74) was identified in patients with autoimmune liver disease,[88] and as outlined above, evidence from cardiac surgery patients suggests that circulating sCD74 levels correlate with better outcome.[54] [56] sCD74-derived strategies could thus represent interesting MIF-targeting approaches in the future.


#

Antibody-Based Anti-MIF Strategies

Antibody-based anti-cytokine/-chemokine strategies are promising therapeutic approaches in inflammatory and cardiovascular diseases. Prominent examples are the IL-1β antibody canakinumab, which was shown in the CANTOS trial to reduce vascular inflammation accompanied by a lower rate of recurrent cardiovascular events,[12] the anti-TNF-α antibody infliximab, successfully used in RA,[89] and numerous chemokine antibodies such as anti-CCL2 (CNTO888/ABN912), which are in phase 1 and 2 clinical studies for various inflammatory conditions and cancer.[90] [91] Chemokine antibodies such as anti-CCL2 have been efficacious in pre-clinical models of atherosclerosis,[7] suggesting their potential in CAD patients.

Neutralization of MIF by blocking antibodies has improved disease exacerbation in numerous pre-clinical inflammation models including atherosclerosis.[18] [37] [92] [93] The most widely used antibody has been the monoclonal NIH/IIID.9, which was raised against full-length mouse MIF. It was initially tested in immunologically induced kidney disease[94] and has been demonstrated to potently block disease progression in numerous inflammatory, autoimmune and cardiovascular conditions.[14] [16] [18] [19] [37] [95] The epitope recognized by NIH/IIID.9 has not been characterized, but it has been suggested that it recognizes a solvent-exposed region of MIF in the middle part of the sequence,[95] similar to mAb clone 1C10 (Bernhagen et al, unpublished), but unlike clone F11, which blocks cecal ligation and puncture-induced sepsis and is directed against the N-terminal of murine MIF.[96]

From the existing anti-MIF antibodies successfully tested in pre-clinical inflammation, cancer and atherosclerosis models, only the MIF antibody imalumab has so far advanced into phase 1/2a clinical trials (NCT01765790) against colorectal cancer and lupus nephritis.[97] Imalumab has an anti-inflammatory capacity as it reduces circulating TNF-α, monocyte chemoattractant protein-1 and IL-6, and attenuates disease progression in mouse models of glomerulonephritis and cancer. Based on biochemical and immunochemical experiments, it was suggested that this antibody recognizes an oxidized form of MIF with the oxidoreductase motif of MIF trapped in an oxidized state, and that Cys-81 serves as a molecular redox switch between the latent reduced form of MIF and its oxidized state.[93] [98] [99] While the antibody was reported to detect an oxidized MIF species termed ‘oxMIF’ in tumour tissue from patients with colorectal, pancreatic, ovarian and lung cancer,[100] the claim that oxMIF is the pathophysiologically relevant MIF species appears speculative and convincing evidence that imalumab targets pathogenic oxMIF species is missing.[67] [95]

MIF-2/D-DT shares with MIF a pronounced pro-inflammatory activity profile and has been reported to promote endotoxaemia, adipose tissue inflammation and tumorigenesis similar to MIF. As discussed, MIF-2 also is involved in early cardioprotection after MI/RI and regulates kidney regeneration, but its role in atherosclerosis has not been studied.[19] [63] [86] [101] [102] [103] Functional studies on MIF-2 have capitalized on Mif-2 gene deletion[63] and a neutralizing antibody.[86] While this polyclonal antibody shows good blocking potency in pre-clinical models,[86] a monoclonal antibody has not been published.

Cytokine/chemokine pathways may be targeted by neutralization of the ligand or may rely on strategies to block the receptor binding site of the ligand and/or receptor signalling. Antibodies against CD74 recapitulate many of the effects seen with neutralizing MIF antibodies, but differences have also been noted. As both MIF and MIF-2 bind to CD74, these may be due to MIF-2-triggered responses that are inhibited by anti-CD74 but not anti-MIF strategies. No other endogenous ligands than MIF or MIF-2 have been identified for CD74. Yet, the blocking phenotype of anti-CD74 may differ from a combined anti-MIF/anti-MIF-2 strategy due to class II-associated functions of endolysosomal-expressed CD74/invariant chain. In fact, specific peptide-based strategies have been developed to block class II-associated functions of CD74 in multiple sclerosis[104] [105] [106] [107] (see below).

A neutralizing CD74 mAb has shown potent inhibitory activity in haematologic cancers such as chronic lymphocytic leukaemia (CLL) and multiple myeloma.[108] [109] One of these mAb clones, a humanized anti-CD74 monoclonal termed milatuzumab/hLL1, is in clinical trials for CLL and multiple myeloma treatments. However, although gene deletion of CD74 significantly attenuates atherosclerosis in atherogenic Ldlr/− mice[110] and although CD74 serves as an accessory molecule in MIF-driven CXCR2/4-mediated leukocyte recruitment responses,[14] neutralizing CD74 antibodies have not been studied in atherosclerosis. It should be emphasized that anti-CD74 strategies may be intrinsically limited regarding their translational potential in atherosclerotic disease due to cardiac protection mediated via the CD74/AMPK pathway in cardiomyocytes following ischaemic stress.[58] It rather seems that MIF-based strategies in cardiovascular disease should aim at sparing CD74-mediated pathways.[19]

CXCR2 and CXCR4 are bona fide GPCRs, but antibody development against GPCRs has been delayed. In fact, of the numerous GPCR-modulating agents available, most are small molecules or peptides. Recently, the first GPCR-directed antibody (erenumab), an antibody against calcitonin gene-related peptide receptor, was Food and Drug Administration (FDA)-approved for treatment of migraine.[111] Meanwhile, the development of additional anti-GPCR antibodies including antibodies against CC and CXC chemokine receptors is underway.[111]

CXCR4 antibodies are in advanced clinical trials for haematologic malignancies,[112] and anti-CXCR4 SMDs such as the bicyclam plerixafor/AMD3100 which promotes CXCR4-dependent haematopoietic stem cell egress from bone marrow is clinically used in autologous stem cell transplantation of cancer patients.[113] Nevertheless, anti-CXCR4 antibody strategies as a means to block MIF-driven pathogenic pathways in atherosclerosis should be pursued with caution. The CXCR4/CXCL12 axis has important homeostatic functions in development and physiology that render generalized anti-CXCR4 strategies difficult. Moreover, disrupting the CXCL12/CXCR4 axis in a mouse model of atherosclerosis promoted lesion formation through dysbalanced neutrophil homeostasis,[114] and a recent study demonstrated a potent atheroprotective effect of vascular CXCR4 via maintaining arterial integrity, endothelial barrier function and preserving contractile VSMC functions.[115]

Cxcr2 gene deficiency reduces the progression of advanced atherosclerosis in mice and, in fact, CXCR2 has been one of the first chemokine receptors implicated in atherogenesis.[116] [117] CXCR2 also is involved in MIF-elicited atherogenic monocyte and neutrophil recruitment,[14] [16] [75] emphasizing the significance of the MIF/CXCR2 axis in leukocyte arrest and atherogenesis. Moreover, anti-MIF antibodies proved superior to anti-CXCL1 (and anti-CXCL12) in an atherosclerosis regression model.[14] An anti-CXCR2 antibody therapy is considered a translatable strategy in solid cancer, for example, through improving the efficacy of checkpoint blockade by preventing trafficking of myeloid-derived suppressor cells to the tumour site[118] and anti-CXCR2 strategies are pursued in various clinical trials,[119] including ACS.[120] An anti-CXCR2 bi-paratopic nanobody is in phase I development for the treatment of inflammation.[111] Antibody strategies for epitopes specifically targeting MIF/CXCR2 receptor pathways in vascular inflammation and atherosclerosis have not been pursued.

[Table 1] summarizes published antibody-based strategies directed at MIF proteins and/or their receptors.

Table 1

Antibodies targeting MIF proteins or their receptors

Antibody

Target/Antigen

Application/Utility in atherosclerosis

References

NIH/IIID.9 (mAb)

Mouse MIF (full-length)

Research and pre-clinical models; blocks atherogenic effects of MIF

[14] [37] [94]

Imalumab (Bax69) (humanized mAb)

Oxidized form of human MIF

Phase IIa trial for metastatic colorectal cancer

[97]

BaxB01, BaxG03, BaxM159

Oxidized form of human, mouse or rat MIF

Research and pre-clinical models

[93] [98] [99] [100] [163]

NbE10-NbAlb8-NbE10 (half-life-extended nanobody)

Human and mouse MIF

Research and pre-clinical sepsis model

[164]

Anti-MIF-2/D-DT

Mouse MIF-2/D-DT (full-length)

Research and pre-clinical models

[86]

Milatuzumab

CD74

Multiple myeloma, NHL, CLL

[108] [165]

i-bodies (AM3–114, AM4–272, AM3–523; single domain antibody)

CXCR4

Research and pre-clinical models

[166]

MEDI3185

CXCR4

Research and pre-clinical models

[112]

Anti-CXCR2 bi-paratopic nanobody

CXCR2

Pre-clinical models and phase I clinical trial

[111]

MAB331

CXCR2

Research and pre-clinical models

[14] [167]

*sCD74

Human and mouse MIF

Research and pre-clinical models

[56] [65] [88]

Abbreviations: CLL, chronic lymphocytic leukaemia; D-DT, D-dopachrome tautomerase; MIF, migration inhibitory factor; NHL, non-Hodgkin lymphoma.


Note: *sCD74 is not an antibody, but the soluble ectodomain of MIF receptor CD74.



#

Small Molecule Drug-Based Anti-MIF Strategies

MIF proteins are structurally unique among cytokines/chemokines in harbouring a conserved catalytic tautomerase cavity that contains the unusually acidic Pro-2 residue. This offers the opportunity to target pathogenic activities of MIF by small molecule approaches. Capitalizing on efficient drug discovery pipelines including in silico and high-throughput screening, numerous anti-MIF SMDs have been identified that bind into or modulate the tautomerase pocket of MIF and/or MIF-2/D-DT by covalent or non-covalent mode.

Small molecule MIF inhibitors have been compiled in several recent review articles.[24] [95] [121] [122] Here, we discuss some of these compounds with a focus on their potential utility in atherosclerosis. [Table 2] summarizes the key features of these inhibitors.

Table 2

Small molecules targeting MIF proteins or their receptors

Small molecule drug (SMD) inhibitor

Target/Binding mode

Ki

IC50/EC50

References

NAPQI

(N-acetyl-p-benzoquinone imine)

MIF

Pro-2

Covalent

N/A

IC50 (dopachrome) = 40 µM

[131] [168]

4-IPP

(4-iodo-6-phenylpyrimidine)

MIF/D-DT

Pro-2

Covalent

N/A

IC50 (HPP) = 0.2–0.5 µM

[169] [170]

ISO-1

(4,5-Dihydro-3-(4-hydroxyphenyl)-5-isoxazoleacetic acid methyl ester)

MIF

Tautomerase site

Competitive

Ki (HPP) =

24 µM

IC50 (dopachrome) = 7 µM

[130] [171]

SCD-19

isocoumarin

MIF

Tautomerase site

Competitive

Not tested

100% inhibition at 100 µM

[172] [173]

4-CPPC

(4-(3-Carboxyphenyl)-2,5-pyridinedicarboxylic acid)

MIF-2/D-DT

C-terminus

V114 − L118

Ki (HPP) =

33 ± 0.7 μM

[125]

Ebselen

(2-Phenyl-1,2-benzisoselenazol-3(2H)-one)

MIF trimer

Cys-81

Covalent

Ki (HPP) = 0.57 μM

IC50 (dopachrome) = 2.4 µM

[123]

p425

6,6'-[(3,3-Dimethoxy[1,1'-biphenyl]-4,4'-diyl)bis(azo)]bis[4-amino-5-hydroxy-1,3-napthalenedisulphonic acid]

MIF trimer

Allosteric

Ki (HPP) ≤ 12 μM

IC50 (CD74 inhibition) = 0.81 µM

[124]

Ibudilast

AV411; 3-isobutyryl-2-isopropylpyrazolo-[1,5-a]pyridine

MIF

Tyr-37

Allosteric

Ki (HPP) = 30.9 μM

[73] [132] [134]

Plerixafor/AMD 3100

(1-[4-(1,4,8,11-Tetrazacyclotetradec-1-ylmethyl)phenyl]methyl)-1,4,8,11-tetrazacyclo-tetradecan

CXCL12/CXCR4

Orthosteric antagonist

MIF/CXCR4

Partial allosteric antagonist

IC50 (CXCR4) = 0.65 µM

EC50 (HIV entry) = 0.4–2 µM

[80] [82] [156] [174]

IT1t

Isothiourea-1t

6,6-dimethyl-5H-imidazo[2,1-b][1,3]thiazol-3-yl)methyl N,N'-dicyclohexylcarbamimidothioate

CXCL12/CXCR4

Orthosteric antagonist

MIF/CXCR4

Partial allosteric antagonist

IC50 (gp120 inhibition) = 8 nM

[80] [82] [174]

Reparixin

(αR)-α-methyl-4-(2-methylpropyl)-N-(methylsulfonyl)benzeneacetamide

CXCL8/CXCR2

Allosteric antagonist

MIF/CXCR2 (?)

IC50 (neutrophil migration) = 1 nM

[174] [175]

Abbreviations: D-DT, D-dopachrome tautomerase; HIV, human immunodeficiency virus; HPP, hydroxyphenylpyruvate; MIF, migration inhibitory factor.


Mechanistically, small molecule MIF inhibitors are classified into different categories: (1) competitive inhibitors that non-covalently bind into the cavity; (2) suicide inhibitors that covalently bind into the cavity; (3) allosteric inhibitors that disrupt the active-site through dissociation of the MIF trimer or an otherwise-induced conformational switch; and (4) allosteric inhibitors that prevent higher-order MIF oligomers.[123] [124] In addition, stabilizers of the MIF monomer have been proposed as inhibitors through prevention of re-association into trimer.[123] These may only qualify as inhibitors of MIF/CD74 interactions, which involve trimeric MIF, whereas it has been assumed that the interaction between MIF and CXCR2/CXCR4 is a function of the monomer.[14] While most inhibitors have been developed against MIF, some also block MIF-2, although significant differences in Ki and IC50 values have been noted.[125] A recent study has identified a selective MIF-2 inhibitor that exhibits 13-fold higher binding to MIF-2 than MIF.[125]

The MIF tautomerase activity is highly conserved across kingdoms, but to date, physiological substrates in mammalians have not been identified, raising the possibility that it is an evolutionary remainder with no function in humans. Thus, it is now thought that by binding to the tautomerase site, these compounds induce conformational changes in MIF that subsequently alter its receptor-binding properties,[82] [126] [127] [128] [129] and therefore ‘indirectly’ influence MIF activities.

The structures of the small molecule MIF inhibitors have been extensively reviewed.[121] [122] Briefly, with the iso-oxazoline compound ISO-1 serving as a reference MIF inhibitor in various inflammation models,[130] they can be grouped into iso-oxazolines (examples: ISO-1, ISO-66, CPSI-1306), chromenes (examples: Orita-13, Kok-17), iminoquinones (example: N-acetyl-p-benzoquinone imine), triazoles (example: Cisneros-3i), benzoxazolones (example: MIF098), pyrimidazoles (example: K664–1) and isocoumarins (e.g. SCD-19).[121] [122] [131] Allosteric MIF inhibitors encompass pyrazolopyridines (example: clinically used PDE4 inhibitor ibudilast), benzoisoselenazolones (example: anti-inflammatory drug ebselen) and azo compounds (example: p425 or Chicago Sky Blue 6b).[73] [123] [124] Of note, ibudilast, which can cross the blood–brain barrier, was recently demonstrated in a phase II clinical trial in progressive multiple sclerosis, to be associated with slower progression of brain atrophy than placebo.[132] Isothiocyanates, such as phenethyl isothiocyanate, are a reactive class of ‘natural’ MIF inhibitors that are present in appreciable amounts in broccoli and water cress. They covalently bind to the acidic Pro-2 residue in the catalytic site of MIF, attenuate MIF antibody binding and inhibit inflammatory MIF activities in vitro, but have not yet been studied in atherosclerosis.[123] [133]

While most of these compounds have not yet been studied in atherosclerosis-relevant test systems or pre-clinical models, some of them may hold promise as pathway-specific MIF inhibitors in atherogenesis. However, a ‘black-and-white’ categorization into receptor-specific blockers appears too simplistic. For example, CD74 is a receptor mediating cardio- and tissue-protective MIF activities[19] [58]; on the other hand, it drives pro-proliferative functions of MIF and can interact with the MIF chemokine receptors, which would be pro-atherogenic.[14] [70] [109] CXCR4 mediates pro-atherogenic lymphocyte recruitment by MIF,[14] [78] but also exhibits homeostatic and atheroprotective activities through CXCL12.[115] Notwithstanding, it may be speculated that MIF trimerization inhibitors such as ebselen might have specificity for MIF/chemokine receptor pathways, while sparing MIF trimer-dependent CD74 signalling,[123] which could lead to an overall atheroprotective effect. Ibudilast was found to block MIF/CD74 interactions in vivo by preventing astrocyte-derived MIF from interacting with CD74+ microglia during the colonization process of brain metastatic tumours resulting in reduced secondary brain tumour loads.[134] Similarly, ISO-1 reduces MIF/CD74 binding albeit at relatively weak IC50 values,[121] but interestingly, was recently also shown to partially interfere with MIF binding to CXCR4, opening up the possibility that certain MIF inhibitors may have utility in cardiovascular disease. However, given their small interaction surface, it remains questionable whether they would sufficiently differentiate between receptor pathways.

[Table 2] summarizes a selection of the developed small molecule MIF inhibitors from different structural classes and compares them to MIF chemokine receptor inhibitors. To this end, AMD3100/Plerixafor and reparixin are established CXCR4 and CXCR2 inhibitors, respectively. AMD3100 also was shown to be a partial—allosteric—inhibitor of MIF/CXCR4 signalling[82]; however, its preferential targeting of CXCL12/CXCR4 responses and its limited application window in autologous transplantation and human immunodeficiency virus (HIV), probably limits an efficacious usage as a specific anti-MIF compound. Reparixin has so far only been studied in MIF-dependent in vitro inflammatory assays.[135]


#

Peptide-Based Anti-MIF Strategies

Peptide therapeutics are a powerful alternative to small molecule and antibody strategies with over 60 peptide drugs approved worldwide. The peptide therapeutic landscape has been reviewed in excellent recent reviews.[136] [137] [138] Advantages of peptide-based inhibitors are: (1) good selectivity and potency, (2) good interaction surface coverage, (3) favourable safety and (4) comparatively low production costs due to standard synthesis protocols. On the other hand, they are prone to degradation and oxidation, but this can be improved by smart mimic chemistry.[136] [139]

Even though peptide-based inhibitors have been pursued as potential therapeutics in atherosclerosis, for example, targeting lipid-regulating and inflammatory pathways such as apolipoproteins, nuclear factor-kappaB and the IL-4 receptor,[140] [141] [142] peptide approaches targeting MIF-specific pathways are still in its infancies. Interestingly, a recent approach established designed peptide inhibitors that specifically disrupt pro-inflammatory CCL5–CXCL4 interactions, attenuating monocyte recruitment and reducing atherosclerosis. Targeting of CCL5–CXCL4 heteromers avoids side effects of generalized anti-CCL5 strategies which would compromise systemic host immunity,[143] [144] thus underscoring the potential of anti-chemokine peptide strategies.

Anti-MIF peptides have been examined in vitro and partially in pre-clinical disease models. Peptides targeting both the MIF/CD74 and MIF/chemokine receptor axes have been considered ([Table 3]); some of them are derived from mapping studies of the interfaces between MIF and its receptors.

Table 3

Peptides and peptide mimics targeting MIF proteins or their receptors

Peptide inhibitor

Target/Binding mode

Application/Utility in atherosclerosis (IC50)

References

MIF[79] [80] [81] [82] [83] [84] [85] [86] (mouse)

LCGLLSDR

MIF/CD74 interface

IC50 = ca. 2–3 µM

[88]

MIF[47] [48] [49] [50] [51] [52] [53] [54] [55] [56] (human)

LMAFGGSSEP

MIF

Competitive

EC50 = ca. 1–2 µM

[76]

MIF[50] [51] [52] [53] [54] [55] [56] [57] [58] [59] [60] [61] [62] [63] [64] [65] (human)

FGGSSEPCALCSLHSI

MIF

Competitive

Not determined

[176] [177]

Conserved CDR peptide C36L1

KSSQSVFYSSNNKNYLA-NH2

CD74

IC50 = upper µM range

[146]

RTL1000

Class II-derived

Drα1β1MEVGWYRSPFSRVVHLYRNGK

CD74 trimer

MIF/CD74 axis

Competitive

IC50 = nanomolar range

[107]

DRα1-MOG-35–55

Class II-derived

DRα1MEVGWYRSPFSRVVHLYRNGK

CD74 trimer

MIF/CD74 axis

Competitive

IC50 = nanomolar range

[104] [107]

CXCL12(22–29)2

KGVSLYR-K-RYSLVGK

CXCL12/CXCR4 axis

[178]

CXCL12a[1] [2] [3] [4] [5] [6] [7] [8] [9] [P2G] dimer

MNAKVVVVL-S-S-LVVVVKANM

CXCL12/CXCR4 axis

IC50 = 2.6 µM

[178]

Ac-Arg-Ala-[D-Cys-Arg-Phe-His-Pen]-COOH

Derivative of CXCL12 N-terminal

CXCL12/CXCR4 axis

IC50 = 1.5 nM

[158] [159]

CVX15

16-residue cyclic peptide analogue of the horseshoe crab peptide polyphemusin

CXCR4

Not known

[80]

MCoTI-based cyclotides

CXCR4

IC50 = 20 nM

EC50 (HIV entry) = 2 nM

[179]

Peptides T22 and T140

Polyphemusin II-related synthetic 14–16-meric derivatives

CXCR4

IC50 = 17 nM

[153] [154]

Cyclopentapeptide FC131

Head-to-tail-cyclized variant of T140

CXCR4

IC50 = 8 nM

[154]

Peptoid 8

Peptoid derivative of FC131

CXCR4

IC50 = 40 pm

EC50 (HIV entry) = 29 nM

[156]

Abbreviations: CDR, complementarity-determining region; HIV, human immunodeficiency virus; MIF, migration inhibitory factor; MOG, myelin oligodendrocyte glycoprotein.


Note: IC50 refers to replacement of ligand (MIF, CXCL12) from receptor (CD74, CXCR4).


MIF-derived peptides targeting interactions with CD74 have not yet been tested in disease models. This is probably due to the complex nature of the MIF/CD74 interaction surface, encompassing all three subunits of the MIF trimer and discontinuous epitopes within an MIF monomer. Nevertheless, a MIF epitope scan for reactivity of the CD74 ectodomain identified MIF peptide 79–86. This octapeptide was able to compete for biotinylated MIF binding to plate-bound CD74,[88] indicating its principal inhibitory utility. A screening for anti-melanoma peptides derived from conserved complementarity-determining region sequences of different immunoglobulins identified peptide C36L1, a 17-mer peptide that binds to CD74 on tumour-associated macrophages and dendritic cells and blocks immunosuppressive activities in melanoma models.[145] [146] Other inhibitors that target the MIF–CD74 interaction have been developed based on antigenic peptide-loaded fragments of class II resulting in reduced severity of experimental autoimmune encephalomyelitis, the experimental mouse model of multiple sclerosis.[104] [105] [107] One such peptide is the DR2-restricted myelin determinant mouse (m) myelin oligodendrocyte glycoprotein (MOG)-35–55 covalently linked to a human leukocyte antigen-DRα1 domain (the ‘DRα1-MOG-35–55’ construct) and has been found to reduce central nervous system inflammation and tissue injury in models of multiple sclerosis, ischaemic stroke and traumatic brain injury.[104] [105] [107] [147] RTL1000 is a variant of this construct additionally containing the β1 domain of DR1 (‘DRα1β1-MOG-35–55’) and is in clinical studies for multiple sclerosis.[148] [149] It is thought that these peptide constructs interfere with MIF/CD74-driven neuroinflammation.[150]

A note of caution should be sounded regarding their potential application in atherosclerosis and cardiovascular disease settings due to the protective role of the MIF/CD74 axis in ischaemic heart disease.[19] [58]

As discussed above, SARs have identified the motifs and residues contributing to the interface between MIF and CXCR2/4 and highlighted differences compared with the classical chemokine ligands CXCL8/1 and CXCL12, respectively. The two-site binding mechanism for MIF and CXCR2 is similar but not identical to that for CXCL1/8, while significant differences were noted for the binding interface of MIF/CXCR4 versus CXCL12/CXCR4.[14] [16] [74] [75] [76] [81] [82] Peptides served as important tools in these studies and some of them might become templates for peptide-based anti-MIF strategies in atherosclerosis and inflammation. For example, MIF peptide 47–56, spanning the N-like loop that contributes to site 1 binding with CXCR2, competes with MIF binding to CXCR2 and MIF-mediated atherogenic leukocyte arrest.[16] [76] Stabilized variants of this peptide or related ones spanning MIF regions contributing to site 1 or 2 binding might qualify as interesting templates for the future development of MIF-based peptide drugs against atherogenic inflammation.

MIF acts as a partial allosteric agonist of CXCL12, consistent with the notion that the binding interface between MIF and CXCR4 differs from that of CXCL12 and CXCR4.[82] Major differences are the contributions of the extended N-like loop and the cavity around Pro-2 in MIF[81] [82] and the RFFESH motif in CXCL12.[80] [151] This SAR information as well as the crystal structures of MIF and of CXCR4 in complex with small molecules and cyclic peptides give valuable hints as to the development of peptide inhibitors that may specifically block MIF-driven responses.[80]

Of note, CXCR4 pathways have already been targeted by peptide inhibitor strategies. CVX15, a 16-residue cyclic peptide analogue of the horseshoe crab peptide polyphemusin, was co-crystallized with CXCR4[80] and characterized as an HIV-inhibiting and anti-metastatic agent.[152] [153] Polyphemusin II-related synthetic peptides T22, T140 and FC131 were pioneered by Fujii et al to adopt a β-hairpin conformation stabilized by disulphide bonds, resulting in high-affinity CXCR4 inhibitory peptides with a low nanomolar IC50.[153] [154] [155] This principle was developed further by Kessler and colleagues, who furthered the principle of protein–epitope mimetics and devised novel classes of super high-affinity CXCR4-targeting cyclopeptides,[156] [157] for example, by ‘freezing’ the conformation of a CXCR4 ligand into a single-active conformation by using a ‘peptoid’ motif.[156] In another approach, peptide inhibitors were derived from the N-terminal of CXCL12 and further optimized and stabilized to give rise to sub-nanomolar serum-stable CXCL12/CXCR4 inhibitors with anti-metastatic activity in vitro.[158] [159] Furthermore, peptides based on the sequence of CXCR4 were linked in an attempt to mimic the ecto-surface of CXCR4 and shown to compete with HIV-gp120 and HIV entry.[160] [161] MIF is not able to inhibit HIV entry[82] and peptides targeting the MIF/CXCR4 axis have not been systematically studied, although a peptide spanning the RLR motif of MIF competes with MIF-mediated lymphocyte migration.[81]


#

Conclusion

MIF is a pivotal mediator of atherosclerosis, MIF-2/D-DT shares critical inflammatory activities with MIF and the MIF receptors CD74, CXCR2 and CXCR4 have all been implicated in atherosclerosis, suggesting that it will be important to develop therapeutic strategies against MIF proteins in atherosclerosis. Importantly, the MIF network is amenable to targeting by all major inhibitor classes, that is, small molecule compounds, antibodies and peptides ([Fig. 1]). In fact, inhibitors against MIF and/or its receptors from all three classes are in clinical development and the CXCR4 blocker AMD3100, which is a partial inhibitor of MIF/CXCR4 binding, is an FDA-approved drug in cancer. However, to make such strategies applicable for cardiovascular disease, MIF pathway-specific concepts need to be developed that specifically target the atheroprogressive activities of MIF.

Zoom Image
Fig. 1 Overview of inhibitory approaches to target the macrophage migration inhibitory factor (MIF)/receptor network in atherosclerosis. The potential utility of all three classes of anti-MIF network inhibitors, that is, antibodies, small molecule drug (SMD) compounds and peptides, in attenuating atherosclerosis and/or atherogenic inflammation is indicated with respect to MIF and MIF-2/D-dopachrome tautomerase (D-DT), as well as the MIF receptors CXCR4, CD74 and CXCR2. The pros and cons for each inhibitor-type regarding each target ligand/receptor or pathway are indicated by scoring their properties (e.g. specificity) with +, +/– or –. The outcome boxes are colour-coded (red, pro-atherogenic; green, athero-/cardioprotective).

Considering the Janus-faced effects of MIF proteins in cardiovascular complications and the complex homeostatic and inflammatory roles of their receptors, this is a challenging task. One strategy might be to specifically target the MIF/CXCR2 interface which is atheropromoting and largely detrimental in the myocardium during an ischaemic insult. Both antibodies and peptide-based compounds such as stabilized derivatives of MIF[47] [48] [49] [50] [51] [52] [53] [54] [55] [56] could potentially qualify as MIF/CXCR2-specific agents. Similarly, inhibitor strategies specifically targeting the MIF/CXCR4 interaction could be envisioned, although great caution would need to be taken to spare the various protective activities of CXCR4 in the atherogenic vasculature and the ischaemic-stressed heart. On the other hand, the cardioprotective effect of MIF and MIF-2 observed in the early phase following MI/RI provides a narrow albeit critical therapeutic window to pharmacologically promote the cardioprotective function of MIF before late-phase inflammatory responses kick in. This could especially be relevant in cardiac surgery patients and one relevant agent is the small molecule agonist MIF-20, which binds near MIF's tautomerase pocket and has been reported to have protective effects in an experimental model of cardiac ischaemic injury.[162] Such a ‘pharmacological augmentation’ strategy would be selective to MIF proteins due to their structurally unique tautomerase cavity and might become particularly important in cohorts of patients identified as MIF ‘low expressers.’[19] [58] Therapeutically, a combinatorial treatment approach might be considered, in which the cardioprotective effect of MIF in the early phase of MI/RI is carefully and phase-specifically enhanced, followed by phase-specific pharmacological inhibition in the late—inflammatory—phase of MI/RI. Whether any anti-MIF strategy would qualify as a treatment regimen to ‘prevent’ or ‘reverse’ chronic atherogenesis similar to canakinumab will have to be subject to comprehensive future investigations. Reversal of atherosclerotic lesions as observed in an experimental mouse model of plaque regression applying anti-MIF but not anti-CXCL12 or anti-CXCL1 antibodies is a promising start in this direction.[14]

In conclusion, for applications in atherosclerotic cardiovascular disease, MIF pathway-specific concepts would need to (1) specifically target the atheroprogressive activities of MIF, (2) preserve homeostatic effects of intracellular MIF, (3) take into account the cardioprotective functions of CD74, and/or (4) spare CXCL12/CXCR4-dependent vascular protection pathways. Molecular characteristics of such agents would need to account for the necessities of chronic treatment over the course of lesion development and/or phase-specificity in the sequelae related to acute cardiac ischaemia.


#
#

Conflict of Interest

J.B. is a co-inventor of patents covering anti-MIF strategies in inflammatory and cardiovascular diseases. Other authors declare no additional conflict of interest.

* Contributed equally.


** Shared last authorship and correspondence.


  • References

  • 1 Ross R. Atherosclerosis--an inflammatory disease. N Engl J Med 1999; 340 (02) 115-126
  • 2 Libby P, Ridker PM, Maseri A. Inflammation and atherosclerosis. Circulation 2002; 105 (09) 1135-1143
  • 3 Dahlöf B. Cardiovascular disease risk factors: epidemiology and risk assessment. Am J Cardiol 2010; 105 (1, Suppl): 3A-9A
  • 4 van der Vorst EP, Döring Y, Weber C. Chemokines and their receptors in atherosclerosis. J Mol Med (Berl) 2015; 93 (09) 963-971
  • 5 Charo IF, Ransohoff RM. The many roles of chemokines and chemokine receptors in inflammation. N Engl J Med 2006; 354 (06) 610-621
  • 6 Bachelerie F, Ben-Baruch A, Burkhardt AM. , et al. International Union of Basic and Clinical Pharmacology. [corrected]. LXXXIX. Update on the extended family of chemokine receptors and introducing a new nomenclature for atypical chemokine receptors. Pharmacol Rev 2013; 66 (01) 1-79
  • 7 Weber C, Noels H. Atherosclerosis: current pathogenesis and therapeutic options. Nat Med 2011; 17 (11) 1410-1422
  • 8 Bäck M, Hansson GK. Anti-inflammatory therapies for atherosclerosis. Nat Rev Cardiol 2015; 12 (04) 199-211
  • 9 Ridker PM, Lüscher TF. Anti-inflammatory therapies for cardiovascular disease. Eur Heart J 2014; 35 (27) 1782-1791
  • 10 Ramji DP, Davies TS. Cytokines in atherosclerosis: key players in all stages of disease and promising therapeutic targets. Cytokine Growth Factor Rev 2015; 26 (06) 673-685
  • 11 Ruparelia N, Chai JT, Fisher EA, Choudhury RP. Inflammatory processes in cardiovascular disease: a route to targeted therapies. Nat Rev Cardiol 2017; 14 (03) 133-144
  • 12 Ridker PM, Everett BM, Thuren T. , et al; CANTOS Trial Group. Antiinflammatory therapy with canakinumab for atherosclerotic disease. N Engl J Med 2017; 377 (12) 1119-1131
  • 13 Bernhagen J, Calandra T, Mitchell RA. , et al. MIF is a pituitary-derived cytokine that potentiates lethal endotoxaemia. Nature 1993; 365 (6448): 756-759
  • 14 Bernhagen J, Krohn R, Lue H. , et al. MIF is a noncognate ligand of CXC chemokine receptors in inflammatory and atherogenic cell recruitment. Nat Med 2007; 13 (05) 587-596
  • 15 Calandra T, Roger T. Macrophage migration inhibitory factor: a regulator of innate immunity. Nat Rev Immunol 2003; 3 (10) 791-800
  • 16 Tillmann S, Bernhagen J, Noels H. Arrest functions of the MIF ligand/receptor axes in atherogenesis. Front Immunol 2013; 4: 115
  • 17 Burger-Kentischer A, Goebel H, Seiler R. , et al. Expression of MIF in different stages of human atherosclerosis. Circulation 2002; 105 (13) 1561-1566
  • 18 Zernecke A, Bernhagen J, Weber C. Macrophage migration inhibitory factor in cardiovascular disease. Circulation 2008; 117 (12) 1594-1602
  • 19 Tilstam PV, Qi D, Leng L, Young L, Bucala R. MIF family cytokines in cardiovascular diseases and prospects for precision-based therapeutics. Expert Opin Ther Targets 2017; 21 (07) 671-683
  • 20 David JR. Delayed hypersensitivity in vitro: its mediation by cell-free substances formed by lymphoid cell-antigen interaction. Proc Natl Acad Sci U S A 1966; 56 (01) 72-77
  • 21 Rich AR, Lewis MR. Migration of neutrophils and macrophages. Bull Johns Hopkins Hosp 1932; 50: 115-131
  • 22 Mischke R, Kleemann R, Brunner H, Bernhagen J. Cross-linking and mutational analysis of the oligomerization state of the cytokine macrophage migration inhibitory factor (MIF). FEBS Lett 1998; 427 (01) 85-90
  • 23 Sun HW, Bernhagen J, Bucala R, Lolis E. Crystal structure at 2.6-A resolution of human macrophage migration inhibitory factor. Proc Natl Acad Sci U S A 1996; 93 (11) 5191-5196
  • 24 Lolis E, Bucala R. Macrophage migration inhibitory factor. Expert Opin Ther Targets 2003; 7 (02) 153-164
  • 25 Flieger O, Engling A, Bucala R, Lue H, Nickel W, Bernhagen J. Regulated secretion of macrophage migration inhibitory factor is mediated by a non-classical pathway involving an ABC transporter. FEBS Lett 2003; 551 (1-3): 78-86
  • 26 Merk M, Baugh J, Zierow S. , et al. The Golgi-associated protein p115 mediates the secretion of macrophage migration inhibitory factor. J Immunol 2009; 182 (11) 6896-6906
  • 27 Hertelendy J, Reumuth G, Simons D. , et al. Macrophage migration inhibitory factor - a favorable marker in inflammatory diseases?. Curr Med Chem 2018; 25 (05) 601-605
  • 28 Morand EF, Leech M, Bernhagen J. MIF: a new cytokine link between rheumatoid arthritis and atherosclerosis. Nat Rev Drug Discov 2006; 5 (05) 399-410
  • 29 Morrison MC, Kleemann R. Role of macrophage migration inhibitory factor in obesity, insulin resistance, type 2 diabetes, and associated hepatic co-morbidities: a comprehensive review of human and rodent studies. Front Immunol 2015; 6: 308
  • 30 Sauler M, Bucala R, Lee PJ. Role of macrophage migration inhibitory factor in age-related lung disease. Am J Physiol Lung Cell Mol Physiol 2015; 309 (01) L1-L10
  • 31 O'Reilly C, Doroudian M, Mawhinney L, Donnelly SC. Targeting MIF in cancer: therapeutic strategies, current developments, and future opportunities. Med Res Rev 2016; 36 (03) 440-460
  • 32 Conroy H, Mawhinney L, Donnelly SC. Inflammation and cancer: macrophage migration inhibitory factor (MIF)--the potential missing link. QJM 2010; 103 (11) 831-836
  • 33 Chesney JA, Mitchell RA, Yaddanapudi K. Myeloid-derived suppressor cells-a new therapeutic target to overcome resistance to cancer immunotherapy. J Leukoc Biol 2017; 102 (03) 727-740
  • 34 Chesney JA, Mitchell RA. 25 years on: a retrospective on migration inhibitory factor in tumor angiogenesis. Mol Med 2015; 21 (Suppl. 01) S19-S24
  • 35 Wirtz TH, Tillmann S, Strüßmann T. , et al. Platelet-derived MIF: a novel platelet chemokine with distinct recruitment properties. Atherosclerosis 2015; 239 (01) 1-10
  • 36 Lin SG, Yu XY, Chen YX. , et al. De novo expression of macrophage migration inhibitory factor in atherogenesis in rabbits. Circ Res 2000; 87 (12) 1202-1208
  • 37 Burger-Kentischer A, Göbel H, Kleemann R. , et al. Reduction of the aortic inflammatory response in spontaneous atherosclerosis by blockade of macrophage migration inhibitory factor (MIF). Atherosclerosis 2006; 184 (01) 28-38
  • 38 Pan JH, Sukhova GK, Yang JT. , et al. Macrophage migration inhibitory factor deficiency impairs atherosclerosis in low-density lipoprotein receptor-deficient mice. Circulation 2004; 109 (25) 3149-3153
  • 39 Amin MA, Haas CS, Zhu K. , et al. Migration inhibitory factor up-regulates vascular cell adhesion molecule-1 and intercellular adhesion molecule-1 via Src, PI3 kinase, and NFkappaB. Blood 2006; 107 (06) 2252-2261
  • 40 Gregory JL, Morand EF, McKeown SJ. , et al. Macrophage migration inhibitory factor induces macrophage recruitment via CC chemokine ligand 2. J Immunol 2006; 177 (11) 8072-8079
  • 41 Atsumi T, Nishihira J, Makita Z, Koike T. Enhancement of oxidised low-density lipoprotein uptake by macrophages in response to macrophage migration inhibitory factor. Cytokine 2000; 12 (10) 1553-1556
  • 42 Kong YZ, Huang XR, Ouyang X. , et al. Evidence for vascular macrophage migration inhibitory factor in destabilization of human atherosclerotic plaques. Cardiovasc Res 2005; 65 (01) 272-282
  • 43 Calandra T, Bernhagen J, Metz CN. , et al. MIF as a glucocorticoid-induced modulator of cytokine production. Nature 1995; 377 (6544): 68-71
  • 44 Luo JY, Xu R, Li XM. , et al. MIF gene polymorphism rs755622 is associated with coronary artery disease and severity of coronary lesions in a Chinese Kazakh population: a case-control study. Medicine (Baltimore) 2016; 95 (04) e2617
  • 45 Herder C, Illig T, Baumert J. , et al. Macrophage migration inhibitory factor (MIF) and risk for coronary heart disease: results from the MONICA/KORA Augsburg case-cohort study, 1984-2002. Atherosclerosis 2008; 200 (02) 380-388
  • 46 Ji K, Wang X, Li J. , et al. Macrophage migration inhibitory factor polymorphism is associated with susceptibility to inflammatory coronary heart disease. BioMed Res Int 2015; 2015: 315174
  • 47 Baugh JA, Chitnis S, Donnelly SC. , et al. A functional promoter polymorphism in the macrophage migration inhibitory factor (MIF) gene associated with disease severity in rheumatoid arthritis. Genes Immun 2002; 3 (03) 170-176
  • 48 Lan MY, Chang YY, Chen WH. , et al. Association between MIF gene polymorphisms and carotid artery atherosclerosis. Biochem Biophys Res Commun 2013; 435 (02) 319-322
  • 49 Müller II, Müller KA, Karathanos A. , et al. Impact of counterbalance between macrophage migration inhibitory factor and its inhibitor Gremlin-1 in patients with coronary artery disease. Atherosclerosis 2014; 237 (02) 426-432
  • 50 Müller II, Müller KAL, Schönleber H. , et al. Macrophage migration inhibitory factor is enhanced in acute coronary syndromes and is associated with the inflammatory response. PLoS One 2012; 7 (06) e38376
  • 51 Deng XN, Wang XY, Yu HY. , et al. Admission macrophage migration inhibitory factor predicts long-term prognosis in patients with ST-elevation myocardial infarction. Eur Heart J Qual Care Clin Outcomes 2018; 4 (03) 208-219
  • 52 Deng F, Zhao Q, Deng Y. , et al. Prognostic significance and dynamic change of plasma macrophage migration inhibitory factor in patients with acute ST-elevation myocardial infarction. Medicine (Baltimore) 2018; 97 (43) e12991
  • 53 Chan W, White DA, Wang XY. , et al. Macrophage migration inhibitory factor for the early prediction of infarct size. J Am Heart Assoc 2013; 2 (05) e000226
  • 54 Stoppe C, Rex S, Goetzenich A. , et al. Interaction of MIF family proteins in myocardial ischemia/reperfusion damage and their influence on clinical outcome of cardiac surgery patients. Antioxid Redox Signal 2015; 23 (11) 865-879
  • 55 Stoppe C, Werker T, Rossaint R. , et al. What is the significance of perioperative release of macrophage migration inhibitory factor in cardiac surgery?. Antioxid Redox Signal 2013; 19 (03) 231-239
  • 56 Stoppe C, Averdunk L, Goetzenich A. , et al. The protective role of macrophage migration inhibitory factor in acute kidney injury after cardiac surgery. Sci Transl Med 2018; 10 (441) eaan4886
  • 57 Luedike P, Hendgen-Cotta UB, Sobierajski J. , et al. Cardioprotection through S-nitros(yl)ation of macrophage migration inhibitory factor. Circulation 2012; 125 (15) 1880-1889
  • 58 Miller EJ, Li J, Leng L. , et al. Macrophage migration inhibitory factor stimulates AMP-activated protein kinase in the ischaemic heart. Nature 2008; 451 (7178): 578-582
  • 59 Pohl J, Hendgen-Cotta UB, Rammos C. , et al. Targeted intracellular accumulation of macrophage migration inhibitory factor in the reperfused heart mediates cardioprotection. Thromb Haemost 2016; 115 (01) 200-212
  • 60 Qi D, Hu X, Wu X. , et al. Cardiac macrophage migration inhibitory factor inhibits JNK pathway activation and injury during ischemia/reperfusion. J Clin Invest 2009; 119 (12) 3807-3816
  • 61 Rassaf T, Weber C, Bernhagen J. Macrophage migration inhibitory factor in myocardial ischaemia/reperfusion injury. Cardiovasc Res 2014; 102 (02) 321-328
  • 62 Koga K, Kenessey A, Powell SR, Sison CP, Miller EJ, Ojamaa K. Macrophage migration inhibitory factor provides cardioprotection during ischemia/reperfusion by reducing oxidative stress. Antioxid Redox Signal 2011; 14 (07) 1191-1202
  • 63 Qi D, Atsina K, Qu L. , et al. The vestigial enzyme D-dopachrome tautomerase protects the heart against ischemic injury. J Clin Invest 2014; 124 (08) 3540-3550
  • 64 Dayawansa NH, Gao X-M, White DA, Dart AM, Du XJ. Role of MIF in myocardial ischaemia and infarction: insight from recent clinical and experimental findings. Clin Sci (Lond) 2014; 127 (03) 149-161
  • 65 Leng L, Metz CN, Fang Y. , et al. MIF signal transduction initiated by binding to CD74. J Exp Med 2003; 197 (11) 1467-1476
  • 66 Ma H, Wang J, Thomas DP. , et al. Impaired macrophage migration inhibitory factor-AMP-activated protein kinase activation and ischemic recovery in the senescent heart. Circulation 2010; 122 (03) 282-292
  • 67 Schindler L, Dickerhof N, Hampton MB. , et al. Post-translational regulation of MIF: basis for functional fine-tuning. Redox Biol 2018; 15: 135-142
  • 68 Liehn EA, Kanzler I, Konschalla S. , et al. Compartmentalized protective and detrimental effects of endogenous MIF mediated by CXCR2 in a mouse model of myocardial ischemia/reperfusion. Arterioscler Thromb Vasc Biol 2013; 33 (09) 2180-2186
  • 69 Liehn EA, Tuchscheerer N, Kanzler I. , et al. Double-edged role of the CXCL12/CXCR4 axis in experimental myocardial infarction. J Am Coll Cardiol 2011; 58 (23) 2415-2423
  • 70 Alampour-Rajabi S, El Bounkari O, Rot A. , et al. MIF interacts with CXCR7 to promote receptor internalization, ERK1/2 and ZAP-70 signaling, and lymphocyte chemotaxis. FASEB J 2015; 29 (11) 4497-4511
  • 71 Chatterjee M, Borst O, Walker B. , et al. Macrophage migration inhibitory factor limits activation-induced apoptosis of platelets via CXCR7-dependent Akt signaling. Circ Res 2014; 115 (11) 939-949
  • 72 Tarnowski M, Grymula K, Liu R. , et al. Macrophage migration inhibitory factor is secreted by rhabdomyosarcoma cells, modulates tumor metastasis by binding to CXCR4 and CXCR7 receptors and inhibits recruitment of cancer-associated fibroblasts. Mol Cancer Res 2010; 8 (10) 1328-1343
  • 73 Cho Y, Crichlow GV, Vermeire JJ. , et al. Allosteric inhibition of MIF revealed by ibudilast. Proc Natl Acad Sci U S A 2010; 107 (25) 11313-11318
  • 74 Xu L, Li Y, Li D. , et al. Exploring the binding mechanisms of MIF to CXCR2 using theoretical approaches. Phys Chem Chem Phys 2015; 17 (05) 3370-3382
  • 75 Weber C, Kraemer S, Drechsler M. , et al. Structural determinants of MIF functions in CXCR2-mediated inflammatory and atherogenic leukocyte recruitment. Proc Natl Acad Sci U S A 2008; 105 (42) 16278-16283
  • 76 Kraemer S, Lue H, Zernecke A. , et al. MIF-chemokine receptor interactions in atherogenesis are dependent on an N-loop-based 2-site binding mechanism. FASEB J 2011; 25 (03) 894-906
  • 77 Klasen C, Ohl K, Sternkopf M. , et al. MIF promotes B cell chemotaxis through the receptors CXCR4 and CD74 and ZAP-70 signaling. J Immunol 2014; 192 (11) 5273-5284
  • 78 Schmitz C, Noels H, El Bounkari O. , et al. Mif-deficiency favors an atheroprotective autoantibody phenotype in atherosclerosis. FASEB J 2018; 32 (08) 4428-4443
  • 79 Qin L, Kufareva I, Holden LG. , et al. Structural biology. Crystal structure of the chemokine receptor CXCR4 in complex with a viral chemokine. Science 2015; 347 (6226): 1117-1122
  • 80 Wu B, Chien EY, Mol CD. , et al. Structures of the CXCR4 chemokine GPCR with small-molecule and cyclic peptide antagonists. Science 2010; 330 (6007): 1066-1071
  • 81 Lacy M, Kontos C, Brandhofer M. , et al. Identification of an Arg-Leu-Arg tripeptide that contributes to the binding interface between the cytokine MIF and the chemokine receptor CXCR4. Sci Rep 2018; 8 (01) 5171
  • 82 Rajasekaran D, Gröning S, Schmitz C. , et al. Macrophage migration inhibitory factor-CXCR4 receptor interactions: evidence for partial allosteric agonism in comparison with CXCL12 chemokine. J Biol Chem 2016; 291 (30) 15881-15895
  • 83 Schwartz V, Lue H, Kraemer S. , et al. A functional heteromeric MIF receptor formed by CD74 and CXCR4. FEBS Lett 2009; 583 (17) 2749-2757
  • 84 Klasen C, Ziehm T, Huber M. , et al. LPS-mediated cell surface expression of CD74 promotes the proliferation of B cells in response to MIF. Cell Signal 2018; 46: 32-42
  • 85 Bertolino P, Rabourdin-Combe C. The MHC class II-associated invariant chain: a molecule with multiple roles in MHC class II biosynthesis and antigen presentation to CD4+ T cells. Crit Rev Immunol 1996; 16 (04) 359-379
  • 86 Merk M, Zierow S, Leng L. , et al. The D-dopachrome tautomerase (DDT) gene product is a cytokine and functional homolog of macrophage migration inhibitory factor (MIF). Proc Natl Acad Sci U S A 2011; 108 (34) E577 –E585
  • 87 Shi X, Leng L, Wang T. , et al. CD44 is the signaling component of the macrophage migration inhibitory factor-CD74 receptor complex. Immunity 2006; 25 (04) 595-606
  • 88 Assis DN, Leng L, Du X. , et al. The role of macrophage migration inhibitory factor in autoimmune liver disease. Hepatology 2014; 59 (02) 580-591
  • 89 Monaco C, Nanchahal J, Taylor P, Feldmann M. Anti-TNF therapy: past, present and future. Int Immunol 2015; 27 (01) 55-62
  • 90 Loberg RD, Ying C, Craig M, Yan L, Snyder LA, Pienta KJ. CCL2 as an important mediator of prostate cancer growth in vivo through the regulation of macrophage infiltration. Neoplasia 2007; 9 (07) 556-562
  • 91 Loberg RD, Ying C, Craig M. , et al. Targeting CCL2 with systemic delivery of neutralizing antibodies induces prostate cancer tumor regression in vivo. Cancer Res 2007; 67 (19) 9417-9424
  • 92 Calandra T, Echtenacher B, Roy DL. , et al. Protection from septic shock by neutralization of macrophage migration inhibitory factor. Nat Med 2000; 6 (02) 164-170
  • 93 Kerschbaumer RJ, Rieger M, Völkel D. , et al. Neutralization of macrophage migration inhibitory factor (MIF) by fully human antibodies correlates with their specificity for the β-sheet structure of MIF. J Biol Chem 2012; 287 (10) 7446-7455
  • 94 Lan HY, Bacher M, Yang N. , et al. The pathogenic role of macrophage migration inhibitory factor in immunologically induced kidney disease in the rat. J Exp Med 1997; 185 (08) 1455-1465
  • 95 Bloom J, Sun S, Al-Abed Y. MIF, a controversial cytokine: a review of structural features, challenges, and opportunities for drug development. Expert Opin Ther Targets 2016; 20 (12) 1463-1475
  • 96 Zhang Y, Zeng X, Chen S. , et al. Characterization, epitope identification and mechanisms of the anti-septic capacity of monoclonal antibodies against macrophage migration inhibitory factor. Int Immunopharmacol 2011; 11 (09) 1333-1340
  • 97 Mahalingam D, Patel M, Sachdev J. , et al. Safety and efficacy analysis of imalumab, an anti-oxidized macrophage migration inhibitory factor (oxMIF) antibody, alone or in combination with 5-fluorouracil/leucovorin (5-FU/LV) or panitumumab, in patients with metastatic colorectal cancer (mCRC). Ann Oncol 2016; 2016: ii105
  • 98 Thiele M, Kerschbaumer RJ, Tam FW. , et al. Selective targeting of a disease-related conformational isoform of MIF ameliorates inflammatory conditions. J Immunol 2015; 195 (05) 2343-2352
  • 99 Schinagl A, Kerschbaumer RJ, Sabarth N. , et al. Role of the cysteine 81 residue of MIF as a molecular redox switch. Biochemistry 2018; 57 (09) 1523-1532
  • 100 Schinagl A, Thiele M, Douillard P. , et al. Oxidized MIF is a potential new tissue marker and drug target in cancer. Oncotarget 2016; 7 (45) 73486-73496
  • 101 Ochi A, Chen D, Schulte W. , et al. MIF-2/D-DT enhances proximal tubular cell regeneration through SLPI- and ATF4-dependent mechanisms. Am J Physiol Renal Physiol 2017; 313 (03) F767-F780
  • 102 Pohl J, Hendgen-Cotta UB, Stock P, Luedike P, Rassaf T. Elevated MIF-2 levels predict mortality in critically ill patients. J Crit Care 2017; 40: 52-57
  • 103 Merk M, Mitchell RA, Endres S, Bucala R. D-dopachrome tautomerase (D-DT or MIF-2): doubling the MIF cytokine family. Cytokine 2012; 59 (01) 10-17
  • 104 Benedek G, Meza-Romero R, Jordan K, Keenlyside L, Offner H, Vandenbark AA. HLA-DRα1-mMOG-35-55 treatment of experimental autoimmune encephalomyelitis reduces CNS inflammation, enhances M2 macrophage frequency, and promotes neuroprotection. J Neuroinflammation 2015; 12: 123
  • 105 Meza-Romero R, Benedek G, Yu X. , et al. HLA-DRα1 constructs block CD74 expression and MIF effects in experimental autoimmune encephalomyelitis. J Immunol 2014; 192 (09) 4164-4173
  • 106 Vandenbark AA, Meza-Romero R, Benedek G. , et al. A novel regulatory pathway for autoimmune disease: binding of partial MHC class II constructs to monocytes reduces CD74 expression and induces both specific and bystander T-cell tolerance. J Autoimmun 2013; 40: 96-110
  • 107 Benedek G, Meza-Romero R, Andrew S. , et al. Partial MHC class II constructs inhibit MIF/CD74 binding and downstream effects. Eur J Immunol 2013; 43 (05) 1309-1321
  • 108 Berkova Z, Tao RH, Samaniego F. Milatuzumab - a promising new immunotherapeutic agent. Expert Opin Investig Drugs 2010; 19 (01) 141-149
  • 109 Borghese F, Clanchy FI. CD74: an emerging opportunity as a therapeutic target in cancer and autoimmune disease. Expert Opin Ther Targets 2011; 15 (03) 237-251
  • 110 Sun J, Hartvigsen K, Chou MY. , et al. Deficiency of antigen-presenting cell invariant chain reduces atherosclerosis in mice. Circulation 2010; 122 (08) 808-820
  • 111 Dolgin E. First GPCR-directed antibody passes approval milestone. Nat Rev Drug Discov 2018; 17 (07) 457-459
  • 112 Peng L, Damschroder MM, Cook KE, Wu H, Dall'Acqua WF. Molecular basis for the antagonistic activity of an anti-CXCR4 antibody. MAbs 2016; 8 (01) 163-175
  • 113 Jantunen E. Novel strategies for blood stem cell mobilization: special focus on plerixafor. Expert Opin Biol Ther 2011; 11 (09) 1241-1248
  • 114 Zernecke A, Bot I, Djalali-Talab Y. , et al. Protective role of CXC receptor 4/CXC ligand 12 unveils the importance of neutrophils in atherosclerosis. Circ Res 2008; 102 (02) 209-217
  • 115 Döring Y, Noels H, van der Vorst EPC. , et al. Vascular CXCR4 limits atherosclerosis by maintaining arterial integrity: evidence from mouse and human studies. Circulation 2017; 136 (04) 388-403
  • 116 Boisvert WA, Santiago R, Curtiss LK, Terkeltaub RA. A leukocyte homologue of the IL-8 receptor CXCR-2 mediates the accumulation of macrophages in atherosclerotic lesions of LDL receptor-deficient mice. J Clin Invest 1998; 101 (02) 353-363
  • 117 Boisvert WA, Curtiss LK, Terkeltaub RA. Interleukin-8 and its receptor CXCR2 in atherosclerosis. Immunol Res 2000; 21 (2-3): 129-137
  • 118 Highfill SL, Cui Y, Giles AJ. , et al. Disruption of CXCR2-mediated MDSC tumor trafficking enhances anti-PD1 efficacy. Sci Transl Med 2014; 6 (237) 237ra67
  • 119 Stadtmann A, Zarbock A. CXCR2: from bench to bedside. Front Immunol 2012; 3: 263
  • 120 Joseph JP, Reyes E, Guzman J. , et al. CXCR2 inhibition - a novel approach to treating CoronAry heart DiseAse (CICADA): study protocol for a randomised controlled trial. Trials 2017; 18 (01) 473
  • 121 Kok T, Wasiel AA, Cool RH. , et al. Small-molecule inhibitors of MIF as an emerging class of therapeutics for immune disorders. Drug Discov Today 2018; 23 (11) 1910-1918
  • 122 Trivedi-Parmar V, Jorgensen WL. Advances and insights for small molecule inhibition of macrophage migration inhibitory factor. J Med Chem 2018; 61 (18) 8104-8119
  • 123 Ouertatani-Sakouhi H, El-Turk F, Fauvet B. , et al. Identification and characterization of novel classes of macrophage migration inhibitory factor (MIF) inhibitors with distinct mechanisms of action. J Biol Chem 2010; 285 (34) 26581-26598
  • 124 Bai F, Asojo OA, Cirillo P. , et al. A novel allosteric inhibitor of macrophage migration inhibitory factor (MIF). J Biol Chem 2012; 287 (36) 30653-30663
  • 125 Pantouris G, Bucala R, Lolis EJ. Structural plasticity in the C-terminal region of macrophage migration inhibitory factor-2 is associated with an induced fit mechanism for a selective inhibitor. Biochemistry 2018; 57 (26) 3599-3605
  • 126 Fingerle-Rowson G, Kaleswarapu DR, Schlander C. , et al. A tautomerase-null macrophage migration-inhibitory factor (MIF) gene knock-in mouse model reveals that protein interactions and not enzymatic activity mediate MIF-dependent growth regulation. Mol Cell Biol 2009; 29 (07) 1922-1932
  • 127 Pantouris G, Syed MA, Fan C. , et al. An analysis of MIF structural features that control functional activation of CD74. Chem Biol 2015; 22 (09) 1197-1205
  • 128 Cournia Z, Leng L, Gandavadi S. , et al. Discovery of human MIF-CD74 antagonists via virtual screening. J Med Chem 2009; 52 (02) 416-424
  • 129 Jorgensen WL, Gandavadi S, Du X. , et al. Receptor agonists of macrophage migration inhibitory factor. Bioorg Med Chem Lett 2010; 20 (23) 7033-7036
  • 130 Lubetsky JB, Dios A, Han J. , et al. The tautomerase active site of macrophage migration inhibitory factor is a potential target for discovery of novel anti-inflammatory agents. J Biol Chem 2002; 277 (28) 24976-24982
  • 131 Senter PD, Al-Abed Y, Metz CN. , et al. Inhibition of macrophage migration inhibitory factor (MIF) tautomerase and biological activities by acetaminophen metabolites. Proc Natl Acad Sci U S A 2002; 99 (01) 144-149
  • 132 Fox RJ, Coffey CS, Conwit R. , et al; NN102/SPRINT-MS Trial Investigators. Phase 2 trial of Ibudilast in progressive multiple sclerosis. N Engl J Med 2018; 379 (09) 846-855
  • 133 Brown KK, Blaikie FH, Smith RA. , et al. Direct modification of the proinflammatory cytokine macrophage migration inhibitory factor by dietary isothiocyanates. J Biol Chem 2009; 284 (47) 32425-32433
  • 134 Priego N, Zhu L, Monteiro C. , et al. STAT3 labels a subpopulation of reactive astrocytes required for brain metastasis. Nat Med 2018; 24 (07) 1024-1035
  • 135 Heinrichs D, Berres M-L, Coeuru M. , et al. Protective role of MIF in nonalcoholic steatohepatitis. FASEB J 2014; 28 (12) 5136-5147
  • 136 Fosgerau K, Hoffmann T. Peptide therapeutics: current status and future directions. Drug Discov Today 2015; 20 (01) 122-128
  • 137 Lau JL, Dunn MK. Therapeutic peptides: historical perspectives, current development trends, and future directions. Bioorg Med Chem 2018; 26 (10) 2700-2707
  • 138 Henninot A, Collins JC, Nuss JM. The current state of peptide drug discovery: back to the future?. J Med Chem 2018; 61 (04) 1382-1414
  • 139 Zorzi A, Deyle K, Heinis C. Cyclic peptide therapeutics: past, present and future. Curr Opin Chem Biol 2017; 38: 24-29
  • 140 Leman LJ, Maryanoff BE, Ghadiri MR. Molecules that mimic apolipoprotein A-I: potential agents for treating atherosclerosis. J Med Chem 2014; 57 (06) 2169-2196
  • 141 Mallavia B, Recio C, Oguiza A. , et al. Peptide inhibitor of NF-κB translocation ameliorates experimental atherosclerosis. Am J Pathol 2013; 182 (05) 1910-1921
  • 142 Hong HY, Lee HY, Kwak W. , et al. Phage display selection of peptides that home to atherosclerotic plaques: IL-4 receptor as a candidate target in atherosclerosis. J Cell Mol Med 2008; 12 (5B): 2003-2014
  • 143 Koenen RR, von Hundelshausen P, Nesmelova IV. , et al. Disrupting functional interactions between platelet chemokines inhibits atherosclerosis in hyperlipidemic mice. Nat Med 2009; 15 (01) 97-103
  • 144 Koenen RR, Weber C. Therapeutic targeting of chemokine interactions in atherosclerosis. Nat Rev Drug Discov 2010; 9 (02) 141-153
  • 145 Figueiredo CR, Matsuo AL, Massaoka MH, Polonelli L, Travassos LR. Anti-tumor activities of peptides corresponding to conserved complementary determining regions from different immunoglobulins. Peptides 2014; 59: 14-19
  • 146 Figueiredo CR, Azevedo RA, Mousdell S. , et al. Blockade of MIF-CD74 signalling on macrophages and dendritic cells restores the antitumour immune response against metastatic melanoma. Front Immunol 2018; 9: 1132
  • 147 Yang L, Liu Z, Ren H. , et al. DRα1-MOG-35-55 treatment reduces lesion volumes and improves neurological deficits after traumatic brain injury. Metab Brain Dis 2017; 32 (05) 1395-1402
  • 148 Meza-Romero R, Benedek G, Leng L, Bucala R, Vandenbark AA. Predicted structure of MIF/CD74 and RTL1000/CD74 complexes. Metab Brain Dis 2016; 31 (02) 249-255
  • 149 Offner H, Sinha S, Burrows GG, Ferro AJ, Vandenbark AA. RTL therapy for multiple sclerosis: a Phase I clinical study. J Neuroimmunol 2011; 231 (1-2): 7-14
  • 150 Meza-Romero R, Benedek G, Jordan K. , et al. Modeling of both shared and distinct interactions between MIF and its homologue D-DT with their common receptor CD74. Cytokine 2016; 88: 62-70
  • 151 Crump MP, Gong JH, Loetscher P. , et al. Solution structure and basis for functional activity of stromal cell-derived factor-1; dissociation of CXCR4 activation from binding and inhibition of HIV-1. EMBO J 1997; 16 (23) 6996-7007
  • 152 Tamamura H, Hori A, Kanzaki N. , et al. T140 analogs as CXCR4 antagonists identified as anti-metastatic agents in the treatment of breast cancer. FEBS Lett 2003; 550 (1-3): 79-83
  • 153 DeMarco SJ, Henze H, Lederer A. , et al. Discovery of novel, highly potent and selective beta-hairpin mimetic CXCR4 inhibitors with excellent anti-HIV activity and pharmacokinetic profiles. Bioorg Med Chem 2006; 14 (24) 8396-8404
  • 154 Fujii N, Oishi S, Hiramatsu K. , et al. Molecular-size reduction of a potent CXCR4-chemokine antagonist using orthogonal combination of conformation- and sequence-based libraries. Angew Chem Int Ed Engl 2003; 42 (28) 3251-3253
  • 155 Ueda S, Oishi S, Wang ZX. , et al. Structure-activity relationships of cyclic peptide-based chemokine receptor CXCR4 antagonists: disclosing the importance of side-chain and backbone functionalities. J Med Chem 2007; 50 (02) 192-198
  • 156 Demmer O, Frank AO, Hagn F. , et al. A conformationally frozen peptoid boosts CXCR4 affinity and anti-HIV activity. Angew Chem Int Ed Engl 2012; 51 (32) 8110-8113
  • 157 Demmer O, Dijkgraaf I, Schumacher U. , et al. Design, synthesis, and functionalization of dimeric peptides targeting chemokine receptor CXCR4. J Med Chem 2011; 54 (21) 7648-7662
  • 158 Di Maro S, Trotta AM, Brancaccio D. , et al. Exploring the N-terminal region of C–X-C motif chemokine 12 (CXCL12): identification of plasma-stable cyclic peptides as novel, potent C–X-C chemokine receptor type 4 (CXCR4) antagonists. J Med Chem 2016; 59 (18) 8369-8380
  • 159 Di Maro S, Di Leva FS, Trotta AM. , et al. Structure-activity relationships and biological characterization of a novel, potent, and serum stable C–X-C chemokine receptor type 4 (CXCR4) antagonist. J Med Chem 2017; 60 (23) 9641-9652
  • 160 Groß A, Möbius K, Haußner C, Donhauser N, Schmidt B, Eichler J. Mimicking protein-protein interactions through peptide-peptide interactions: HIV-1 gp120 and CXCR4. Front Immunol 2013; 4: 257
  • 161 Möbius K, Dürr R, Haussner C, Dietrich U, Eichler J. A functionally selective synthetic mimic of the HIV-1 co-receptor CXCR4. Chemistry 2012; 18 (27) 8292-8295
  • 162 Wang J, Tong C, Yan X. , et al. Limiting cardiac ischemic injury by pharmacological augmentation of macrophage migration inhibitory factor-AMP-activated protein kinase signal transduction. Circulation 2013; 128 (03) 225-236
  • 163 Höllriegl W, Bauer A, Baumgartner B. , et al. Pharmacokinetics, disease-modifying activity, and safety of an experimental therapeutic targeting an immunological isoform of macrophage migration inhibitory factor, in rat glomerulonephritis. Eur J Pharmacol 2018; 820: 206-216
  • 164 Sparkes A, De Baetselier P, Brys L. , et al. Novel half-life extended anti-MIF nanobodies protect against endotoxic shock. FASEB J 2018; 32 (06) 3411-3422
  • 165 Kaufman JL, Niesvizky R, Stadtmauer EA. , et al. Phase I, multicentre, dose-escalation trial of monotherapy with milatuzumab (humanized anti-CD74 monoclonal antibody) in relapsed or refractory multiple myeloma. Br J Haematol 2013; 163 (04) 478-486
  • 166 Griffiths K, Dolezal O, Cao B. , et al. i-bodies, human single domain antibodies that antagonize chemokine receptor CXCR4. J Biol Chem 2016; 291 (24) 12641-12657
  • 167 Korkolopoulou P, Levidou G, El-Habr EA. , et al. Expression of interleukin-8 receptor CXCR2 and suppressor of cytokine signaling-3 in astrocytic tumors. Mol Med 2012; 18: 379-388
  • 168 Crichlow GV, Lubetsky JB, Leng L, Bucala R, Lolis EJ. Structural and kinetic analyses of macrophage migration inhibitory factor active site interactions. Biochemistry 2009; 48 (01) 132-139
  • 169 Rajasekaran D, Zierow S, Syed M, Bucala R, Bhandari V, Lolis EJ. Targeting distinct tautomerase sites of D-DT and MIF with a single molecule for inhibition of neutrophil lung recruitment. FASEB J 2014; 28 (11) 4961-4971
  • 170 Winner M, Meier J, Zierow S. , et al. A novel, MIF suicide substrate inhibits motility and growth of lung cancer cells. Cancer Res 2008; 68 (18) 7253-7257
  • 171 Al-Abed Y, Dabideen D, Aljabari B. , et al. ISO-1 binding to the tautomerase active site of MIF inhibits its pro-inflammatory activity and increases survival in severe sepsis. J Biol Chem 2005; 280 (44) 36541-36544
  • 172 Tynan A, Mawhinney L, Armstrong ME. , et al. Macrophage migration inhibitory factor enhances Pseudomonas aeruginosa biofilm formation, potentially contributing to cystic fibrosis pathogenesis. FASEB J 2017; 31 (11) 5102-5110
  • 173 Mawhinney L, Armstrong ME, O' Reilly C. , et al. Macrophage migration inhibitory factor (MIF) enzymatic activity and lung cancer. Mol Med 2015; 20: 729-735
  • 174 Pease J, Horuk R. Chemokine receptor antagonists. J Med Chem 2012; 55 (22) 9363-9392
  • 175 Bizzarri C, Beccari AR, Bertini R, Cavicchia MR, Giorgini S, Allegretti M. ELR+ CXC chemokines and their receptors (CXC chemokine receptor 1 and CXC chemokine receptor 2) as new therapeutic targets. Pharmacol Ther 2006; 112 (01) 139-149
  • 176 Kleemann R, Hausser A, Geiger G. , et al. Intracellular action of the cytokine MIF to modulate AP-1 activity and the cell cycle through Jab1. Nature 2000; 408 (6809): 211-216
  • 177 Nguyen MT, Beck J, Lue H. , et al. A 16-residue peptide fragment of macrophage migration inhibitory factor, MIF-(50-65), exhibits redox activity and has MIF-like biological functions. J Biol Chem 2003; 278 (36) 33654-33671
  • 178 Liang X. CXCR4, inhibitors and mechanisms of action. Chem Biol Drug Des 2008; 72 (02) 97-110
  • 179 Aboye TL, Ha H, Majumder S. , et al. Design of a novel cyclotide-based CXCR4 antagonist with anti-human immunodeficiency virus (HIV)-1 activity. J Med Chem 2012; 55 (23) 10729-10734

Address for correspondence

Jürgen Bernhagen, PhD
Chair of Vascular Biology, Institute for Stroke and Dementia Research (ISD)
Klinikum der Universität München (KUM), Ludwig-Maximilians-University (LMU) Munich, Feodor-Lynen-Straße 17, 81377 Munich
Germany   
Aphrodite Kapurniotu, PhD
Division of Peptide Biochemistry, Technische Universität München (TUM)
Emil-Erlenmeyer-Forum 5, 85354 Freising
Germany   

  • References

  • 1 Ross R. Atherosclerosis--an inflammatory disease. N Engl J Med 1999; 340 (02) 115-126
  • 2 Libby P, Ridker PM, Maseri A. Inflammation and atherosclerosis. Circulation 2002; 105 (09) 1135-1143
  • 3 Dahlöf B. Cardiovascular disease risk factors: epidemiology and risk assessment. Am J Cardiol 2010; 105 (1, Suppl): 3A-9A
  • 4 van der Vorst EP, Döring Y, Weber C. Chemokines and their receptors in atherosclerosis. J Mol Med (Berl) 2015; 93 (09) 963-971
  • 5 Charo IF, Ransohoff RM. The many roles of chemokines and chemokine receptors in inflammation. N Engl J Med 2006; 354 (06) 610-621
  • 6 Bachelerie F, Ben-Baruch A, Burkhardt AM. , et al. International Union of Basic and Clinical Pharmacology. [corrected]. LXXXIX. Update on the extended family of chemokine receptors and introducing a new nomenclature for atypical chemokine receptors. Pharmacol Rev 2013; 66 (01) 1-79
  • 7 Weber C, Noels H. Atherosclerosis: current pathogenesis and therapeutic options. Nat Med 2011; 17 (11) 1410-1422
  • 8 Bäck M, Hansson GK. Anti-inflammatory therapies for atherosclerosis. Nat Rev Cardiol 2015; 12 (04) 199-211
  • 9 Ridker PM, Lüscher TF. Anti-inflammatory therapies for cardiovascular disease. Eur Heart J 2014; 35 (27) 1782-1791
  • 10 Ramji DP, Davies TS. Cytokines in atherosclerosis: key players in all stages of disease and promising therapeutic targets. Cytokine Growth Factor Rev 2015; 26 (06) 673-685
  • 11 Ruparelia N, Chai JT, Fisher EA, Choudhury RP. Inflammatory processes in cardiovascular disease: a route to targeted therapies. Nat Rev Cardiol 2017; 14 (03) 133-144
  • 12 Ridker PM, Everett BM, Thuren T. , et al; CANTOS Trial Group. Antiinflammatory therapy with canakinumab for atherosclerotic disease. N Engl J Med 2017; 377 (12) 1119-1131
  • 13 Bernhagen J, Calandra T, Mitchell RA. , et al. MIF is a pituitary-derived cytokine that potentiates lethal endotoxaemia. Nature 1993; 365 (6448): 756-759
  • 14 Bernhagen J, Krohn R, Lue H. , et al. MIF is a noncognate ligand of CXC chemokine receptors in inflammatory and atherogenic cell recruitment. Nat Med 2007; 13 (05) 587-596
  • 15 Calandra T, Roger T. Macrophage migration inhibitory factor: a regulator of innate immunity. Nat Rev Immunol 2003; 3 (10) 791-800
  • 16 Tillmann S, Bernhagen J, Noels H. Arrest functions of the MIF ligand/receptor axes in atherogenesis. Front Immunol 2013; 4: 115
  • 17 Burger-Kentischer A, Goebel H, Seiler R. , et al. Expression of MIF in different stages of human atherosclerosis. Circulation 2002; 105 (13) 1561-1566
  • 18 Zernecke A, Bernhagen J, Weber C. Macrophage migration inhibitory factor in cardiovascular disease. Circulation 2008; 117 (12) 1594-1602
  • 19 Tilstam PV, Qi D, Leng L, Young L, Bucala R. MIF family cytokines in cardiovascular diseases and prospects for precision-based therapeutics. Expert Opin Ther Targets 2017; 21 (07) 671-683
  • 20 David JR. Delayed hypersensitivity in vitro: its mediation by cell-free substances formed by lymphoid cell-antigen interaction. Proc Natl Acad Sci U S A 1966; 56 (01) 72-77
  • 21 Rich AR, Lewis MR. Migration of neutrophils and macrophages. Bull Johns Hopkins Hosp 1932; 50: 115-131
  • 22 Mischke R, Kleemann R, Brunner H, Bernhagen J. Cross-linking and mutational analysis of the oligomerization state of the cytokine macrophage migration inhibitory factor (MIF). FEBS Lett 1998; 427 (01) 85-90
  • 23 Sun HW, Bernhagen J, Bucala R, Lolis E. Crystal structure at 2.6-A resolution of human macrophage migration inhibitory factor. Proc Natl Acad Sci U S A 1996; 93 (11) 5191-5196
  • 24 Lolis E, Bucala R. Macrophage migration inhibitory factor. Expert Opin Ther Targets 2003; 7 (02) 153-164
  • 25 Flieger O, Engling A, Bucala R, Lue H, Nickel W, Bernhagen J. Regulated secretion of macrophage migration inhibitory factor is mediated by a non-classical pathway involving an ABC transporter. FEBS Lett 2003; 551 (1-3): 78-86
  • 26 Merk M, Baugh J, Zierow S. , et al. The Golgi-associated protein p115 mediates the secretion of macrophage migration inhibitory factor. J Immunol 2009; 182 (11) 6896-6906
  • 27 Hertelendy J, Reumuth G, Simons D. , et al. Macrophage migration inhibitory factor - a favorable marker in inflammatory diseases?. Curr Med Chem 2018; 25 (05) 601-605
  • 28 Morand EF, Leech M, Bernhagen J. MIF: a new cytokine link between rheumatoid arthritis and atherosclerosis. Nat Rev Drug Discov 2006; 5 (05) 399-410
  • 29 Morrison MC, Kleemann R. Role of macrophage migration inhibitory factor in obesity, insulin resistance, type 2 diabetes, and associated hepatic co-morbidities: a comprehensive review of human and rodent studies. Front Immunol 2015; 6: 308
  • 30 Sauler M, Bucala R, Lee PJ. Role of macrophage migration inhibitory factor in age-related lung disease. Am J Physiol Lung Cell Mol Physiol 2015; 309 (01) L1-L10
  • 31 O'Reilly C, Doroudian M, Mawhinney L, Donnelly SC. Targeting MIF in cancer: therapeutic strategies, current developments, and future opportunities. Med Res Rev 2016; 36 (03) 440-460
  • 32 Conroy H, Mawhinney L, Donnelly SC. Inflammation and cancer: macrophage migration inhibitory factor (MIF)--the potential missing link. QJM 2010; 103 (11) 831-836
  • 33 Chesney JA, Mitchell RA, Yaddanapudi K. Myeloid-derived suppressor cells-a new therapeutic target to overcome resistance to cancer immunotherapy. J Leukoc Biol 2017; 102 (03) 727-740
  • 34 Chesney JA, Mitchell RA. 25 years on: a retrospective on migration inhibitory factor in tumor angiogenesis. Mol Med 2015; 21 (Suppl. 01) S19-S24
  • 35 Wirtz TH, Tillmann S, Strüßmann T. , et al. Platelet-derived MIF: a novel platelet chemokine with distinct recruitment properties. Atherosclerosis 2015; 239 (01) 1-10
  • 36 Lin SG, Yu XY, Chen YX. , et al. De novo expression of macrophage migration inhibitory factor in atherogenesis in rabbits. Circ Res 2000; 87 (12) 1202-1208
  • 37 Burger-Kentischer A, Göbel H, Kleemann R. , et al. Reduction of the aortic inflammatory response in spontaneous atherosclerosis by blockade of macrophage migration inhibitory factor (MIF). Atherosclerosis 2006; 184 (01) 28-38
  • 38 Pan JH, Sukhova GK, Yang JT. , et al. Macrophage migration inhibitory factor deficiency impairs atherosclerosis in low-density lipoprotein receptor-deficient mice. Circulation 2004; 109 (25) 3149-3153
  • 39 Amin MA, Haas CS, Zhu K. , et al. Migration inhibitory factor up-regulates vascular cell adhesion molecule-1 and intercellular adhesion molecule-1 via Src, PI3 kinase, and NFkappaB. Blood 2006; 107 (06) 2252-2261
  • 40 Gregory JL, Morand EF, McKeown SJ. , et al. Macrophage migration inhibitory factor induces macrophage recruitment via CC chemokine ligand 2. J Immunol 2006; 177 (11) 8072-8079
  • 41 Atsumi T, Nishihira J, Makita Z, Koike T. Enhancement of oxidised low-density lipoprotein uptake by macrophages in response to macrophage migration inhibitory factor. Cytokine 2000; 12 (10) 1553-1556
  • 42 Kong YZ, Huang XR, Ouyang X. , et al. Evidence for vascular macrophage migration inhibitory factor in destabilization of human atherosclerotic plaques. Cardiovasc Res 2005; 65 (01) 272-282
  • 43 Calandra T, Bernhagen J, Metz CN. , et al. MIF as a glucocorticoid-induced modulator of cytokine production. Nature 1995; 377 (6544): 68-71
  • 44 Luo JY, Xu R, Li XM. , et al. MIF gene polymorphism rs755622 is associated with coronary artery disease and severity of coronary lesions in a Chinese Kazakh population: a case-control study. Medicine (Baltimore) 2016; 95 (04) e2617
  • 45 Herder C, Illig T, Baumert J. , et al. Macrophage migration inhibitory factor (MIF) and risk for coronary heart disease: results from the MONICA/KORA Augsburg case-cohort study, 1984-2002. Atherosclerosis 2008; 200 (02) 380-388
  • 46 Ji K, Wang X, Li J. , et al. Macrophage migration inhibitory factor polymorphism is associated with susceptibility to inflammatory coronary heart disease. BioMed Res Int 2015; 2015: 315174
  • 47 Baugh JA, Chitnis S, Donnelly SC. , et al. A functional promoter polymorphism in the macrophage migration inhibitory factor (MIF) gene associated with disease severity in rheumatoid arthritis. Genes Immun 2002; 3 (03) 170-176
  • 48 Lan MY, Chang YY, Chen WH. , et al. Association between MIF gene polymorphisms and carotid artery atherosclerosis. Biochem Biophys Res Commun 2013; 435 (02) 319-322
  • 49 Müller II, Müller KA, Karathanos A. , et al. Impact of counterbalance between macrophage migration inhibitory factor and its inhibitor Gremlin-1 in patients with coronary artery disease. Atherosclerosis 2014; 237 (02) 426-432
  • 50 Müller II, Müller KAL, Schönleber H. , et al. Macrophage migration inhibitory factor is enhanced in acute coronary syndromes and is associated with the inflammatory response. PLoS One 2012; 7 (06) e38376
  • 51 Deng XN, Wang XY, Yu HY. , et al. Admission macrophage migration inhibitory factor predicts long-term prognosis in patients with ST-elevation myocardial infarction. Eur Heart J Qual Care Clin Outcomes 2018; 4 (03) 208-219
  • 52 Deng F, Zhao Q, Deng Y. , et al. Prognostic significance and dynamic change of plasma macrophage migration inhibitory factor in patients with acute ST-elevation myocardial infarction. Medicine (Baltimore) 2018; 97 (43) e12991
  • 53 Chan W, White DA, Wang XY. , et al. Macrophage migration inhibitory factor for the early prediction of infarct size. J Am Heart Assoc 2013; 2 (05) e000226
  • 54 Stoppe C, Rex S, Goetzenich A. , et al. Interaction of MIF family proteins in myocardial ischemia/reperfusion damage and their influence on clinical outcome of cardiac surgery patients. Antioxid Redox Signal 2015; 23 (11) 865-879
  • 55 Stoppe C, Werker T, Rossaint R. , et al. What is the significance of perioperative release of macrophage migration inhibitory factor in cardiac surgery?. Antioxid Redox Signal 2013; 19 (03) 231-239
  • 56 Stoppe C, Averdunk L, Goetzenich A. , et al. The protective role of macrophage migration inhibitory factor in acute kidney injury after cardiac surgery. Sci Transl Med 2018; 10 (441) eaan4886
  • 57 Luedike P, Hendgen-Cotta UB, Sobierajski J. , et al. Cardioprotection through S-nitros(yl)ation of macrophage migration inhibitory factor. Circulation 2012; 125 (15) 1880-1889
  • 58 Miller EJ, Li J, Leng L. , et al. Macrophage migration inhibitory factor stimulates AMP-activated protein kinase in the ischaemic heart. Nature 2008; 451 (7178): 578-582
  • 59 Pohl J, Hendgen-Cotta UB, Rammos C. , et al. Targeted intracellular accumulation of macrophage migration inhibitory factor in the reperfused heart mediates cardioprotection. Thromb Haemost 2016; 115 (01) 200-212
  • 60 Qi D, Hu X, Wu X. , et al. Cardiac macrophage migration inhibitory factor inhibits JNK pathway activation and injury during ischemia/reperfusion. J Clin Invest 2009; 119 (12) 3807-3816
  • 61 Rassaf T, Weber C, Bernhagen J. Macrophage migration inhibitory factor in myocardial ischaemia/reperfusion injury. Cardiovasc Res 2014; 102 (02) 321-328
  • 62 Koga K, Kenessey A, Powell SR, Sison CP, Miller EJ, Ojamaa K. Macrophage migration inhibitory factor provides cardioprotection during ischemia/reperfusion by reducing oxidative stress. Antioxid Redox Signal 2011; 14 (07) 1191-1202
  • 63 Qi D, Atsina K, Qu L. , et al. The vestigial enzyme D-dopachrome tautomerase protects the heart against ischemic injury. J Clin Invest 2014; 124 (08) 3540-3550
  • 64 Dayawansa NH, Gao X-M, White DA, Dart AM, Du XJ. Role of MIF in myocardial ischaemia and infarction: insight from recent clinical and experimental findings. Clin Sci (Lond) 2014; 127 (03) 149-161
  • 65 Leng L, Metz CN, Fang Y. , et al. MIF signal transduction initiated by binding to CD74. J Exp Med 2003; 197 (11) 1467-1476
  • 66 Ma H, Wang J, Thomas DP. , et al. Impaired macrophage migration inhibitory factor-AMP-activated protein kinase activation and ischemic recovery in the senescent heart. Circulation 2010; 122 (03) 282-292
  • 67 Schindler L, Dickerhof N, Hampton MB. , et al. Post-translational regulation of MIF: basis for functional fine-tuning. Redox Biol 2018; 15: 135-142
  • 68 Liehn EA, Kanzler I, Konschalla S. , et al. Compartmentalized protective and detrimental effects of endogenous MIF mediated by CXCR2 in a mouse model of myocardial ischemia/reperfusion. Arterioscler Thromb Vasc Biol 2013; 33 (09) 2180-2186
  • 69 Liehn EA, Tuchscheerer N, Kanzler I. , et al. Double-edged role of the CXCL12/CXCR4 axis in experimental myocardial infarction. J Am Coll Cardiol 2011; 58 (23) 2415-2423
  • 70 Alampour-Rajabi S, El Bounkari O, Rot A. , et al. MIF interacts with CXCR7 to promote receptor internalization, ERK1/2 and ZAP-70 signaling, and lymphocyte chemotaxis. FASEB J 2015; 29 (11) 4497-4511
  • 71 Chatterjee M, Borst O, Walker B. , et al. Macrophage migration inhibitory factor limits activation-induced apoptosis of platelets via CXCR7-dependent Akt signaling. Circ Res 2014; 115 (11) 939-949
  • 72 Tarnowski M, Grymula K, Liu R. , et al. Macrophage migration inhibitory factor is secreted by rhabdomyosarcoma cells, modulates tumor metastasis by binding to CXCR4 and CXCR7 receptors and inhibits recruitment of cancer-associated fibroblasts. Mol Cancer Res 2010; 8 (10) 1328-1343
  • 73 Cho Y, Crichlow GV, Vermeire JJ. , et al. Allosteric inhibition of MIF revealed by ibudilast. Proc Natl Acad Sci U S A 2010; 107 (25) 11313-11318
  • 74 Xu L, Li Y, Li D. , et al. Exploring the binding mechanisms of MIF to CXCR2 using theoretical approaches. Phys Chem Chem Phys 2015; 17 (05) 3370-3382
  • 75 Weber C, Kraemer S, Drechsler M. , et al. Structural determinants of MIF functions in CXCR2-mediated inflammatory and atherogenic leukocyte recruitment. Proc Natl Acad Sci U S A 2008; 105 (42) 16278-16283
  • 76 Kraemer S, Lue H, Zernecke A. , et al. MIF-chemokine receptor interactions in atherogenesis are dependent on an N-loop-based 2-site binding mechanism. FASEB J 2011; 25 (03) 894-906
  • 77 Klasen C, Ohl K, Sternkopf M. , et al. MIF promotes B cell chemotaxis through the receptors CXCR4 and CD74 and ZAP-70 signaling. J Immunol 2014; 192 (11) 5273-5284
  • 78 Schmitz C, Noels H, El Bounkari O. , et al. Mif-deficiency favors an atheroprotective autoantibody phenotype in atherosclerosis. FASEB J 2018; 32 (08) 4428-4443
  • 79 Qin L, Kufareva I, Holden LG. , et al. Structural biology. Crystal structure of the chemokine receptor CXCR4 in complex with a viral chemokine. Science 2015; 347 (6226): 1117-1122
  • 80 Wu B, Chien EY, Mol CD. , et al. Structures of the CXCR4 chemokine GPCR with small-molecule and cyclic peptide antagonists. Science 2010; 330 (6007): 1066-1071
  • 81 Lacy M, Kontos C, Brandhofer M. , et al. Identification of an Arg-Leu-Arg tripeptide that contributes to the binding interface between the cytokine MIF and the chemokine receptor CXCR4. Sci Rep 2018; 8 (01) 5171
  • 82 Rajasekaran D, Gröning S, Schmitz C. , et al. Macrophage migration inhibitory factor-CXCR4 receptor interactions: evidence for partial allosteric agonism in comparison with CXCL12 chemokine. J Biol Chem 2016; 291 (30) 15881-15895
  • 83 Schwartz V, Lue H, Kraemer S. , et al. A functional heteromeric MIF receptor formed by CD74 and CXCR4. FEBS Lett 2009; 583 (17) 2749-2757
  • 84 Klasen C, Ziehm T, Huber M. , et al. LPS-mediated cell surface expression of CD74 promotes the proliferation of B cells in response to MIF. Cell Signal 2018; 46: 32-42
  • 85 Bertolino P, Rabourdin-Combe C. The MHC class II-associated invariant chain: a molecule with multiple roles in MHC class II biosynthesis and antigen presentation to CD4+ T cells. Crit Rev Immunol 1996; 16 (04) 359-379
  • 86 Merk M, Zierow S, Leng L. , et al. The D-dopachrome tautomerase (DDT) gene product is a cytokine and functional homolog of macrophage migration inhibitory factor (MIF). Proc Natl Acad Sci U S A 2011; 108 (34) E577 –E585
  • 87 Shi X, Leng L, Wang T. , et al. CD44 is the signaling component of the macrophage migration inhibitory factor-CD74 receptor complex. Immunity 2006; 25 (04) 595-606
  • 88 Assis DN, Leng L, Du X. , et al. The role of macrophage migration inhibitory factor in autoimmune liver disease. Hepatology 2014; 59 (02) 580-591
  • 89 Monaco C, Nanchahal J, Taylor P, Feldmann M. Anti-TNF therapy: past, present and future. Int Immunol 2015; 27 (01) 55-62
  • 90 Loberg RD, Ying C, Craig M, Yan L, Snyder LA, Pienta KJ. CCL2 as an important mediator of prostate cancer growth in vivo through the regulation of macrophage infiltration. Neoplasia 2007; 9 (07) 556-562
  • 91 Loberg RD, Ying C, Craig M. , et al. Targeting CCL2 with systemic delivery of neutralizing antibodies induces prostate cancer tumor regression in vivo. Cancer Res 2007; 67 (19) 9417-9424
  • 92 Calandra T, Echtenacher B, Roy DL. , et al. Protection from septic shock by neutralization of macrophage migration inhibitory factor. Nat Med 2000; 6 (02) 164-170
  • 93 Kerschbaumer RJ, Rieger M, Völkel D. , et al. Neutralization of macrophage migration inhibitory factor (MIF) by fully human antibodies correlates with their specificity for the β-sheet structure of MIF. J Biol Chem 2012; 287 (10) 7446-7455
  • 94 Lan HY, Bacher M, Yang N. , et al. The pathogenic role of macrophage migration inhibitory factor in immunologically induced kidney disease in the rat. J Exp Med 1997; 185 (08) 1455-1465
  • 95 Bloom J, Sun S, Al-Abed Y. MIF, a controversial cytokine: a review of structural features, challenges, and opportunities for drug development. Expert Opin Ther Targets 2016; 20 (12) 1463-1475
  • 96 Zhang Y, Zeng X, Chen S. , et al. Characterization, epitope identification and mechanisms of the anti-septic capacity of monoclonal antibodies against macrophage migration inhibitory factor. Int Immunopharmacol 2011; 11 (09) 1333-1340
  • 97 Mahalingam D, Patel M, Sachdev J. , et al. Safety and efficacy analysis of imalumab, an anti-oxidized macrophage migration inhibitory factor (oxMIF) antibody, alone or in combination with 5-fluorouracil/leucovorin (5-FU/LV) or panitumumab, in patients with metastatic colorectal cancer (mCRC). Ann Oncol 2016; 2016: ii105
  • 98 Thiele M, Kerschbaumer RJ, Tam FW. , et al. Selective targeting of a disease-related conformational isoform of MIF ameliorates inflammatory conditions. J Immunol 2015; 195 (05) 2343-2352
  • 99 Schinagl A, Kerschbaumer RJ, Sabarth N. , et al. Role of the cysteine 81 residue of MIF as a molecular redox switch. Biochemistry 2018; 57 (09) 1523-1532
  • 100 Schinagl A, Thiele M, Douillard P. , et al. Oxidized MIF is a potential new tissue marker and drug target in cancer. Oncotarget 2016; 7 (45) 73486-73496
  • 101 Ochi A, Chen D, Schulte W. , et al. MIF-2/D-DT enhances proximal tubular cell regeneration through SLPI- and ATF4-dependent mechanisms. Am J Physiol Renal Physiol 2017; 313 (03) F767-F780
  • 102 Pohl J, Hendgen-Cotta UB, Stock P, Luedike P, Rassaf T. Elevated MIF-2 levels predict mortality in critically ill patients. J Crit Care 2017; 40: 52-57
  • 103 Merk M, Mitchell RA, Endres S, Bucala R. D-dopachrome tautomerase (D-DT or MIF-2): doubling the MIF cytokine family. Cytokine 2012; 59 (01) 10-17
  • 104 Benedek G, Meza-Romero R, Jordan K, Keenlyside L, Offner H, Vandenbark AA. HLA-DRα1-mMOG-35-55 treatment of experimental autoimmune encephalomyelitis reduces CNS inflammation, enhances M2 macrophage frequency, and promotes neuroprotection. J Neuroinflammation 2015; 12: 123
  • 105 Meza-Romero R, Benedek G, Yu X. , et al. HLA-DRα1 constructs block CD74 expression and MIF effects in experimental autoimmune encephalomyelitis. J Immunol 2014; 192 (09) 4164-4173
  • 106 Vandenbark AA, Meza-Romero R, Benedek G. , et al. A novel regulatory pathway for autoimmune disease: binding of partial MHC class II constructs to monocytes reduces CD74 expression and induces both specific and bystander T-cell tolerance. J Autoimmun 2013; 40: 96-110
  • 107 Benedek G, Meza-Romero R, Andrew S. , et al. Partial MHC class II constructs inhibit MIF/CD74 binding and downstream effects. Eur J Immunol 2013; 43 (05) 1309-1321
  • 108 Berkova Z, Tao RH, Samaniego F. Milatuzumab - a promising new immunotherapeutic agent. Expert Opin Investig Drugs 2010; 19 (01) 141-149
  • 109 Borghese F, Clanchy FI. CD74: an emerging opportunity as a therapeutic target in cancer and autoimmune disease. Expert Opin Ther Targets 2011; 15 (03) 237-251
  • 110 Sun J, Hartvigsen K, Chou MY. , et al. Deficiency of antigen-presenting cell invariant chain reduces atherosclerosis in mice. Circulation 2010; 122 (08) 808-820
  • 111 Dolgin E. First GPCR-directed antibody passes approval milestone. Nat Rev Drug Discov 2018; 17 (07) 457-459
  • 112 Peng L, Damschroder MM, Cook KE, Wu H, Dall'Acqua WF. Molecular basis for the antagonistic activity of an anti-CXCR4 antibody. MAbs 2016; 8 (01) 163-175
  • 113 Jantunen E. Novel strategies for blood stem cell mobilization: special focus on plerixafor. Expert Opin Biol Ther 2011; 11 (09) 1241-1248
  • 114 Zernecke A, Bot I, Djalali-Talab Y. , et al. Protective role of CXC receptor 4/CXC ligand 12 unveils the importance of neutrophils in atherosclerosis. Circ Res 2008; 102 (02) 209-217
  • 115 Döring Y, Noels H, van der Vorst EPC. , et al. Vascular CXCR4 limits atherosclerosis by maintaining arterial integrity: evidence from mouse and human studies. Circulation 2017; 136 (04) 388-403
  • 116 Boisvert WA, Santiago R, Curtiss LK, Terkeltaub RA. A leukocyte homologue of the IL-8 receptor CXCR-2 mediates the accumulation of macrophages in atherosclerotic lesions of LDL receptor-deficient mice. J Clin Invest 1998; 101 (02) 353-363
  • 117 Boisvert WA, Curtiss LK, Terkeltaub RA. Interleukin-8 and its receptor CXCR2 in atherosclerosis. Immunol Res 2000; 21 (2-3): 129-137
  • 118 Highfill SL, Cui Y, Giles AJ. , et al. Disruption of CXCR2-mediated MDSC tumor trafficking enhances anti-PD1 efficacy. Sci Transl Med 2014; 6 (237) 237ra67
  • 119 Stadtmann A, Zarbock A. CXCR2: from bench to bedside. Front Immunol 2012; 3: 263
  • 120 Joseph JP, Reyes E, Guzman J. , et al. CXCR2 inhibition - a novel approach to treating CoronAry heart DiseAse (CICADA): study protocol for a randomised controlled trial. Trials 2017; 18 (01) 473
  • 121 Kok T, Wasiel AA, Cool RH. , et al. Small-molecule inhibitors of MIF as an emerging class of therapeutics for immune disorders. Drug Discov Today 2018; 23 (11) 1910-1918
  • 122 Trivedi-Parmar V, Jorgensen WL. Advances and insights for small molecule inhibition of macrophage migration inhibitory factor. J Med Chem 2018; 61 (18) 8104-8119
  • 123 Ouertatani-Sakouhi H, El-Turk F, Fauvet B. , et al. Identification and characterization of novel classes of macrophage migration inhibitory factor (MIF) inhibitors with distinct mechanisms of action. J Biol Chem 2010; 285 (34) 26581-26598
  • 124 Bai F, Asojo OA, Cirillo P. , et al. A novel allosteric inhibitor of macrophage migration inhibitory factor (MIF). J Biol Chem 2012; 287 (36) 30653-30663
  • 125 Pantouris G, Bucala R, Lolis EJ. Structural plasticity in the C-terminal region of macrophage migration inhibitory factor-2 is associated with an induced fit mechanism for a selective inhibitor. Biochemistry 2018; 57 (26) 3599-3605
  • 126 Fingerle-Rowson G, Kaleswarapu DR, Schlander C. , et al. A tautomerase-null macrophage migration-inhibitory factor (MIF) gene knock-in mouse model reveals that protein interactions and not enzymatic activity mediate MIF-dependent growth regulation. Mol Cell Biol 2009; 29 (07) 1922-1932
  • 127 Pantouris G, Syed MA, Fan C. , et al. An analysis of MIF structural features that control functional activation of CD74. Chem Biol 2015; 22 (09) 1197-1205
  • 128 Cournia Z, Leng L, Gandavadi S. , et al. Discovery of human MIF-CD74 antagonists via virtual screening. J Med Chem 2009; 52 (02) 416-424
  • 129 Jorgensen WL, Gandavadi S, Du X. , et al. Receptor agonists of macrophage migration inhibitory factor. Bioorg Med Chem Lett 2010; 20 (23) 7033-7036
  • 130 Lubetsky JB, Dios A, Han J. , et al. The tautomerase active site of macrophage migration inhibitory factor is a potential target for discovery of novel anti-inflammatory agents. J Biol Chem 2002; 277 (28) 24976-24982
  • 131 Senter PD, Al-Abed Y, Metz CN. , et al. Inhibition of macrophage migration inhibitory factor (MIF) tautomerase and biological activities by acetaminophen metabolites. Proc Natl Acad Sci U S A 2002; 99 (01) 144-149
  • 132 Fox RJ, Coffey CS, Conwit R. , et al; NN102/SPRINT-MS Trial Investigators. Phase 2 trial of Ibudilast in progressive multiple sclerosis. N Engl J Med 2018; 379 (09) 846-855
  • 133 Brown KK, Blaikie FH, Smith RA. , et al. Direct modification of the proinflammatory cytokine macrophage migration inhibitory factor by dietary isothiocyanates. J Biol Chem 2009; 284 (47) 32425-32433
  • 134 Priego N, Zhu L, Monteiro C. , et al. STAT3 labels a subpopulation of reactive astrocytes required for brain metastasis. Nat Med 2018; 24 (07) 1024-1035
  • 135 Heinrichs D, Berres M-L, Coeuru M. , et al. Protective role of MIF in nonalcoholic steatohepatitis. FASEB J 2014; 28 (12) 5136-5147
  • 136 Fosgerau K, Hoffmann T. Peptide therapeutics: current status and future directions. Drug Discov Today 2015; 20 (01) 122-128
  • 137 Lau JL, Dunn MK. Therapeutic peptides: historical perspectives, current development trends, and future directions. Bioorg Med Chem 2018; 26 (10) 2700-2707
  • 138 Henninot A, Collins JC, Nuss JM. The current state of peptide drug discovery: back to the future?. J Med Chem 2018; 61 (04) 1382-1414
  • 139 Zorzi A, Deyle K, Heinis C. Cyclic peptide therapeutics: past, present and future. Curr Opin Chem Biol 2017; 38: 24-29
  • 140 Leman LJ, Maryanoff BE, Ghadiri MR. Molecules that mimic apolipoprotein A-I: potential agents for treating atherosclerosis. J Med Chem 2014; 57 (06) 2169-2196
  • 141 Mallavia B, Recio C, Oguiza A. , et al. Peptide inhibitor of NF-κB translocation ameliorates experimental atherosclerosis. Am J Pathol 2013; 182 (05) 1910-1921
  • 142 Hong HY, Lee HY, Kwak W. , et al. Phage display selection of peptides that home to atherosclerotic plaques: IL-4 receptor as a candidate target in atherosclerosis. J Cell Mol Med 2008; 12 (5B): 2003-2014
  • 143 Koenen RR, von Hundelshausen P, Nesmelova IV. , et al. Disrupting functional interactions between platelet chemokines inhibits atherosclerosis in hyperlipidemic mice. Nat Med 2009; 15 (01) 97-103
  • 144 Koenen RR, Weber C. Therapeutic targeting of chemokine interactions in atherosclerosis. Nat Rev Drug Discov 2010; 9 (02) 141-153
  • 145 Figueiredo CR, Matsuo AL, Massaoka MH, Polonelli L, Travassos LR. Anti-tumor activities of peptides corresponding to conserved complementary determining regions from different immunoglobulins. Peptides 2014; 59: 14-19
  • 146 Figueiredo CR, Azevedo RA, Mousdell S. , et al. Blockade of MIF-CD74 signalling on macrophages and dendritic cells restores the antitumour immune response against metastatic melanoma. Front Immunol 2018; 9: 1132
  • 147 Yang L, Liu Z, Ren H. , et al. DRα1-MOG-35-55 treatment reduces lesion volumes and improves neurological deficits after traumatic brain injury. Metab Brain Dis 2017; 32 (05) 1395-1402
  • 148 Meza-Romero R, Benedek G, Leng L, Bucala R, Vandenbark AA. Predicted structure of MIF/CD74 and RTL1000/CD74 complexes. Metab Brain Dis 2016; 31 (02) 249-255
  • 149 Offner H, Sinha S, Burrows GG, Ferro AJ, Vandenbark AA. RTL therapy for multiple sclerosis: a Phase I clinical study. J Neuroimmunol 2011; 231 (1-2): 7-14
  • 150 Meza-Romero R, Benedek G, Jordan K. , et al. Modeling of both shared and distinct interactions between MIF and its homologue D-DT with their common receptor CD74. Cytokine 2016; 88: 62-70
  • 151 Crump MP, Gong JH, Loetscher P. , et al. Solution structure and basis for functional activity of stromal cell-derived factor-1; dissociation of CXCR4 activation from binding and inhibition of HIV-1. EMBO J 1997; 16 (23) 6996-7007
  • 152 Tamamura H, Hori A, Kanzaki N. , et al. T140 analogs as CXCR4 antagonists identified as anti-metastatic agents in the treatment of breast cancer. FEBS Lett 2003; 550 (1-3): 79-83
  • 153 DeMarco SJ, Henze H, Lederer A. , et al. Discovery of novel, highly potent and selective beta-hairpin mimetic CXCR4 inhibitors with excellent anti-HIV activity and pharmacokinetic profiles. Bioorg Med Chem 2006; 14 (24) 8396-8404
  • 154 Fujii N, Oishi S, Hiramatsu K. , et al. Molecular-size reduction of a potent CXCR4-chemokine antagonist using orthogonal combination of conformation- and sequence-based libraries. Angew Chem Int Ed Engl 2003; 42 (28) 3251-3253
  • 155 Ueda S, Oishi S, Wang ZX. , et al. Structure-activity relationships of cyclic peptide-based chemokine receptor CXCR4 antagonists: disclosing the importance of side-chain and backbone functionalities. J Med Chem 2007; 50 (02) 192-198
  • 156 Demmer O, Frank AO, Hagn F. , et al. A conformationally frozen peptoid boosts CXCR4 affinity and anti-HIV activity. Angew Chem Int Ed Engl 2012; 51 (32) 8110-8113
  • 157 Demmer O, Dijkgraaf I, Schumacher U. , et al. Design, synthesis, and functionalization of dimeric peptides targeting chemokine receptor CXCR4. J Med Chem 2011; 54 (21) 7648-7662
  • 158 Di Maro S, Trotta AM, Brancaccio D. , et al. Exploring the N-terminal region of C–X-C motif chemokine 12 (CXCL12): identification of plasma-stable cyclic peptides as novel, potent C–X-C chemokine receptor type 4 (CXCR4) antagonists. J Med Chem 2016; 59 (18) 8369-8380
  • 159 Di Maro S, Di Leva FS, Trotta AM. , et al. Structure-activity relationships and biological characterization of a novel, potent, and serum stable C–X-C chemokine receptor type 4 (CXCR4) antagonist. J Med Chem 2017; 60 (23) 9641-9652
  • 160 Groß A, Möbius K, Haußner C, Donhauser N, Schmidt B, Eichler J. Mimicking protein-protein interactions through peptide-peptide interactions: HIV-1 gp120 and CXCR4. Front Immunol 2013; 4: 257
  • 161 Möbius K, Dürr R, Haussner C, Dietrich U, Eichler J. A functionally selective synthetic mimic of the HIV-1 co-receptor CXCR4. Chemistry 2012; 18 (27) 8292-8295
  • 162 Wang J, Tong C, Yan X. , et al. Limiting cardiac ischemic injury by pharmacological augmentation of macrophage migration inhibitory factor-AMP-activated protein kinase signal transduction. Circulation 2013; 128 (03) 225-236
  • 163 Höllriegl W, Bauer A, Baumgartner B. , et al. Pharmacokinetics, disease-modifying activity, and safety of an experimental therapeutic targeting an immunological isoform of macrophage migration inhibitory factor, in rat glomerulonephritis. Eur J Pharmacol 2018; 820: 206-216
  • 164 Sparkes A, De Baetselier P, Brys L. , et al. Novel half-life extended anti-MIF nanobodies protect against endotoxic shock. FASEB J 2018; 32 (06) 3411-3422
  • 165 Kaufman JL, Niesvizky R, Stadtmauer EA. , et al. Phase I, multicentre, dose-escalation trial of monotherapy with milatuzumab (humanized anti-CD74 monoclonal antibody) in relapsed or refractory multiple myeloma. Br J Haematol 2013; 163 (04) 478-486
  • 166 Griffiths K, Dolezal O, Cao B. , et al. i-bodies, human single domain antibodies that antagonize chemokine receptor CXCR4. J Biol Chem 2016; 291 (24) 12641-12657
  • 167 Korkolopoulou P, Levidou G, El-Habr EA. , et al. Expression of interleukin-8 receptor CXCR2 and suppressor of cytokine signaling-3 in astrocytic tumors. Mol Med 2012; 18: 379-388
  • 168 Crichlow GV, Lubetsky JB, Leng L, Bucala R, Lolis EJ. Structural and kinetic analyses of macrophage migration inhibitory factor active site interactions. Biochemistry 2009; 48 (01) 132-139
  • 169 Rajasekaran D, Zierow S, Syed M, Bucala R, Bhandari V, Lolis EJ. Targeting distinct tautomerase sites of D-DT and MIF with a single molecule for inhibition of neutrophil lung recruitment. FASEB J 2014; 28 (11) 4961-4971
  • 170 Winner M, Meier J, Zierow S. , et al. A novel, MIF suicide substrate inhibits motility and growth of lung cancer cells. Cancer Res 2008; 68 (18) 7253-7257
  • 171 Al-Abed Y, Dabideen D, Aljabari B. , et al. ISO-1 binding to the tautomerase active site of MIF inhibits its pro-inflammatory activity and increases survival in severe sepsis. J Biol Chem 2005; 280 (44) 36541-36544
  • 172 Tynan A, Mawhinney L, Armstrong ME. , et al. Macrophage migration inhibitory factor enhances Pseudomonas aeruginosa biofilm formation, potentially contributing to cystic fibrosis pathogenesis. FASEB J 2017; 31 (11) 5102-5110
  • 173 Mawhinney L, Armstrong ME, O' Reilly C. , et al. Macrophage migration inhibitory factor (MIF) enzymatic activity and lung cancer. Mol Med 2015; 20: 729-735
  • 174 Pease J, Horuk R. Chemokine receptor antagonists. J Med Chem 2012; 55 (22) 9363-9392
  • 175 Bizzarri C, Beccari AR, Bertini R, Cavicchia MR, Giorgini S, Allegretti M. ELR+ CXC chemokines and their receptors (CXC chemokine receptor 1 and CXC chemokine receptor 2) as new therapeutic targets. Pharmacol Ther 2006; 112 (01) 139-149
  • 176 Kleemann R, Hausser A, Geiger G. , et al. Intracellular action of the cytokine MIF to modulate AP-1 activity and the cell cycle through Jab1. Nature 2000; 408 (6809): 211-216
  • 177 Nguyen MT, Beck J, Lue H. , et al. A 16-residue peptide fragment of macrophage migration inhibitory factor, MIF-(50-65), exhibits redox activity and has MIF-like biological functions. J Biol Chem 2003; 278 (36) 33654-33671
  • 178 Liang X. CXCR4, inhibitors and mechanisms of action. Chem Biol Drug Des 2008; 72 (02) 97-110
  • 179 Aboye TL, Ha H, Majumder S. , et al. Design of a novel cyclotide-based CXCR4 antagonist with anti-human immunodeficiency virus (HIV)-1 activity. J Med Chem 2012; 55 (23) 10729-10734

Zoom Image
Fig. 1 Overview of inhibitory approaches to target the macrophage migration inhibitory factor (MIF)/receptor network in atherosclerosis. The potential utility of all three classes of anti-MIF network inhibitors, that is, antibodies, small molecule drug (SMD) compounds and peptides, in attenuating atherosclerosis and/or atherogenic inflammation is indicated with respect to MIF and MIF-2/D-dopachrome tautomerase (D-DT), as well as the MIF receptors CXCR4, CD74 and CXCR2. The pros and cons for each inhibitor-type regarding each target ligand/receptor or pathway are indicated by scoring their properties (e.g. specificity) with +, +/– or –. The outcome boxes are colour-coded (red, pro-atherogenic; green, athero-/cardioprotective).