Macrophage migration inhibitory factor (MIF), also known as glycosylation-inhibiting factor, is a small, multifunctional cytokine and enzyme that serves as a key regulator of innate and adaptive immune responses, inflammation, and stress pathways.[1] Originally identified in the late 1960s as a soluble factor secreted by activated T lymphocytes that inhibits the random migration of macrophages during delayed-type hypersensitivity reactions, MIF was cloned in 1989 and identified in 1993 as a pituitary hormone released in response to stress.[2] Encoded by a single gene on human chromosome 22q11.2, MIF is a 12.5 kDa, 114-amino-acid polypeptide that forms a homotrimeric structure with intrinsic tautomerase enzymatic activity, catalyzing the tautomerization of substrates such as D-dopachrome.[3]MIF is constitutively expressed in a wide array of cells and tissues, including monocytes, macrophages, T cells, epithelial cells, and endocrine organs like the pituitary gland, with rapid secretion triggered by microbial products, inflammatory stimuli, or physiological stress.[1] Its expression is tightly regulated at transcriptional and post-transcriptional levels, often induced by cytokines such as TNF-α and counteracted by glucocorticoids, though MIF uniquely overrides glucocorticoid-mediated immune suppression to sustain proinflammatory responses.[2] Structurally conserved across species from prokaryotes to eukaryotes, MIF exerts its effects by binding to cell surface receptors, including CD74 (the invariant chain of MHC class II), CXCR2, CXCR4, and CXCR7, thereby activating downstream signaling pathways like ERK1/2 MAPK and promoting leukocyte recruitment, survival, and effector functions.[3]In innate immunity, MIF plays a pivotal role by upregulating Toll-like receptor 4 (TLR4) expression on macrophages, enhancing endotoxin responsiveness, and driving the production of proinflammatory mediators such as TNF-α, IL-1β, and nitric oxide, which are crucial for pathogen clearance.[1] Beyond immunity, MIF influences diverse physiological processes, including glucose metabolism, cell proliferation, and apoptosis suppression via p53 inhibition, while its enzymatic activity may contribute to oxidative stress modulation.[2] However, dysregulated MIF activity is implicated in numerous pathologies; elevated levels correlate with severity in conditions like sepsis, acute respiratory distress syndrome (ARDS), rheumatoid arthritis, and glomerulonephritis, where it exacerbates inflammation and tissue damage.[1]In infectious diseases, MIF exhibits a dual nature: protective in controlling intracellular pathogens such as Mycobacterium tuberculosis and Toxoplasma gondii by bolstering macrophage bactericidal activity, yet detrimental in overwhelming infections like pneumococcal meningitis or HIV, where it amplifies cytokine storms and viral replication.[3] Similarly, in cancer and autoimmune disorders, MIF promotes tumor growth, angiogenesis, and metastasis while sustaining chronic inflammation, positioning it as a promising therapeutic target with inhibitors like ISO-1 and SCD-19 under investigation for sepsis, rheumatoid arthritis, and oncology applications.[2] As of 2025, MIF inhibitors continue to be investigated in combination with immunotherapies for cancer and other inflammatory conditions.[4] Genetic polymorphisms, such as the -173 G/C SNP in the MIF promoter, further modulate its expression and disease susceptibility, underscoring its clinical relevance as a biomarker and intervention point.[2]
Discovery and History
Initial Identification
Macrophage migration inhibitory factor (MIF) was first identified in 1966 by Barry R. Bloom and B. Bennett as a soluble factor secreted by sensitized lymphocytes that inhibits the random migration of macrophages in vitro. Independently in the same year, John R. David described a comparable activity derived from lymphoid cells interacting with antigens, further establishing MIF as a key mediator in cellular immune responses.[5]The discovery arose from investigations into delayed-type hypersensitivity reactions using guinea pig models. In these early experiments, researchers sensitized guinea pigs to specific antigens and then stimulated their lymphocytes in vitro with the corresponding antigen; the resulting supernatants were collected and added to cultures containing macrophages packed into capillary tubes. These supernatants prevented the macrophages from migrating out of the tubes and spreading on the culture surface, demonstrating a specific inhibition of macrophage motility dependent on prior lymphocyte sensitization and antigen stimulation.This inhibitory effect led to the naming of the factor as "macrophage migration inhibitory factor," reflecting its primary observed function in restricting macrophage movement during cell culture assays. MIF was promptly recognized as a T-cell-derived lymphokine, marking it as one of the earliest identified soluble mediators in the burgeoning field of cytokine research at the time.[5]
Molecular Characterization
In the late 1980s, efforts to purify macrophage migration inhibitory factor (MIF) from supernatants of lectin-stimulated human T-cell hybridomas led to the isolation of a protein exhibiting MIF activity, estimated at approximately 12.5 kDa by SDS-PAGE analysis.[6] Partial amino acid sequencing of the N-terminal region provided 13 residues, enabling the design of oligonucleotide probes for subsequent cloning efforts and confirming MIF as a distinct polypeptide without a classical signal peptide for secretion.[6]The human MIF cDNA was cloned in 1989 through functional expression screening in COS-1 cells, using a library derived from the same T-cell hybridoma (T-CEMB).[6] The isolated cDNA encoded a 115-amino acid protein that, when expressed, produced bioactive MIF capable of inhibiting macrophage migration in vitro, distinguishing it from other known lymphokines due to its unique sequence and lack of homology to previously identified cytokines.[6] This cloning revealed MIF as a proinflammatory cytokine with broad implications beyond initial immune functions.In 1991, MIF was rediscovered as a pituitary hormone secreted in response to stress, linking its immune regulatory role to endocrine functions.[2]Subsequent biochemical studies in the mid-1990s uncovered unexpected enzymatic activities for MIF, expanding its functional profile to non-immune roles. In 1996, MIF was demonstrated to possess tautomerase activity, catalyzing the conversion of D-dopachrome methyl ester to 5,6-dihydroxyindole-2-carboxylic acid methyl ester, with related activity on L-dopachrome methyl ester substrates, though native L-dopachrome was not a direct substrate.[7] A year later, MIF was identified as a phenylpyruvate tautomerase (EC 5.3.2.1), efficiently interconverting phenylpyruvate and its enoltautomer, as well as p-hydroxyphenylpyruvate, suggesting potential involvement in metabolic pathways like tyrosinecatabolism.[8]The MIF gene was mapped to human chromosome 22q11.2 in the mid-1990s through fluorescence in situ hybridization and somatic cell hybrid analysis, confirming its location in a region syntenic with the mouse Mif locus on chromosome 10.[9] This small gene, spanning less than 1 kb with three exons and short introns of 189 bp and 95 bp, exhibits constitutive expression across diverse human tissues, including immune cells, epithelial tissues, and endocrine organs, as evidenced by Northern blot analyses detecting an ~800 nt mRNA transcript ubiquitously.[10]
Structure and Expression
Gene and Protein Structure
The human MIF gene is located on chromosome 22q11.23 and spans less than 1 kb in length. It consists of three exons measuring 107 bp, 172 bp, and 66 bp, separated by two short introns of 189 bp and 95 bp, respectively.[10][11]The MIF protein is synthesized as a 115-amino acid polypeptide with a calculated molecular weight of approximately 12.3 kDa. It is nonglycosylated and assembles into stable homotrimers primarily through hydrophobic interactions at the subunit interfaces.[12][13] Each monomer adopts a compact fold featuring two antiparallel α-helices packed against a twisted four-stranded β-sheet, with the trimer forming a central β-barrel topology lined by the β-sheets from all three subunits. The protein lacks intramolecular disulfide bonds, though it contains two conserved cysteine residues (Cys-57 and Cys-60) that can form an intersubunit disulfide under oxidative stress. A key structural feature is the conserved N-terminal Pro1-Leu2 motif, which constitutes the active site for the protein's intrinsic tautomerase activity.[13][14]MIF undergoes limited post-translational modifications, with no significant glycosylation observed in its mature form. Notably, the protein is synthesized without a classical N-terminal signal peptide, instead employing unconventional secretion pathways—such as ABC transporter-mediated export or direct translocation across the plasma membrane—to achieve extracellular release.[15][16]
Expression Patterns
Macrophage migration inhibitory factor (MIF) is constitutively expressed across a wide range of immune cells, including monocytes, macrophages, T cells, B cells, neutrophils, eosinophils, and dendritic cells, as well as in non-immune tissues such as the pituitary gland, liver, kidney, lung, skin, and gastrointestinal tract.[1][17]63797-2/fulltext) This baseline expression enables rapid release upon cellular activation, distinguishing MIF from many other cytokines that require de novo synthesis.[18]MIF expression is dynamically upregulated in response to inflammatory and stress stimuli, including lipopolysaccharide (LPS) from bacterial sources, tumor necrosis factor-alpha (TNF-α), and interferon-gamma (IFN-γ).[18][19] The MIF gene promoter includes glucocorticoid-responsive elements (GREs), which facilitate induction during glucocorticoid-mediated stress responses, such as those involving cortisol.[20]63797-2/fulltext)As a leaderless protein lacking a classical signal peptide, MIF is secreted via non-conventional pathways, including ABC transporter-mediated export and endosomal or vesicular trafficking involving proteins like p115.[21][22] In some cellular contexts, MIF is retained intracellularly, particularly within cytoplasmic vesicles or the nucleus, prior to release.[23][1]In healthy individuals, circulating plasma MIF levels typically range from 5 to 10 ng/mL.[24] These concentrations can elevate substantially during inflammatory states, such as sepsis, reaching over 100 ng/mL in severe cases.[25]
Macrophage migration inhibitory factor (MIF) plays a central role in orchestrating both innate and adaptive immune responses, acting as a key regulator that promotes antimicrobial defense and sustains immune cell function during infection. Constitutively expressed by immune cells such as macrophages and T lymphocytes, MIF is rapidly secreted upon pathogen recognition, enabling it to amplify early inflammatory signals and counteract suppressive mechanisms that could impair host defense.[18] This positions MIF as an essential component of the innate immune alarm system, where it enhances the proinflammatory activities of leukocytes to facilitate pathogen clearance.[1]In innate immunity, MIF activates macrophages by enhancing phagocytosis, nitric oxide (NO) production, and antigen presentation, often in synergy with Toll-like receptor 4 (TLR4) signaling. Specifically, MIF stimulates macrophage engulfment of pathogens and boosts inducible NO synthase expression, thereby increasing bactericidal activity against intracellular microbes like Mycobacterium tuberculosis.[26] Through upregulation of TLR4 expression via ETS transcription factors, MIF potentiates lipopolysaccharide (LPS)-induced responses, including improved major histocompatibility complex (MHC) class II-mediated antigen presentation to T cells, which bridges innate and adaptive immunity.[27] These effects sustain macrophage survival by inhibiting p53-dependent apoptosis, ensuring prolonged antimicrobial function at infection sites.MIF also modulates adaptive immunity by influencing T-cell responses, promoting differentiation into proinflammatory Th1 and Th17 subsets while countering apoptosis to enhance T-cell persistence. It drives Th1 polarization by supporting interleukin-2 (IL-2) production and CD74-CD44 signaling, which favors interferon-γ secretion and cell-mediated immunity against intracellular pathogens.[28] Similarly, MIF promotes Th17 differentiation through NF-κB pathway activation and interactions with CD74, amplifying IL-17 responses critical for defense against extracellular bacteria and fungi.[29] By overriding apoptosis via MAPK activation, MIF sustains activated T-cell survival, preventing premature exhaustion during chronic infections.A hallmark of MIF's immune regulatory function is its counteraction of glucocorticoid-mediated suppression, preserving host defense under stress conditions like infection or trauma. Glucocorticoids, such as cortisol, typically inhibit cytokine production and immune cell migration to dampen inflammation, but MIF overrides these effects by sustaining TNF-α and IL-1β secretion from macrophages, thereby maintaining proinflammatory responses. This antagonism occurs at the transcriptional level, where MIF blocks glucocorticoid receptor-induced NF-κB inhibition, ensuring immune competence during physiological stress.[30] In early infection, MIF's rapid release from pathogen-exposed immune cells amplifies cytokine storms, particularly by upregulating IL-1β and TNF-α production, which recruits additional leukocytes and initiates a robust innate response.[18]
Role in Inflammation and Beyond
Macrophage migration inhibitory factor (MIF) exhibits prominent pro-inflammatory actions by sustaining nuclear factor kappa B (NF-κB) activity, which drives the expression of pro-inflammatory cytokines such as tumor necrosis factor-alpha (TNF-α) and interleukin-6 (IL-6) in immune cells.[31] Additionally, MIF inhibits p53-mediated apoptosis in macrophages, thereby prolonging their survival and enhancing sustained inflammatory responses during innate immune activation, as demonstrated in lipopolysaccharide-challenged models where MIF deficiency leads to increased macrophage apoptosis and reduced pro-inflammatory cytokine production.[32] MIF also promotes leukocyte recruitment to sites of inflammation by upregulating adhesion molecules and chemokines, facilitating the influx of neutrophils and monocytes to amplify local immune reactions.[33]MIF displays a dual, Janus-faced role in inflammation, providing protective effects in acute settings by supporting rapid immune containment and tissue repair, while contributing to pathogenesis in chronic inflammation, such as in atherosclerosis where it exacerbates plaque instability through persistent monocyte accumulation and cytokine release.[31] This duality arises from context-dependent signaling, where MIF's pro-survival effects aid resolution in short-term responses but fuel unresolved inflammation in prolonged conditions.[34] Notably, MIF counteracts the anti-inflammatory actions of glucocorticoids, overriding their suppression of cytokine production in activated immune cells.[30]Beyond core immunity, MIF supports non-immune functions including angiogenesis through induction of vascular endothelial growth factor (VEGF) expression in various cell types, promoting new vessel formation essential for tissue vascularization.[35] It also stimulates cell proliferation in fibroblasts and endothelial cells; for instance, MIF enhances fibroblast migration and proliferation in wound healing models, accelerating closure of scratch-wounded monolayers to 86.6% at 24 hours compared to 54.7% in controls, associated with an increase to approximately 62% Ki67-positive cells.[36] Similarly, MIF acts as a potent mitogen for endothelial cells, driving their growth and contributing to angiogenic processes in tissue remodeling.[37] These effects collectively aid wound healing by facilitating the transition from inflammation to proliferative repair phases.In metabolic contexts, MIF regulates insulin signaling in adipocytes by inhibiting tyrosine phosphorylation of insulin receptor substrate-1 (IRS-1) and AKT activation, thereby impairing glucose uptake and contributing to insulin resistance in obesity.[38] Elevated MIF levels in obese individuals correlate with higher body mass index and adipose inflammation, while MIF deficiency improves glucose homeostasis and reduces diet-induced insulin resistance in mouse models.[39] These metabolic links highlight MIF's broader influence on energy balance and adipose tissue function beyond inflammatory roles.[40]
Mechanism of Action
Receptor Binding and Signaling
Macrophage migration inhibitory factor (MIF) primarily engages cells through binding to CD74, the invariant chain of major histocompatibility complex class II molecules, which serves as its main receptor. This interaction occurs at the extracellular domain of CD74, initiating signal transduction by inducing phosphorylation of CD74 and subsequent recruitment of CD44 as a co-receptor. The CD74-CD44 complex then activates Src-family kinases, such as Lyn and Fyn, leading to downstream intracellular signaling cascades essential for MIF's proinflammatory effects.[41][42]In addition to CD74, MIF interacts with chemokine receptors, including CXCR2, CXCR4, and CXCR7, often forming heterocomplexes with CD74 to modulate cell migration and other responses. Binding to CXCR2 promotes pro-angiogenic activities by enhancing endothelial cell proliferation and vascularization in inflammatory contexts. Interactions with CXCR4 and CXCR7 facilitate cell migration, particularly in immune and tumor cells, where MIF-CXCR4 engagement drives chemotaxis and metastasis. These receptor complexes enable MIF to exert chemokine-like functions, distinct from its primary CD74-mediated pathway.[43][44][45]Upon receptor binding, MIF triggers several key downstream signaling pathways, notably sustained activation of the ERK1/2 mitogen-activated protein kinase (MAPK) pathway, characterized by prolonged phosphorylation that persists for hours, unlike the transient activation by many growth factors. This sustained ERK1/2 signaling promotes cell proliferation and survival. MIF also activates the PI3K/Akt pathway, enhancing anti-apoptotic effects and cell viability in various cell types, including melanoma and immune cells. Furthermore, MIF induces activation of NF-κB, contributing to inflammatory gene expression.[46][47] Recent studies (as of 2025) have further elucidated MIF's role in modulating AMP-activated protein kinase (AMPK) signaling and its contributions to metabolic regulation in inflammation.[48]At the CXCR4 receptor, MIF functions as a partial allosteric agonist, binding to an allosteric site distinct from the orthosteric site occupied by the full agonist CXCL12, resulting in weaker G-protein coupling and β-arrestin recruitment compared to CXCL12. This modulation allows MIF to fine-tune CXCR4 signaling, supporting sustained but attenuated responses in migration and inflammation. The trimeric structure of MIF facilitates these receptor interactions by presenting multiple binding interfaces.[43]
Enzymatic Activity
Macrophage migration inhibitory factor (MIF) possesses intrinsic tautomerase activity, catalyzing the keto-enol tautomerization of specific substrates, including phenylpyruvate and L-dopachrome methyl ester.[49] This enzymatic function relies on the N-terminal proline residue (Pro1), which acts as the catalytic base by facilitating proton abstraction and transfer during the reaction. Site-directed mutagenesis studies, such as substituting Pro1 with serine (P1S), completely abolish this tautomerase activity, confirming its essential role.[50]The physiological relevance of MIF's tautomerase activity remains under investigation but may involve regulation of melanogenesis through the conversion of L-dopachrome methyl ester to its keto form, a step in melanin synthesis pathways, and modulation of small molecule metabolism by processing oxidized catecholamines like 3,4-dihydroxyphenylaminechrome and norepinephrinechrome, which are toxic byproducts in oxidative environments.[49][51] This activity can be selectively inhibited by small molecules such as ISO-1 (4-iodo-6-phenyl-3-hydroxy-1-tetralone), a covalent modifier that binds the active site with an IC50 of approximately 7 μM, thereby blocking tautomerization without directly targeting cytokine functions.MIF's enzymatic site, centered around Pro1, is distinct yet partially overlapping with the trimer interface required for its cytokine oligomerization, allowing for functional separation.[52] Mutations that eliminate tautomerase activity, such as the P1G variant, do not fully impair MIF's signaling capabilities, as demonstrated in models where tautomerase-null MIF retains growth-regulatory and proinflammatory effects.[53] This independence suggests the tautomerase function operates as a non-cytokine role, potentially contributing to intracellular metabolitehomeostasis.Emerging evidence links MIF's tautomerase activity to oxidative stress responses in inflamed tissues, where it may mitigate damage by modulating reactive metabolites derived from catecholamine oxidation, thus protecting cells from quinone-induced toxicity during inflammatory episodes.[51] In tautomerase-deficient models, reduced capacity to handle such metabolites correlates with exacerbated oxidative injury, highlighting a protective enzymatic mechanism alongside MIF's immune roles.[53]
Molecular Interactions
Receptor Interactions
Macrophage migration inhibitory factor (MIF) primarily interacts with the cell surface receptor CD74, a type IIintegral membrane protein, through high-affinity binding to its extracellular domain. This interaction exhibits a dissociation constant (Kd) of approximately 9 × 10^{-9} M, enabling stable complex formation that facilitates subsequent intramembrane proteolysis of CD74.[54] The binding engages specific residues on MIF's N-terminal region, promoting the recruitment of accessory proteins to form a signalosome complex on the cell membrane.MIF also binds to members of the CXC chemokine receptor family, including CXCR2, CXCR4, and CXCR7, with nanomolar affinities that support its role as a noncognate ligand. For instance, MIF interacts with CXCR2 to mediate neutrophilchemotaxis, while binding to CXCR4 influences stem cell homing; MIF binding to CXCR7 promotes receptor internalization, ERK1/2 signaling, and lymphocytechemotaxis.[55][56] These interactions occur via a two-site binding model on the receptors' extracellular loops, distinct from classical chemokine engagement.[57]Co-receptor dynamics are critical for MIF's receptor specificity, with CD74 forming heterodimers with CXCR2 or CXCR4 on macrophages and endothelial cells to enable full activation. These complexes, confirmed through co-immunoprecipitation and microscopy, require surface expression of CD74 variants lacking endoplasmic reticulum retention signals.[58] MIF distinguishes itself from classical ELR+ chemokines, which possess an N-terminal Glu-Leu-Arg motif for CXCR2 binding, by lacking this sequence yet achieving similar effects through allosteric modulation via a pseudo-ELR motif (Asp44-Xaa-Arg11). This structural adaptation allows MIF to promote angiogenesis via the MIF-CXCR2 axis without direct sequence homology.[55]
Intracellular Partners
Macrophage migration inhibitory factor (MIF) primarily interacts intracellularly with JAB1 (also known as CSN5), a subunit of the COP9 signalosome complex, in both cytosolic and nuclear compartments. This binding sequesters JAB1, thereby inhibiting its function as a coactivator of the AP-1 transcription factor and its role in promoting the degradation of c-Jun, a key component of AP-1. By antagonizing JAB1-mediated activation of AP-1 and stabilization of phospho-c-Jun, MIF reduces AP-1 transcriptional activity, which modulates gene expression related to cell proliferation and survival.[59] The interaction is mediated by specific residues in MIF, including amino acids 50–65 and cysteine 60, and occurs independently of MIF's enzymatic tautomerase activity.[60]MIF's binding to JAB1/CSN5 also influences proteasomal degradation pathways, as the COP9 signalosome regulates cullin-RING ubiquitin ligases involved in protein turnover. Through this association, MIF modulates the degradation of substrates such as the cyclin-dependent kinase inhibitor p27^{Kip1}, leading to its stabilization and subsequent cell cycle regulation. Additionally, MIF suppresses p53-mediated apoptosis by directly interacting with p53 to inhibit its transcriptional activity and nuclear translocation, while enhancing the p53-Mdm2 association that promotes p53 ubiquitination and degradation; this effect is independent of but complementary to JAB1 binding. These mechanisms collectively promote cell survival by counteracting pro-apoptotic signals.[61]In T cells, the MIF-JAB1 axis sustains proliferation by suppressing activation-induced apoptosis through p53 inhibition, allowing continued immune responses. Disruptions in this interaction, such as by MIF inhibitors or knockdown, restore p53 function, leading to increased apoptosis in cancer cell models, including lymphomas and pancreatic tumors, highlighting the axis's role in tumor survival.[62] Internalized MIF, following CD74-mediated endocytosis, can translocate to the nucleus where it further engages JAB1 to fine-tune gene expression, though specific transcriptional targets like TNF-α enhancement remain context-dependent.[63]
Clinical Significance
Involvement in Diseases
Macrophage migration inhibitory factor (MIF) plays a pathological role in various inflammatory and autoimmune diseases, where its dysregulation contributes to disease progression. In rheumatoid arthritis (RA), MIF is overexpressed in synovial tissues and fluids, with concentrations often exceeding 50 ng/mL in patients with erosive disease, promoting joint destruction through enhanced production of matrix metalloproteinases and angiogenic factors by synovial fibroblasts.[64][65] In sepsis, elevated serum MIF levels correlate with increased mortality risk, serving as an early biomarker of poor outcome due to its amplification of systemic inflammation and counter-regulation of glucocorticoid-mediated immunosuppression.[66][67]In cardiovascular diseases, MIF drives atherosclerosis by promoting monocyte recruitment and plaque instability through inflammatory signaling in vascular endothelium and smooth muscle cells.[68] Post-myocardial infarction, MIF exacerbates adverse ventricular remodeling, while recent 2025 studies link elevated circulating MIF to systolic dysfunction and progression in heart failure with reduced ejection fraction.[69][70]MIF contributes to oncogenesis in multiple cancers by fostering a tumor-supportive microenvironment. In colorectal cancer (CRC), breast cancer, and prostate cancer, MIF promotes tumor growth and metastasis via induction of angiogenesis and suppression of apoptosis, with high serum MIF levels emerging as a marker of poor prognosis, including reduced overall survival in advanced CRC.[71][72][73]In infections and neurological disorders, MIF exacerbates septic shock by sustaining hyperinflammation and impairing immune resolution, while in viral infections like COVID-19, it boosts SARS-CoV-2 replication through activation of pro-inflammatory pathways in infected cells.[1][74] Emerging evidence highlights MIF's role in neuroinflammation, where it inhibits microglial clearance of amyloid-beta plaques, contributing to Alzheimer's disease progression and neuronal damage.[75][76]Beyond these, MIF is implicated in metabolic and gastrointestinal pathologies. In obesity and type 2 diabetes, adipose-derived MIF induces insulin resistance by promoting chronic low-grade inflammation and impairing glucose uptake in adipocytes and hepatocytes.[77] In inflammatory bowel disease (IBD), elevated MIF in active lesions drives mucosal inflammation and epithelial barrier dysfunction, correlating with flare-ups and increased disease severity in both ulcerative colitis and Crohn's disease.[78][79]
Therapeutic Potential
Macrophage migration inhibitory factor (MIF) has emerged as a promising therapeutic target due to its overexpression in various inflammatory and neoplastic conditions, prompting the development of targeted inhibitors to mitigate its pro-inflammatory and tumor-promoting effects. Small molecule inhibitors such as ISO-1, which blocks MIF's tautomerase activity, have demonstrated efficacy in preclinical models of sepsis and cancer by reducing proinflammatory cytokine release and tumor cell proliferation. Similarly, CPSI-1306, a small moleculeantagonist of the MIF-CD74 interaction, has shown potential in inhibiting T-cell suppression within the tumor microenvironment and reducing tumor growth in head and neck squamous cell carcinoma models.[80][81][82]Monoclonal antibodies targeting MIF, particularly imalumab (BAX69), an anti-oxidized MIF antibody, have advanced to clinical testing. Imalumab completed Phase I trials in advanced solid tumors, including colorectal cancer (CRC), where it exhibited acceptable tolerability and preliminary antitumor activity, leading to a Phase Ib/IIa proof-of-concept study in CRC patients. As of 2025, a novel small molecule MIF inhibitor, IPG1094, has entered Phase I/II trials for advanced solid tumors, including glioblastoma and lung cancer brain metastases, demonstrating tumor size reduction and blood-brain barrier penetration in pilot studies. In sepsis models, anti-MIF strategies, including small molecule inhibitors, have reduced cytokine storm severity, though human trials remain preclinical or early-phase. For cardiovascular disease (CVD), MIF blockade in post-myocardial infarction (MI) models has been shown to ameliorate adverse cardiac remodeling, highlighting potential in limiting post-infarct remodeling.[83][84][84][85]Alternative strategies include gene silencing approaches, such as siRNA-mediated MIF knockdown, which has attenuated tumor maintenance in CRC models by disrupting epithelial cell survival pathways. Combining MIF inhibition with immune checkpoint inhibitors, like anti-PD-1, has enhanced antitumor immunity in preclinical melanoma and mesothelioma models, leading to complete tumor regression and improved survival by countering MIF's immunosuppressive effects. These combinations address resistance to single-agent immunotherapy, promoting T-cell infiltration and activation.[71][86]Despite these advances, MIF's dual pro- and anti-inflammatory roles necessitate context-specific targeting to avoid exacerbating protective functions, such as in acute immune responses. Elevated plasma MIF levels serve as a biomarker for patient stratification, with concentrations above typical thresholds (e.g., >10-20 ng/mL in sepsis or cancer) correlating with disease severity and predicting responsiveness to MIF-targeted therapies in observational studies. Ongoing challenges include optimizing delivery for tissue-specific inhibition and validating biomarkers in larger cohorts to refine therapeutic selection.[87][88][29]
Related Proteins and Homologs
D-DT as a Functional Homolog
D-dopachrome tautomerase (D-DT), also known as MIF-2, is encoded by the DDTgene located on humanchromosome 22q11.23.[89] The gene produces a protein of 118 amino acids with a molecular weight of approximately 13 kDa.[89] D-DT shares 35% amino acid sequence identity with macrophage migration inhibitory factor (MIF) and exhibits a conserved gene structure, including identical exon-intron organization.[90] Structurally, D-DT folds into a homotrimeric assembly with a beta-barrel motif akin to MIF, facilitating similar quaternary interactions and functional properties.[91]D-DT and MIF overlap in several biological activities, particularly in immune regulation. Both proteins bind the cell surface receptor CD74, though D-DT exhibits threefold lower affinity (higher Kd), leading to rapid association and dissociation kinetics.[92] This binding activates downstream signaling pathways, including phosphorylation of ERK1/2 and Akt, which promote proinflammatory responses and inhibit random macrophage migration.[93] Additionally, D-DT possesses intrinsic tautomerase activity on substrates like D-dopachrome methyl ester, albeit approximately tenfold less potent than MIF.[92]While sharing functions with MIF, D-DT displays distinct tissue-specific roles, notably in neural and adipose contexts. D-DT is constitutively expressed in the brain and other neural tissues, where it contributes to anatomical distribution patterns observed in murine models, potentially influencing neuroinflammatory processes.[94] In adipose tissue, D-DT acts as an adipokine that regulates lipid metabolism via AMPK and PKA pathways, enhancing glucose tolerance in obesity models and modulating adipogenesis.[95] D-DT also synergizes with MIF in systemic inflammation; studies using Mif/D-dt double-knockout mice demonstrate reduced sepsis severity, including lower proinflammatory cytokine levels and improved survival compared to single knockouts, underscoring their cooperative role in endotoxemia.[96][93]Recent investigations as of 2025 highlight D-DT's compensatory mechanisms in oncogenesis, particularly when MIF is depleted. In tumor models, D-DT sustains proliferation and angiogenesis in MIF-deficient settings by similarly antagonizing p53 and activating survival pathways, thereby mitigating the antitumor effects of MIF inhibition alone.[97] This redundancy supports the rationale for co-targeting both proteins in cancer therapies; dual inhibition via small molecules or antibodies enhances immune infiltration, suppresses tumor growth, and improves responses to checkpoint blockade in melanoma and other solid tumors.[98][86] As of November 2025, recent studies in melanoma, head and neck squamous cell carcinoma, and colorectal cancer reinforce the benefits of dual inhibition for enhanced antitumor immunity and survival.[99][100][71]
Parasite-Derived MIF Homologs
Macrophage migration inhibitory factor (MIF) homologs have been identified in various parasites, including nematodes, protozoans, and apicomplexans, where they contribute to immune modulation during infection. Notable examples include the filarial nematode Brugia malayi, which produces two isoforms, MIF-1 and MIF-2; the malaria parasite Plasmodium falciparum; the kinetoplastid Leishmania species; and the apicomplexan Toxoplasma gondii. These homologs enable parasites to interfere with host immune responses, promoting survival within the host.[101][102][103][104]Structurally, parasite-derived MIF homologs exhibit 26-44% amino acid sequence identity to human MIF, with conserved features such as the N-terminal proline residue essential for enzymatic function. They typically form homotrimers, as observed in P. falciparum MIF (PfMIF) and T. gondii MIF (TgMIF), and retain tautomerase activity, converting substrates like L-dopachrome methyl ester at varying rates relative to human MIF (e.g., ~1% for TgMIF and ~80% for some filarial homologs), though some, like TgMIF, lack oxidoreductase activity due to the absence of a CXXC motif. While many preserve cytokine-like properties, certain homologs show reduced or altered signaling capacity compared to the human protein.[102][104][101][105]These homologs modulate host immunity by binding to the human MIF receptor CD74, activating pathways such as ERK1/2 and PI3K/Akt to suppress macrophage activation and promote Th2-biased responses. For instance, B. malayi MIF isoforms bind CD74 and inhibit nitric oxide (NO) production in macrophages by downregulating iNOS expression, thereby reducing inflammatory killing of parasites. Similarly, Leishmania major MIF (LmMIF) prevents activation-induced apoptosis in host macrophages, while PfMIF and TgMIF enhance proinflammatory cytokine release like TNF-α and IL-8 but dampen adaptive responses such as T-cell proliferation.[105][106][103][104]In their pathogenic roles, these MIF homologs aid parasite persistence by dampening excessive inflammation and evading innate immunity, as seen in Leishmania infections where they promote intracellular survival. Recent studies highlight their vaccine potential; for example, deletion of MIF genes from live-attenuated Leishmania donovani strains enhances CD4+ T-cell immunity and reduces parasite burdens in challenged mice, suggesting targeting parasite MIF could improve clearance in leishmaniasis.[105][107] As of 2025, recombinant EtMIF shows promise as a vaccine against chicken coccidiosis, boosting immunity.[108]