Skip to main content

MIF contribution to progressive brain diseases

Abstract

Progressive brain diseases create a huge social and economic burden on modern societies as a major cause of disability and death. Incidence of brain diseases has a significantly increasing trend and merits new therapeutic strategies. At the base of many progressive brain malfunctions is a process of unresolved, chronic inflammation. Macrophage migration inhibitory factor, MIF, is an inflammatory mediator that recently gained interest of neuro-researchers due to its varied effects on the CNS such as participation of nervous system development, neuroendocrine functions, and modulation of neuroinflammation. MIF appears to be a candidate as a new biomarker and target of novel therapeutics against numerous neurologic diseases ranging from cancer, autoimmune diseases, vascular diseases, neurodegenerative pathology to psychiatric disorders. In this review, we will focus on MIF’s crucial role in neurological diseases such as multiple sclerosis (MS), Alzheimer’s disease (AD) and glioblastoma (GBM).

Introduction

During immune responses, a galaxy of factors are activated to maintain homeostasis. Disease occurs when the homeostasis is broken, and inflammation persists. One of the inflammatory mediators with broad influence on host immune responses is MIF. MIF, a cytokine, hormone and enzyme, was first described almost a half century ago in association with delayed-type hypersensitivity [1, 2]. At that time, Bloom and Bennett in an in vitro study showed the inhibition of macrophage migration mediated by a soluble factor produced by sensitized lymphocytes upon interaction with a specific antigen [1]. In 1989, Weiser et al. isolated a cDNA encoding human MIF, which allowed further studies on its physiological and pathological characteristics [3]. The identification of MIF’s receptors confirmed its chemokine-like functions and major regulatory abilities in host immune defense and inflammation [4]. According to Human Protein Atlas (HPA) (www.proteinatlas.org), MIF and its partners (e.g., CXCR4, CD74, p53, JAB1) are expressed in all tissues with high levels in bone marrow, reproductive, brain and lymphoid tissues [5,6,7]. In the Th2 subset of lymphocytes and monocytes/macrophages, in contrast to other cytokines, MIF is present in high intracytoplasmic levels in unstimulated cells. Notably, at least two orders in magnitude lower LPS is needed to stimulate expression of MIF than that in case of TNF-α [8]. In the CNS, MIF is secreted by neurons, Schwann cells, microglia, astrocytes and oligodendrocytes, predominantly in the cortex layer [9, 10]. MIF, a multifaceted cytokine is involved in immune response and numerous biological processes like proliferation, angiogenesis, anti-oxidant signaling and tissue repair [11, 12]. MIF has been identified as a critical controller of inflammation and immune responses by counter-regulating immunosuppressive effect of glucocorticoids [8, 13,14,15]. Production during systemic inflammation and glucocorticoid regulatory abilities of MIF are in tight conjunction with the hypothalamic–pituitary–adrenal axis. As a pituitary derived stress hormone, MIF became recognized as a key neuroendocrine regulator of immune responses in the CNS [13, 16, 17]. Part of MIF’s pathologic role in inflammatory diseases may be based on neuroendocrine mechanisms [18]. MIF functions in cell proliferation and glucocorticoid action by MAPK (mitogen-activated protein kinases) and cytoplasmic phospholipase A2 activation (cPLA2) [19] (Fig. 1). MIF directly interacts with intracellular protein, JAB1, a coactivator of AP-1 transcription, negatively modulating JAB-1 controlled pathways including cell-cycle regulation [20]. The discovery that MIF is capable of influencing cell proliferation by inactivating p53 tumor suppressor activity may suggest a direct link between inflammation and cancer [21]. MIF exerts its activity via specific receptors including its cognate receptor CD74, and via non-cognate interactions with CXCR2, CXCR4 or CXCR7, or possibly via endocytic engulfment to the cytoplasm [22,23,24]. MIF binds to the extracellular domain of CD74, initiating CD44 activation and the extracellular signal-regulated kinase-1/2 MAP (ERK1/2 MAP) kinase cascade and downstream inflammatory pathways leading to cell proliferation, cell survival and PGE2 production [25] (Fig. 1). Particularly, linkage of MIF with CD74/CD44 regulates the immune response by controlling maintenance, proliferation and survival of macrophages and B cells [26]. Recently, a product of CD74 cleavage, a CD74 cytosolic intracellular domain (CD74-ICD) has been found to induce cell–cell signaling and cell survival in B cells [27]. MIF shares biological functions with its homologue, D-dopachrome tautomerase (D-DT or MIF-2), which in a similar way activates the inflammatory responses [28, 29]. According to HPA, D-DT but not MIF, is greatly produced by white blood cells which suggests that D-DT possesses an especially important role in innate inflammation through CD74 [12]. As for the genetic regulation of MIF expression, there are two promoter polymorphisms located in the MIF gene: (1) functional alleles of a − 794 CATT5–8 microsatellite repeat that modulates MIF mRNA transcription and correlates with MIF expression levels. MIF promoter activity is proportional to increased numbers of the CATT repeats at position − 794 [30]: and (2) the nearby − 173 SNP C allele may be associated with increased MIF promoter activity by its linkage disequilibrium with the high-expression − 794 CATT7 variant.

Fig. 1
figure 1

MIF function and signaling. MIF fulfills its biological functions through membrane receptors and via binding to intracellular molecules. MIF’s binding to membrane receptor CD74 recruits CD44 and leads to activation of Src/MAPK signaling. MIF via CXCR2/4 activates PI3K/Akt downstream signaling and induces cell migration. Sustained activation of ERK1/2 phosphorylation is mediated by JUN activation domain binding protein-1 (JAB1) and leads to cytoplasmic phospholipase A2 (cPLA2) activity (blocked by glucocorticoids) and further to arachidonate/prostaglandin production. MIF production can be stimulated via TLRs by e.g., LPS stimulation. MIF regulates innate immune responses through modulation of TLR4. In response to LPS and Gram-negative bacteria MIF upregulates TLR-4 expression and in consequence induces the production of pro-inflammatory cytokines. MIF overrules glucocorticoid effects such as the nuclear factor-κB (NF-κB) inhibitor IκB which downregulates inflammatory responses. MIF via p53 inhibits activation-induced apoptosis, increase cellular survival and proliferation. MIF’s functions include: 1. stimulation of proinflammatory and co-stimulatory factors; 2. activation of adhesion molecules; 3. increase of cell trafficking to the sites of inflammation; 4. increase of cell proliferation and survival and inhibition of apoptosis

MIF expression has long been associated with certain diseases such as rheumatoid arthritis, asthma and cancer with increased levels found in more aggressive forms of such diseases [31,32,33,34]. In SLE, MIF has been found to play a dual role. High-expression MIF polymorphisms have been found to be associated with a lower incidence of SLE. However, in patients with established disease, low-expression MIF polymorphisms have been linked with a lower incidence of the end-organ injury [35]. Brain disorders from autoimmunity, stroke, cancer and dementia are characterized by an inflammatory component, and MIF has a detrimental effect on these pathologies. In MS, AD and GBM, MIF contributes to the severity of disease [36]. However, its pathological role in brain diseases became challenged based on some recent studies. In Parkinson’s disease (PD), MIF has been found to mediate a neuroprotective effect by suppressing inflammatory responses, inhibiting apoptosis, and inducing autophagy [37]. Moreover, the protective effect of MIF has been reported in amyotrophic lateral sclerosis where elevated MIF levels inhibited the accumulation of misfolded SOD1 [38]. In stroke, cerebral ischemia and depression, MIF has protective as well as pathological roles [36]. From the accumulating data, MIF possesses diverse functions within the CNS and more research is needed to decipher its specific role in normal and pathological conditions.

This review focuses on MIF research and actions in progressive brain diseases such as MS, AD and GBM. As a molecule broadly involved in many biological events and variety of autoimmune or inflammatory conditions, MIF can become a new potential biomarker and therapeutic target for the development of new prognostic, diagnostic as well as treatment strategies. Of note, different approaches to modulating MIF-dependent pathways are now in advanced clinical testing, including neurologic disorders.

Multiple sclerosis

MS is an autoimmune inflammatory disorder of the CNS characterized by demyelination and permanent neurological disability in young adults with prevalence in women [39, 40]. The most common form of MS at the beginning of disease is relapsing–remitting (RRMS) characterized by spontaneous episodes and partial recovery in disease severity with accumulating neurological dysfunctions over time [39]. After several years, disease mostly progresses into secondary progressive MS (SPMS) with slow permanent advancement of neurological, physical, and mental disfunction [41]. The pathogenesis of MS possesses a strong immune component and MIF as an inflammatory cytokine with potent control over innate and adaptive arms of immunity contributes to the development and progression of the disease [29]. Activation of MIF is essential for regulation of leukocyte migration across the blood–brain barrier [42]. Infiltration of immune cells to the brain tissue causes inflammation, demyelination, and formation of sclerotic plaques, hallmarks of MS.

MIF profiles in MS patients

MIF drives T cell and macrophage activation and may play a pivotal role in MS. Several different studies have been performed in order to decipher a role of MIF in MS pathogenesis. However, information about MIF expression in MS patients is limited with some contradictions due to variations in groups of MS patients with respect to different stages and severity of disease. In 2000, Niino et al. determined the level of MIF in the cerebrospinal fluid (CSF) of patients with conventional-form multiple sclerosis (C-MS), optic-spinal form multiple sclerosis (OpS-MS), and neuro-Behcet's disease (NBD) [43]. The highest levels of MIF have been found in the CSF of OpS-MS patients in relapse. Elevated levels of MIF were also found in relapsed but not in remission cases of C-MS. In NBD patients, the concentration of MIF in CSF was significantly elevated compared with control samples [43]. Similarly, increased levels of MIF were found in sera of untreated patients with MS relapse indicating their association with MS disease activity [44]. To that point, the high levels of MIF correlated with clinical MRI findings with a worsening EDSS score in different subtypes of MS including clinically isolated syndrome (CIS) [45]. A recent study in CIS patients revealed that observed overexpression of MIF, D-DT, and CD44 appeared to be unique for CD4( +)T cells [46]. In healthy blood MIF is predominantly expressed by B cells [47]. In early MS patients, B cells have been found to have downregulated MIF and MIF receptor (CD74) and upregulated the MIF receptor CXCR4 as compared to healthy controls, potentially reflecting a functional state of anergy that may contribute to the persistence of pathogenic immature B cells in the periphery [47]. In another study, MIF was shown to be highly expressed in human active white matter MS lesions predominantly associated with reactive hypertrophic GFAP + astrocytes and macrophages, suggesting MIF may contribute to the actively demyelinating lesion [48]. A more recent study showed increased levels of MIF both in CSF and in serum of RRMS patients [49]. In contrast, the study by Hjaeresen et al. shows that MIF is decreased during RRMS and increased in SPMS [50]. Additionally, MIF levels were significantly decreased in females with CIS and RRMS as compared to males suggesting sex-dependent regulation of MIF production. These findings are in accordance with our previous study and demonstrate the importance of estrogens and estrogen receptor in inhibition of MIF expression, as well as the binding between MIF and its CD74 receptor in the monocyte sub-population [51, 52]. The findings on how MIF exerts its effect on MS progression in males and females require further clarifications.

MIF genetic polymorphisms in MS

MIF genetic polymorphism studies support the role of MIF in the pathogenesis and severity of MS. Two polymorphisms in the promoter region of MIF gene have shown correlation with the worsening of autoimmune and inflammatory diseases such as systemic lupus erythematosus (SLE), rheumatoid arthritis (RA) and psoriatic arthritis (PsA) in Mexican-Mestizo population [53,54,55,56,57]. As studied in the white Turkish population, a MIF polymorphism has been associated with younger age of MS disease onset. Patients with multiple sclerosis had an increase in the MIF-173 CC genotype and exhibited significantly lower age of disease onset compared with those with the MIF-173 CG and MIF-173 GG genotypes. The MIF-794 CATT 6/7 genotype had a significantly lower progression index compared with MIF-794 CATT 6/6 and patients with the MIF-794 CATT 5/6 genotype had a significantly later age of disease onset [58]. In another study in the Turkish population, the results suggested no relation between MS susceptibility and MIF gene-173G>C polymorphism [59]. In Mexican patients, the MIF-173 GC genotype was associated with a higher clinical severity of MS [60]. Our study found a correlation between a high expression −794CATT5-8 and associated −173G/C SNP with increased MIF and D-DT levels in males with progressive disease [52]. These findings on the sex-specific contribution of MIF polymorphisms were supported by studies on MS patients in Western Mexico. When grouping by sex, an effect of both MIF polymorphisms (−794 CATT5-8 and − 173 G > C) was found with high MIF serum levels, increased severity and progression in male MS patients [61]. Both studies suggest that MIF polymorphisms could act as sex-specific disease modifiers that increase the severity and progression of MS in male patients. Further confirmation that 173G > C polymorphism can also regulate DDT expression in a sex-specific way and that the DDT is highly expressed in MS brain tissues and promotes MS progression in males but not females has been reported recently [62].

MIF in the EAE animal model

In experimental autoimmune encephalitis (EAE), an animal model for MS, MIF has been shown to accelerate disease progression, mostly by activation of macrophages and microglia and upregulation of the inflammatory responses in the CNS [48, 63,64,65,66,67]. Moreover, MIF promotes resistance of pathogenic CD4( +) T cells to glucocorticoid treatment in EAE [68]. In SJL mice with severe EAE, treatment with anti-MIF antibodies reduced disease severity and improved recovery by impairment of CNS homing of pathogenic T cells [69]. In another study, MIF-deficient C57BL/6 mice were protected from EAE and treatment with a small-molecule inhibitor of MIF reduced EAE severity in SJL mice [70]. Using MIF-/- mice, it was reported that MIF is necessary for progression of EAE, possibly due to significant decreases in inflammatory cytokines [64]. In our previous study, we demonstrated that MIF or D-DT deficiency ameliorates EAE severity and that D-DT absence is associated with reduced migration of memory and activated mononuclear cells into the CNS. We also showed that genetically controlled high expression of both molecules promotes MS progression in males and that both molecules are important sex-specific disease modifiers [52]. A novel role for MIF in inducing microglial C/EBP-beta, a transcription factor shown to regulate myeloid cell function has also been proposed in a rodent model of MS [48].

MIF-oriented MS treatments

At present there are no sufficient and satisfactory treatments for MS. The most important caveat in systematically administrated drugs is the limited penetration through BBB. Drugs such as monomethyl fumarate (MMF), a product of dimethyl fumarate (DMF) hydrolysis after absorption inside the small intestine and MTX (mitoxantrone) have only limited access to the CNS. Thus, these drugs would likely have little influence on MIF levels in CNS-resident cells and limited effect on increased MIF levels in CSF as found in RRMS patients [50]. A newer anti-inflammatory drug, ibudilast, a non-selective inhibitor of various cyclic nucleotide phosphodiesterases often used as a bronchodilator for bronchial asthma treatment, plays a central role in processes like inflammation and synaptic plasticity. Ibudilast suppresses pro-inflammatory cytokines, upregulates anti-inflammatory cytokines and blocks TLR4 and acts as a noncompetitive and allosteric inhibitor of MIF tautomerase activity and its chemotactic effects [71]. Additionally, ibudilast possesses an enhanced ability to pass the BBB, and was found in a successful PMS Phase 2 clinical trial to inhibit glial activity, support the production of neurotrophic factors and influence CNS production of MIF [72]. Other therapeutic approaches such as a small molecule inhibitor (ISO-1) and MHC constructs (DRQ) will be discussed below. That said, we are not aware of any studies using MS approved drugs that have evaluated MIF levels.

Alzheimer disease

Alzheimer disease (AD) is the most common neurodegenerative disease affecting predominantly the hippocampus and cerebral cortex characterized by the aggregation of extracellular Aβ proteins and intracellular neurofibrillary tangles (NFT), which are composed of hyperphosphorylated tau proteins in neurons. Neuroinflammation plays a pivotal role in AD pathogenesis leading to neuronal loss, alterations in glial cells and severe cognitive decline.

MIF’s effects on glia and neurons in AD

One of the first reports on the MIF’s involvement in AD identified MIF as a new Aβ-binding protein in a soluble fraction of the cerebral cortex of AD brain by immunoprecipitation [73, 74]. Some early reports using immunohistochemistry reported elevated expression of CD74, a MIF receptor in AD [75, 76]. CD74 was found to be increased in microglia in AD cases compared to age-matched controls [76]. Following study revealed a significant increase in CD74 primarily in neurofibrillary tangles, amyloid-beta plaques, microglia and for the first time in neurons of AD cases [75]. Toxic involvement of MIF within amyloid-aggregates was established by studies in brains of transgenic APP mice where MIF has been found to be produced by activated microglia near Aβ plaques [77]. The co-localization of MIF and activated microglia to amyloid deposits has been further confirmed by using mass spectrometry-based imaging technique [78]. Besides microglia, MIF possesses strong influence also on astrocyte activation (Fig. 3). MIF in astrocytes plays an important role in elevated tau phosphorylation, which involves mediators released by the activated astrocytes in AD animal model [79]. In animal models, the expression of MIF in astrocytes and the number of reactive astrocytes were noticeably increased in contrast to MIF knockout mice. Additionally, it has been found that conditioned medium from activated astrocytes could stimulate tau hyperphosphorylation in neurons in a MIF-dependent manner. Recently it has been shown that MIF displays neurotoxicity similar to Aβ [1,2,3,4,5,6,7,8,9,10,11,12,13,14,15,16,17,18,19,20,21,22,23,24,25,26,27,28,29,30,31,32,33,34,35,36,37,38,39,40,41,42], which was associated with the MIF-induced increase in apoptosis in human neuroblastoma cells [80]. It has been found that MIF’s involvement in neuronal cytotoxicity can be reversed by ISO-1 that blocks the enzymatic and biologic activities of MIF [81]. In vitro data using murine and human neuronal cell lines revealed that ISO-1 almost completely antagonized the neurotoxic effects of Aβ [77]. A recent study on the effect of MIF on inflammatory markers and spatial learning in a mouse model of sporadic AD and on tau pathology in AD patients showed that MIF inhibition resulted in reduced cytokine production in vitro and in vivo [82].

Sugar metabolic malfunction affects MIFs role in high glucose-induced AD

Some recent studies implicate MIF with progression of high glucose-induced AD. AGEs (advanced glycation endproducts) are neurotoxic, foster the deposition of Aβ and the hyperphosphorylation of tau protein and the expression of proinflammatory mediators in glial cells [83, 84]. It has been demonstrated that AGEs promoted the expression of MIF and aggravated the neuroinflammatory response at the cell level [85]. In PC12 cells, (an AD-cell model), ISO‑1 reduced AGE‑mediated damage by decreasing the expression of neuroinflammatory mediators. Previously, MIF has been found to be glycated and oxidized in AD brain homogenates. The glycation completely inhibited the enzymatic activity of MIF and was harmful to the signaling effects of MIF on glia, strongly weakening MIF-induced ERK phosphorylation [86] (Fig. 2). This might be especially important at the beginning of AD where microglia are actively involved in removing Aβ plaques and MIF signaling is crucial for this beneficial microglia’ function. Thus, dysregulation of glucose homeostasis or insulin regulation leads to MIF conformational changes and severely affects MIF activity with implications for impaired innate immune response during progression of AD [86].

Fig. 2
figure 2

MIF in AD. Activated glia and especially neurons are the main sources of MIF in AD. Dysregulated glucose metabolism and AGE promote MIF conformational changes leading to disrupted ERK signaling in glia cells. In AD increased CD74 has been found in microglia, neurofibrillary tangles, amyloid β plaques and neurons

MIF in AD patients

In human subjects with AD at early clinical stages, cerebrospinal fluid levels of MIF were increased in comparison with age-matched controls, and correlated with biomarkers of tau hyper-phosphorylation and neuronal injury suggesting that MIF can play a role as biomarker for early-stage AD. First report from clinical studies showed significantly increased levels of MIF in the CSF of AD patients in comparison to age matched controls [87]. A possible link between MIF and TNF-α release in AD group is suggested as a correlation between MIF and TNF-a concentrations has been found. The next study by this group showed the highest levels of MIF in the brain cytosol and CSF in a mild cognitive impairment group of patients (MCI) that has a high risk to develop AD over time, thus providing evidence that the neuroinflammation occurs early, at predementia stages of AD [77]. A recent study based on measurements of MIF levels in plasma and CSF in MCI or mild dementia (cognitive impairment, CI) patients established a significant role for MIF as biomarker in AD pathology for predicting cognitive failure in MCI and CI [88]. Additionally, this study provided evidence that MIF-related inflammation is related to amyloid pathology, tau hyperphosphorylation, and neuronal injury at the early clinical stages of AD. Further usefulness of MIF as a potential AD biomarker has been proposed by Zhang et al. [89]. In this study, elevated MIF levels were detected in CSF of AD patients but not in MCI or vascular dementia patients. Neurons but not glia cells stimulated with Aβ oligomers were the main source of MIF. Interestingly, reduced MIF expression impaired learning and memory in the AD mouse model thus supporting the conclusion that neuronal secretion of MIF may serve as a defense mechanism to compensate for declining cognitive function in AD. MIF has been found to have neuroprotective abilities on neuronal cells by inducing expression of BDNF, an essential modulator of synaptic plasticity related to learning and memory [90]. MIF administration protected neurons from hypoxic injury by upregulation of mature BDNF and anti-apoptotic molecules in human neuroblastoma cells. Previously, BDNF, serotonin and THP2, a critical enzyme in the biosynthesis of serotonin in the brain have been found to be upregulated by MIF in vitro as well as during both exercise and electroconvulsive seizure in vivo [91].

Glioblastoma

Glioblastoma (GBM) is a grade IV astrocytoma derived from astrocytes according to WHO classification [92]. GBM is the most common and the most deadly brain tumor with low treatment efficacy after surgery, chemotherapy and radiation. One of the main reasons for poor therapeutic outcome in this type of cancer is marked cellular heterogeneity with genetic and epigenetic variability [93]. Recent genome-wide association studies (GWAS) showed that genetic susceptibility to GBM and non-GBM tumors are highly distinct with possible different etiologies [94].

Pro-carcinogenic abilities of MIF

Despite MIF’s critical role in the pathogenesis of inflammatory and immune responses, it has been documented that MIF promotes carcinogenesis and its elevated levels have been found in various tumors [95,96,97]. MIF is able to promote cellular processes related to tumorigenesis such as cell cycle progression, tumor growth, blockage of apoptosis, induction of angiogenesis and tumor spread [98] (Fig. 3). Moreover, MIF has profound and detrimental effects on anti-cancer immune responses. This includes inhibition of NK anti-cancer cytotoxic effect, inhibition of T cell activation, promotion of MDSCs (myeloid-derived suppressor cells) and polarization of macrophages towards an anti-inflammatory phenotype [99,100,101,102,103]. Blockade of MIF by shRNA in glioma cells restores cytotoxic activity of NK and CD8 + T cells downregulating the immune receptor NKG2D [104]. In contrast to abundant studies showing that MIF is a key factor in tumor immune response, recently it has been found that its cognate receptor CD74 is confined to human microglia/macrophages and is positively associated with pro-inflammatory anti-tumor immune responses and improved patients’ survival [105]. Considering the extremely high diversity of microglia subpopulations with unique gene expression profiles and different roles, more studies are needed to decipher the role of CD74 in microglia anti-tumor responses.

Fig. 3
figure 3

Main MIF sources and its effects in GBM. One of the key stressors in development of GBM are hypoxia and hypoglycemic states which induce production of high MIF levels in primary GBM cells leading to neovascularization. MIF major sources are: 1. cancer stem cells (CSC) located nearby new blood vessels; 2. transformed astrocytes and 3. MCs. All sources add to promotion of immune suppression mostly by MDSCs, angiogenesis and increased cell proliferation

High levels of MIF predict poor survival rates in cancer patients [106, 107]. In prostate cancer and in many solid tumors, increased risk is associated with polymorphism in the C allele in the MIF-173 G/C promoter single-nucleotide polymorphism (SNP) as documented by meta-analysis. MIF gene promoter polymorphisms also are associated with lymphatic metastasis and cervical cancer [108, 109]. MIF is highly expressed by GBM cells and higher levels of MIF are associated with increased cancer grade [104, 110] as well as poor prognosis and tumor recurrence [103, 111]. One of the critical stressors in the development of GBM are hypoxia and hypoglycemic states which induce production of high MIF levels in primary GBM cells leading to neovascularization [110, 112] (Fig. 3). Hypoxia-induced MIF expression is driven by HIF-1, but amplified by hypoxia-induced degradation of cAMP-responsive element binding protein (CREB) [113]. In the hypoxic area of glioma specimens, MIF co-localized with CXCR4 where MIF promotes vasculogenic mimicry formation [114].

MIF effects on the cellular environment in glioma

In the CNS, MIF is produced by neurons, astrocytes, oligodendrocytes, microglia and Schwann cells [10, 115]. Activated microglia and macrophages play a central role in the delivery of MIF and signaling molecules to nearby neurons and transformed astrocytes [116, 117]. An increased number of activated microglia that reside in the tumor periphery and infiltrating macrophages that occupy the perivascular area account up to 30–50% of all cell types in gliomas and correlate with higher grade and worse prognosis [118]. The number of glioma-associated microglia/macrophages (GAMs) and MDSCs is the highest in malignant gliomas and inversely correlates with patient survival [103, 117]. MDSCs have been identified nearby cancer stem cells (CSC) which are responsible for high MIF production and promotion of immune suppression activities of MDSCs [119]. Depletion of MDSCs by 5-fluorouracil (5-FU) increases the survival rate in mice with glioma [119]. In vitro treatment with sulforaphane, a MIF inhibitor, reverses the transition of normal monocytes to MDSCs [120]. Higher numbers of the circulating monocytic subset of MDSCs (M-MDSC), in contrast to macrophages and regulatory T cells, have been associated with higher tumor grade and poorer prognosis in GBM [103, 121]. Recently it has been found that in both murine and human models, M-MDSCs express high levels of the CD74 MIF receptor and are localized in the glioma microenvironment, in contrast to granulocyte-MDSCs (G-MDSCs) which show minimal accumulation in the tumor environment [103]. Ibudilast which appears effective in an experimental model of glioma, suppresses MDSCs and boosts activity of CD8 T cells [103]. Additionally, mast cells (MCs) have a crucial role in glioma pathogenesis that contribute to angiogenic processes and immune regulation [122]. MIF attracts mast cells to the tumor environment and accumulation of MCs and their pSTAT5 expression correlates with the level of MIF [123]. MIF expression was frequently associated with the presence of the tumor-suppressor gene p53, which supports stem cell tumorigenic activities [110]. Many stem cells, including human pluripotent stem cells (iPCS), were the source of MIF [124]. Blockade of MIF by miR-608 in glioma stem cells reduced the proliferation, translocation, and invasion [125]. MIF gene knockdown effected brain tumor initiating cell (BTIC) function through direct, intracellular inhibition of p53, reducing cell proliferation and increasing apoptosis in an in vitro setting and in human mouse xenograft model [126].

MIF and D-DT signaling in cancer

MIF signals though several different receptors primarily via cognate CD74 and non-cognate receptors CXCR2, CXCR4 and CXCR7 [4, 25, 127, 128]. Upon binding to CD74 in combination with the CD44 co-receptor, the signal is transduced via src/MAPK signaling pathway, whereas binding to CXCR2, MIF signaling is PI3K/Akt-dependent [128]. MIF activates MAPK and PI3K pathways involved in signal transduction cascades in many cancers [103, 121, 129,130,131]. Additionally, the PI3K pathway contributes to the MIF-mediated suppression of apoptosis linked with p53 activation [132]. It has been found that in melanoma, CD74-MIF interaction promotes tumor survival via the PI3K/AKT pathway in response to IFN-γ [133]. A second ligand for CD74 is D-DT (a.k.a. MIF-2) with close structural and functional similarities with MIF and an overlapping spectrum of activities such as tumor formation, growth and survival of cancer cells and tumor invasion [134, 135]. Both molecules cooperate in tumorigenesis. For example, in non-small cell lung carcinoma both molecules via CD74 negatively regulated AMP-activated protein kinase (AMPK) leading to increased expression of CXCL8 and VEGF [136, 137]. In pancreatic cancer tissue, D-DT was over-expressed together with MIF and knockdown of D-DT and MIF in a pancreatic cell line, PANC-1, cooperatively inhibited ERK1/2 and AKT phosphorylation, increased p53 expression, and reduced cell proliferation, invasion and tumor formation. A covalent tautomerase inhibitor of both DDT and MIF, a 4-iodo-6-phenylpyrimidine (4-IPP), attenuated cell proliferation and colony formation in vitro and tumor growth in vivo [138]. 4-IPP irreversibly binds to Pro1 of MIF or D-DT via nucleophilic dislocation by Pro1 of an aromatic iodo group and was shown in vitro to have anticancer properties in the head and neck squamous cell carcinoma cell line, SCCVII, and in the human A549 lung adenocarcinoma [139]. In GBM, targeted inhibition of MIF and D-DT by 4-IPP might improve radiation therapy [140]. Additive effects have been shown in dual inhibition of MIF and D-DT by shRNA treatment in clear cell renal cell carcinomas [141]. Current treatment strategies are focused on simultaneous counteracting of both cytokines.

Anti-MIF treatment strategies in glioma models

Several different anti-cancer and GBM treatment approaches based on MIF inhibition have been proposed and include competitive, irreversible and endogenous inhibitors, molecules that destabilize MIF, and monoclonal antibodies blocking MIF or CD74 [34, 103, 142]. One of the biggest caveats in glioma treatment strategies is the inability of drugs to traverse the blood–brain barrier (BBB). Attempts have been made to design more lipophilic compounds with better ability to reach CNS tumors. Recently this direction of research has become focused on nanotechnology [143, 144]. So far, only liposomes have reached phase I/II clinical trials [143]. One of the well-known MIF competitive inhibitors is ISO-1. ISO-1 reduces the proliferation of human glioblastoma cell lines, especially the human LN18 cell line, in a dose-dependent manner and was able to restore contact inhibition, reduce proliferation and mitogenic signaling [112, 145]. Moreover, ISO-1 was able to sensitize glioma cells to glucocorticoids, and when applied together with dexamethasone, cell migration and invasion were diminished in Hs683 glioma cells [146]. MIF knockdown by antisense transfection allowed for restoration of contact inhibition in human glioblastoma cell lines [145]. Blockade of MIF with shRNA resulted in an increase of CD8-positive CTLs and reduction of Treg lymphocytes in the brain in animal models of glioma [119]. Silencing of CD74 by shRNA was associated with reduced AKT and ERK1/2 pathways and in the human glioma U87 cell line, significantly suppressed proliferation and increased temozolomide sensitivity [147]. Monoclonal antibodies against MIF have been tested in in vitro settings where they were able to reduce growth of glioma cell lines, the migration of cells and arginase-1 assembly in MDSCs in a CXCR2-dependent manner [119, 123, 145]. Treatment with 4-IPP (inhibitor of MIF or D-DT) showed the potential to improve radiotherapy by inhibiting the stemness and intracellular signaling pathways and inducing apoptosis in vitro and in vivo glioma models [140]. A common chemotherapy in glioma can be efficiently enhanced by using combined treatments. Synergism in the inhibition of cell cycle and increased apoptosis has been observed in ex vivo and in vivo models when ibudilast was combined with temozolomide leading to significant increased overall survival [148]. Despite some successful results in the in vitro experiments mentioned, more studies evaluating molecules and their receptors with known genetic polymorphisms are needed to help establish the clinical relevance of potential therapeutics in GBM.

Discussion of MIF and CD74 inhibitors for possible clinical use in MS, AD and astrocytomas

MIF-inhibitors as novel neurotherapeutics

Once CD74 was determined to be the cognate receptor for MIF [25], additional CD74 targeted pharmacologic approaches became available for consideration in CNS disease. A monoclonal antibody targeting CD74 (milatuzumab) has received US FDA approval for multiple myeloma [149], a malignancy of B cells where MIF–CD74 survival pathways are highly active [150]. Milatuzumab further showed evidence of efficacy in lupus nephritis, a highly inflammatory condition, in one phase 1B study [151]. Anti-MIF (imalumab) completed phase II testing in heavily pre-treated cancer patients with a favorable safety profile [152], but has yet to be evaluated further. The rational design of small-molecule antagonists of MIF/CD74 interaction has been facilitated by MIF’s vestigial tautomerase active site, which overlaps structurally with the CD74 binding domain [153, 154]. This feature has been exploited by several academic and industry groups to design small-molecule MIF tautomerase inhibitors that target this site [155,156,157,158] and a subset of such inhibitors shows therapeutic activity in mouse models of autoimmunity [159, 160]. The small molecule MIF antagonist that is furthest advanced in clinical development is ibudilast, which was originally developed as a phosphodiesterase inhibitor but was discovered to inhibit MIF allosterically [71]. Remarkably, ibudilast binds to a dynamic site that is not present in the (apo) crystal form of MIF; that is, this site is only revealed when ibudilast binds to MIF. Once bound, the ensuing conformational changes eliminate MIF activity. Ibudilast has shown efficacy in a phase II study of MS, where high-expression MIF genotype is a risk for progressive disease [52, 161]. Ibudilast is used for asthma in Japan and is in clinical testing in the US for additional inflammatory conditions, as well as in oncology and in neurodegenerative disease conditions.

The recent appreciation that the MIF family member D-DT is co-expressed with MIF in many disease conditions, and also activates CD74 has prompted interest in dual MIF/D-DT inhibitors [162, 163], and possibilities exist for targeting both MIF and D-DT together by either small molecule or bispecific antibody approaches [164, 165].

Major histocompatibility complex (pMHC) constructs represent a novel therapeutic approach for treatment of PMS and other conditions involving activation of the CD74 pathway

RTL1000: In the mid-1990s, our lab designed recombinant T cell receptor ligands (RTLs) that mimic the major histocompatibility complex (MHC) class II/peptide interface with cognate T cell receptors [166,167,168]. The initial RTL1000 construct, designed to partially mimic only the extracellular components of MOG-35-55 specific T cell receptor (TCR) ligands present on antigen presenting cells, consisted of the alpha-1 and beta-1 domains of HLA-DR2 linked to the MOG-35-55 peptide. Interaction of the soluble RTL1000 with the TCR of MOG-35-55 specific CD4 + T cells produced partial, but incomplete, T cell activation that selectively blocked T cell proliferation and IL-2 secretion of MOG-35-55 specific T cells. In a series of studies, we demonstrated that partial major histocompatibility complex (pMHC) constructs bind to human and mouse monocytes through cell surface CD74 and that this interaction inhibits MIF binding and signaling [65, 169]. The RTL1000 construct could strongly inhibit EAE, a mouse model of MS in HLA-DR2*1501 transgenic mice [170]. This model was relevant to ~ 50% of MS patients that were DR2 positive. RTL1000 was approved by FDA for a Phase 1 clinical trial carried out in 2007–2009 which demonstrated safety and tolerability of RTL1000 at a ≤ 60 mg dose given i.v. [171, 172].

DRα1-MOG-35–55: We later learned that another major receptor for RTL1000 (besides the TCR) was the highly conserved CD74 molecule [169], which also was the receptor for MIF and D-DT. By slightly altering the RTL1000 design (removal of the DR2 beta 1 domain that contained histocompatibility markers), we produced a simplified ligand, DRα1-MOG-35-55, for CD74 that could competitively inhibit MIF/D-DT signaling, block downstream inflammatory activity (ERK-1/2 phosphorylation and MAPK activation) and treat EAE in both DR2*1501 transgenic and wild-type C57BL/6 mice comparably with RTL1000 that was active only in the DR2 transgenic mice [66, 166]. We found that the DRα1-MOG-35-55 construct displayed a stronger effect in downregulating CD74 expression on male CD11b + cells as compared to female cells and the treatment of chronic EAE in female mice was dependent on signaling through ER-α [51], a restriction that we could overcome with higher doses of DRα1-MOG-35-55. Due to conservation of DRα1 in humans and mice and molecular modeling implicating DRα1 but not DRβ1 amino acid residues as the major binding region for CD74, the DRα1-MOG-35-55 construct was confirmed to retain ability to block CD74 signaling and reverse clinical signs of EAE independent of MHC barriers [173]. The DRα1 construct can inhibit the activation and recruitment of brain-infiltrating T cells and CD11b + CD45hi myeloid cells and expression of the co-stimulatory CD86 marker on CD11b + CD45hi cells, and at the same time increase expression of the anti-inflammatory CD206 marker on CD11b + CD45int microglial cells [173]. Moreover, DRα1-MOG35-55 treatment could significantly reduce severe disease enhancing effects of MIF and D-DT in both male and female wild-type and knockout mice with chronic EAE [174, 175].

DRQ: DRQ is a third generation 13.5 kDa protein construct comprising the human (h)MOG-35-55 peptide covalently linked to the L50Q modified DR alpha-1 domain of MHC class II [176]. This modified construct had eightfold higher inhibitory activity for blocking MIF/D-DT binding and downstream signaling through CD74 and enhanced ability to reverse EAE induced paralysis in wild-type mice. DRQ is thus our prime candidate for treatment of progressive MS as well other MIF/CD74-dependent inflammatory conditions, including stroke, methamphetamine abuse, traumatic brain injury, Alzheimer’s disease and GBM without the need for histocompatibility testing of DRQ recipients [173]. Inhibition of MIF and/or D-DT signaling by DRQ, ISO-1 or ibudilast may have the potential to slow MS progression (Fig. 4).

Fig. 4
figure 4

MIF and D-DT inhibition in MS. MIF and D-DT signaling through CD74 is involved in MS progression by increasing inflammatory cell migration to the CNS, enhancing secretion of pro-inflammatory cytokines and prolonging survival of pro-inflammatory cells. Inhibition of MIF or D-DT signaling by partial MHC constructs (DRQ), ISO-1 or ibudilast attenuates signs of MS progression

Concluding remarks

MIF is a pleiotropic protein that functions as a glucocorticoid-induced immunoregulator, pituitary hormone, inflammatory cytokine, and immune and growth response regulator, with pathologic roles in autoimmunity, neurologic disorders, and oncology. There is a growing interest in the therapeutic application of MIF antagonists in different diseases. Increased studies of D-DT, a MIF homolog molecule with overlapping biological functions and possible synergism as MIF, has focused attention on the utility of dual inhibition for the full potential of therapy. Possible pharmacological strategies include small molecules that disrupt both MIF and D-DT interaction with CD74, either as competitive or irreversible inhibitors. In this regard, 4-IPP, is one dually active, irreversible inhibitor. Anti-CD74 similarly prevents both MIF and D-DT signaling, and the bio-engineering of dual specificity antibodies (e.g., binding to both MIF and D-Dt) can be envisioned. A novel therapeutic approach utilizing our DRQ construct to inhibit dual effects of MIF and D-DT signaling has the potential to treat all progressive brain diseases involving CD74-dependent neuroinflammatory pathways. In Table 1, we indicate the operative mechanisms of action of MIF and D-DT in MS, Alzheimer disease and glioblastoma and potential therapeutic anti-MIF/D-DT drugs that could block their pathogenic effects. Additionally, the development of new technologies that identify genetic heterogeneity of cellular subpopulations responsible for pathology such as single cell analysis, cellular bar coding, CRISPR-Cas 9 and CyTOF hold future promise for new therapeutics. Finally, the circumstance that approximately 20% of individuals express a high expression MIF allele [52] supports the possibility that MIF/CD74 directed therapies would be most effectively used in such subjects, thus providing a more precise pharmacogenomic for treatment of a number of MIF-dependent illnesses. Further studies are needed to decipher how MIF inhibitors block the hyperactivation of cells, including glia cells in the CNS, and exert anti‑inflammatory and neuroprotective effects.

Table 1 The operative mechanisms of action of MIF and D-DT in MS, Alzheimer disease and glioblastoma and potential therapeutic anti-MIF/D-DT drugs that could block their pathogenic effects

Availability of data and materials

Not applicable.

References

  1. Bloom BR, Bennett B. Mechanism of a reaction in vitro associated with delayed-type hypersensitivity. Science. 1966;153(3731):80–2.

    Article  CAS  PubMed  Google Scholar 

  2. 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(1):72–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Weiser WY, Temple PA, Witek-Giannotti JS, Remold HG, Clark SC, David JR. Molecular cloning of a cDNA encoding a human macrophage migration inhibitory factor. Proc Natl Acad Sci U S A. 1989;86(19):7522–6.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Bernhagen J, Krohn R, Lue H, Gregory JL, Zernecke A, Koenen RR, et al. MIF is a noncognate ligand of CXC chemokine receptors in inflammatory and atherogenic cell recruitment. Nat Med. 2007;13(5):587–96.

    Article  CAS  PubMed  Google Scholar 

  5. Uhlen M, Fagerberg L, Hallstrom BM, Lindskog C, Oksvold P, Mardinoglu A, et al. Proteomics. Tissue-based map of the human proteome. Science. 2015;347(6220):1260419.

    Article  PubMed  Google Scholar 

  6. Thul PJ, Åkesson L, Wiking M, Mahdessian D, Geladaki A, Ait Blal H, Alm T, Asplund A, Björk L, Breckels LM, Bäckström A, Danielsson F, Fagerberg L, Fall J, Gatto L, Gnann C, Hober S, Hjelmare M, Johansson F, Lee S, Lindskog C, Mulder J, Mulvey CM, Nilsson P, Oksvold P, Rockberg J, Schutten R, Schwenk JM, Sivertsson Å, Sjöstedt E, Skogs M, Stadler C, Sullivan DP, Tegel H, Winsnes C, Zhang C, Zwahlen M, Mardinoglu A, Pontén F, von Feilitzen K, Lilley KS, Uhlén M, Lundberg E. A subcellular map of the human proteome. Science. 2017;356(6340):eaal3321. https://doi.org/10.1126/science.aal3321.

  7. Uhlen M, Zhang C, Lee S, Sjöstedt E, Fagerberg L, Bidkhori G, Benfeitas R, Arif M, Liu Z, Edfors F, Sanli K, von Feilitzen K, Oksvold P, Lundberg E, Hober S, Nilsson P, Mattsson J, Schwenk JM, Brunnström H, Glimelius B, Sjöblom T, Edqvist PH, Djureinovic D, Micke P, Lindskog C, Mardinoglu A, Ponten F. A pathology atlas of the human cancer transcriptome. Science. 2017;357(6352):eaan2507. https://doi.org/10.1126/science.aan2507.

  8. Calandra T, Bucala R. Macrophage migration inhibitory factor (MIF): a glucocorticoid counter-regulator within the immune system. Crit Rev Immunol. 2017;37(2–6):359–70.

    Article  PubMed  Google Scholar 

  9. Michell-Robinson MA, Touil H, Healy LM, Owen DR, Durafourt BA, Bar-Or A, et al. Roles of microglia in brain development, tissue maintenance and repair. Brain. 2015;138(Pt 5):1138–59.

    Article  PubMed  PubMed Central  Google Scholar 

  10. Su Y, Wang Y, Zhou Y, Zhu Z, Zhang Q, Zhang X, et al. Macrophage migration inhibitory factor activates inflammatory responses of astrocytes through interaction with CD74 receptor. Oncotarget. 2017;8(2):2719–30.

    Article  PubMed  Google Scholar 

  11. Sumaiya K, Langford D, Natarajaseenivasan K, Shanmughapriya S. Macrophage migration inhibitory factor (MIF): a multifaceted cytokine regulated by genetic and physiological strategies. Pharmacol Ther. 2022;233: 108024.

    Article  CAS  PubMed  Google Scholar 

  12. Song S, Xiao Z, Dekker FJ, Poelarends GJ, Melgert BN. Macrophage migration inhibitory factor family proteins are multitasking cytokines in tissue injury. Cell Mol Life Sci. 2022;79(2):105.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Calandra T, Bernhagen J, Metz CN, Spiegel LA, Bacher M, Donnelly T, et al. MIF as a glucocorticoid-induced modulator of cytokine production. Nature. 1995;377(6544):68–71.

    Article  CAS  PubMed  Google Scholar 

  14. Bacher M, Metz CN, Calandra T, Mayer K, Chesney J, Lohoff M, et al. An essential regulatory role for macrophage migration inhibitory factor in T-cell activation. Proc Natl Acad Sci U S A. 1996;93(15):7849–54.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Yao J, Leng L, Fu W, Li J, Bronner C, Bucala R. ICBP90 regulates MIF expression, glucocorticoid sensitivity, and apoptosis at the MIF immune susceptibility locus. Arthritis Rheumatol. 2021;73(10):1931–42.

    Article  CAS  PubMed  Google Scholar 

  16. Bernhagen J, Calandra T, Cerami A, Bucala R. Macrophage migration inhibitory factor is a neuroendocrine mediator of endotoxaemia. Trends Microbiol. 1994;2(6):198–201.

    Article  CAS  PubMed  Google Scholar 

  17. Bernhagen J, Calandra T, Mitchell RA, Martin SB, Tracey KJ, Voelter W, et al. MIF is a pituitary-derived cytokine that potentiates lethal endotoxaemia. Nature. 1993;365(6448):756–9.

    Article  CAS  PubMed  Google Scholar 

  18. Lerch JK, Puga DA, Bloom O, Popovich PG. Glucocorticoids and macrophage migration inhibitory factor (MIF) are neuroendocrine modulators of inflammation and neuropathic pain after spinal cord injury. Semin Immunol. 2014;26(5):409–14.

    Article  CAS  PubMed  Google Scholar 

  19. Mitchell RA, Metz CN, Peng T, Bucala R. Sustained mitogen-activated protein kinase (MAPK) and cytoplasmic phospholipase A2 activation by macrophage migration inhibitory factor (MIF). Regulatory role in cell proliferation and glucocorticoid action. J Biol Chem. 1999;274(25):18100–6.

    Article  CAS  PubMed  Google Scholar 

  20. Kleemann R, Hausser A, Geiger G, Mischke R, Burger-Kentischer A, Flieger O, et al. Intracellular action of the cytokine MIF to modulate AP-1 activity and the cell cycle through Jab1. Nature. 2000;408(6809):211–6.

    Article  CAS  PubMed  Google Scholar 

  21. Hudson JD, Shoaibi MA, Maestro R, Carnero A, Hannon GJ, Beach DH. A proinflammatory cytokine inhibits p53 tumor suppressor activity. J Exp Med. 1999;190(10):1375–82.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Su H, Na N, Zhang X, Zhao Y. The biological function and significance of CD74 in immune diseases. Inflamm Res. 2017;66(3):209–16.

    Article  CAS  PubMed  Google Scholar 

  23. Sanchez-Nino MD, Sanz AB, Ruiz-Andres O, Poveda J, Izquierdo MC, Selgas R, et al. MIF, CD74 and other partners in kidney disease: tales of a promiscuous couple. Cytokine Growth Factor Rev. 2013;24(1):23–40.

    Article  CAS  PubMed  Google Scholar 

  24. Kleemann R, Grell M, Mischke R, Zimmermann G, Bernhagen J. Receptor binding and cellular uptake studies of macrophage migration inhibitory factor (MIF): use of biologically active labeled MIF derivatives. J Interferon Cytokine Res. 2002;22(3):351–63.

    Article  CAS  PubMed  Google Scholar 

  25. Leng L, Metz CN, Fang Y, Xu J, Donnelly S, Baugh J, et al. MIF signal transduction initiated by binding to CD74. J Exp Med. 2003;197(11):1467–76.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Bucala R, Shachar I. The integral role of CD74 in antigen presentation, MIF signal transduction, and B cell survival and homeostasis. Mini Rev Med Chem. 2014;14(14):1132–8.

    Article  CAS  PubMed  Google Scholar 

  27. Gil-Yarom N, Radomir L, Sever L, Kramer MP, Lewinsky H, Bornstein C, et al. CD74 is a novel transcription regulator. Proc Natl Acad Sci U S A. 2017;114(3):562–7.

    Article  CAS  PubMed  Google Scholar 

  28. Kim BS, Tilstam PV, Hwang SS, Simons D, Schulte W, Leng L, et al. D-dopachrome tautomerase in adipose tissue inflammation and wound repair. J Cell Mol Med. 2017;21(1):35–45.

    Article  CAS  PubMed  Google Scholar 

  29. Merk M, Mitchell RA, Endres S, Bucala R. D-dopachrome tautomerase (D-DT or MIF-2): doubling the MIF cytokine family. Cytokine. 2012;59(1):10–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Yao J, Leng L, Sauler M, Fu W, Zheng J, Zhang Y, et al. Transcription factor ICBP90 regulates the MIF promoter and immune susceptibility locus. J Clin Invest. 2016;126(2):732–44.

    Article  PubMed  PubMed Central  Google Scholar 

  31. Radstake TR, Sweep FC, Welsing P, Franke B, Vermeulen SH, Geurts-Moespot A, et al. Correlation of rheumatoid arthritis severity with the genetic functional variants and circulating levels of macrophage migration inhibitory factor. Arthritis Rheum. 2005;52(10):3020–9.

    Article  CAS  PubMed  Google Scholar 

  32. Mizue Y, Ghani S, Leng L, McDonald C, Kong P, Baugh J, et al. Role for macrophage migration inhibitory factor in asthma. Proc Natl Acad Sci U S A. 2005;102(40):14410–5.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Mitchell RA, Bucala R. Tumor growth-promoting properties of macrophage migration inhibitory factor (MIF). Semin Cancer Biol. 2000;10(5):359–66.

    Article  CAS  PubMed  Google Scholar 

  34. 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(3):440–60.

    Article  CAS  PubMed  Google Scholar 

  35. Bucala R. MIF, MIF alleles, and prospects for therapeutic intervention in autoimmunity. J Clin Immunol. 2013;33(Suppl 1):S72–8.

    Article  PubMed  Google Scholar 

  36. Leyton-Jaimes MF, Kahn J, Israelson A. Macrophage migration inhibitory factor: a multifaceted cytokine implicated in multiple neurological diseases. Exp Neurol. 2018;301(Pt B):83–91.

    Article  CAS  PubMed  Google Scholar 

  37. Li S, Nie K, Zhang Q, Guo M, Qiu Y, Li Y, et al. Macrophage migration inhibitory factor mediates neuroprotective effects by regulating inflammation, apoptosis and autophagy in Parkinson’s disease. Neuroscience. 2019;416:50–62.

    Article  CAS  PubMed  Google Scholar 

  38. Israelson A, Ditsworth D, Sun S, Song S, Liang J, Hruska-Plochan M, et al. Macrophage migration inhibitory factor as a chaperone inhibiting accumulation of misfolded SOD1. Neuron. 2015;86(1):218–32.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Nylander A, Hafler DA. Multiple sclerosis. J Clin Invest. 2012;122(4):1180–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Dendrou CA, Fugger L, Friese MA. Immunopathology of multiple sclerosis. Nat Rev Immunol. 2015;15(9):545–58.

    Article  CAS  PubMed  Google Scholar 

  41. Barzegar M, Najdaghi S, Afshari-Safavi A, Nehzat N, Mirmosayyeb O, Shaygannejad V. Early predictors of conversion to secondary progressive multiple sclerosis. Mult Scler Relat Disord. 2021;54: 103115.

    Article  PubMed  Google Scholar 

  42. Liu YC, Tsai YH, Tang SC, Liou HC, Kang KH, Liou HH, et al. Cytokine MIF enhances blood-brain barrier permeability: impact for therapy in ischemic stroke. Sci Rep. 2018;8(1):743.

    Article  PubMed  PubMed Central  Google Scholar 

  43. Niino M, Ogata A, Kikuchi S, Tashiro K, Nishihira J. Macrophage migration inhibitory factor in the cerebrospinal fluid of patients with conventional and optic-spinal forms of multiple sclerosis and neuro-Behcet’s disease. J Neurol Sci. 2000;179(S 1–2):127–31.

    Article  CAS  PubMed  Google Scholar 

  44. Rinta S, Kuusisto H, Raunio M, Paalavuo R, Levula M, Lehtimaki T, et al. Apoptosis-related molecules in blood in multiple sclerosis. J Neuroimmunol. 2008;205(1–2):135–41.

    Article  CAS  PubMed  Google Scholar 

  45. Hagman S, Raunio M, Rossi M, Dastidar P, Elovaara I. Disease-associated inflammatory biomarker profiles in blood in different subtypes of multiple sclerosis: prospective clinical and MRI follow-up study. J Neuroimmunol. 2011;234(1–2):141–7.

    Article  CAS  PubMed  Google Scholar 

  46. Cavalli E, Mazzon E, Basile MS, Mangano K, Di Marco R, Bramanti P, Nicoletti F, Fagone P, Petralia MC. Upregulated Expression of Macrophage Migration Inhibitory Factor, Its Analogue D-Dopachrome Tautomerase, and the CD44 Receptor in Peripheral CD4 T Cells from Clinically Isolated Syndrome Patients with Rapid Conversion to Clinical Defined Multiple Sclerosis. Medicina (Kaunas). 2019;55(10):667. https://doi.org/10.3390/medicina55100667.

  47. Rijvers L, Melief MJ, van der Vuurst de Vries RM, Stephant M, van Langelaar J, Wierenga-Wolf AF, et al. The macrophage migration inhibitory factor pathway in human B cells is tightly controlled and dysregulated in multiple sclerosis. Eur J Immunol. 2018;48(11):1861–71.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Cox GM, Kithcart AP, Pitt D, Guan Z, Alexander J, Williams JL, et al. Macrophage migration inhibitory factor potentiates autoimmune-mediated neuroinflammation. J Immunol. 2013;191(3):1043–54.

    Article  CAS  PubMed  Google Scholar 

  49. Khaibullin T, Ivanova V, Martynova E, Cherepnev G, Khabirov F, Granatov E, et al. Elevated levels of proinflammatory cytokines in cerebrospinal fluid of multiple sclerosis patients. Front Immunol. 2017;8:531.

    Article  PubMed  PubMed Central  Google Scholar 

  50. Hjaeresen S, Sejbaek T, Axelsson M, Mortensen SK, Vinslov-Jensen H, Pihl-Jensen G, et al. MIF in the cerebrospinal fluid is decreased during relapsing-remitting while increased in secondary progressive multiple sclerosis. J Neurol Sci. 2022;439: 120320.

    Article  PubMed  Google Scholar 

  51. Benedek G, Chaudhary P, Meza-Romero R, Calkins E, Kent G, Offner H, et al. Sex-dependent treatment of chronic EAE with partial MHC class II constructs. J Neuroinflamm. 2017;14(1):100.

    Article  Google Scholar 

  52. Benedek G, Meza-Romero R, Jordan K, Zhang Y, Nguyen H, Kent G, et al. MIF and D-DT are potential disease severity modifiers in male MS subjects. Proc Natl Acad Sci U S A. 2017;114(40):E8421–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. De la Cruz-Mosso U, Bucala R, Palafox-Sanchez CA, Parra-Rojas I, Padilla-Gutierrez JR, Pereira-Suarez AL, et al. Macrophage migration inhibitory factor: association of -794 CATT5-8 and -173 G>C polymorphisms with TNF-alpha in systemic lupus erythematosus. Hum Immunol. 2014;75(5):433–9.

    Article  PubMed  PubMed Central  Google Scholar 

  54. Llamas-Covarrubias MA, Valle Y, Bucala R, Navarro-Hernandez RE, Palafox-Sanchez CA, Padilla-Gutierrez JR, et al. Macrophage migration inhibitory factor (MIF): genetic evidence for participation in early onset and early stage rheumatoid arthritis. Cytokine. 2013;61(3):759–65.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. Baugh JA, Chitnis S, Donnelly SC, Monteiro J, Lin X, Plant BJ, 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(3):170–6.

    Article  CAS  PubMed  Google Scholar 

  56. Sreih A, Ezzeddine R, Leng L, LaChance A, Yu G, Mizue Y, et al. Dual effect of the macrophage migration inhibitory factor gene on the development and severity of human systemic lupus erythematosus. Arthritis Rheum. 2011;63(12):3942–51.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  57. Morales-Zambrano R, Bautista-Herrera LA, De la Cruz-Mosso U, Villanueva-Quintero GD, Padilla-Gutierrez JR, Valle Y, et al. Macrophage migration inhibitory factor (MIF) promoter polymorphisms (-794 CATT5-8 and -173 G>C): association with MIF and TNFalpha in psoriatic arthritis. Int J Clin Exp Med. 2014;7(9):2605–14.

    PubMed  PubMed Central  Google Scholar 

  58. Akcali A, Pehlivan S, Pehlivan M, Sever T, Neyal M. Association of macrophage migration inhibitory factor gene promoter polymorphisms with multiple sclerosis in Turkish patients. J Int Med Res. 2010;38(1):69–77.

    Article  CAS  PubMed  Google Scholar 

  59. Cevik B, Yigit S, Karakus N, Aksoy D, Ates O, Kurt S. Lack of association between MIF gene -173G>C polymorphism with multiple sclerosis. In Vivo. 2015;29(1):71–6.

    CAS  PubMed  Google Scholar 

  60. Castaneda Moreno VA, Muñoz-Valle JF, Torres Carrillo N, Gonzalez Perez OP, Macias Islas MA, Ruiz Sandoval JL, Padilla De La Torre O, Trujillo Trujillo XA, Huerta Vieyra M. A case-control study on the association of MIF-794 CATT5-8 and-173 G> C polymorphisms and its serum levels and the clinical severity of multiple sclerosis in Mexican patients. Front Immunol. 2015. https://doi.org/10.3389/conf.fimmu.2015.05.00227.

  61. Castaneda-Moreno VA, De la Cruz-Mosso U, Torres-Carrillo N, Macias-Islas MA, Padilla-De la Torre O, Mireles-Ramirez MA, et al. MIF functional polymorphisms (-794 CATT(5–8) and -173 G>C) are associated with MIF serum levels, severity and progression in male multiple sclerosis from western Mexican population. J Neuroimmunol. 2018;320:117–24.

    Article  CAS  PubMed  Google Scholar 

  62. Han Z, Qu J, Zhao J, Zou X. Genetic variant rs755622 regulates expression of the multiple sclerosis severity modifier D-dopachrome tautomerase in a sex-specific way. Biomed Res Int. 2018;2018:8285653.

    Article  PubMed  PubMed Central  Google Scholar 

  63. Fagone P, Mazzon E, Cavalli E, Bramanti A, Petralia MC, Mangano K, et al. Contribution of the macrophage migration inhibitory factor superfamily of cytokines in the pathogenesis of preclinical and human multiple sclerosis: in silico and in vivo evidences. J Neuroimmunol. 2018;322:46–56.

    Article  CAS  PubMed  Google Scholar 

  64. Powell ND, Papenfuss TL, McClain MA, Gienapp IE, Shawler TM, Satoskar AR, et al. Cutting edge: macrophage migration inhibitory factor is necessary for progression of experimental autoimmune encephalomyelitis. J Immunol. 2005;175(9):5611–4.

    Article  CAS  PubMed  Google Scholar 

  65. Benedek G, Meza-Romero R, Andrew S, Leng L, Burrows GG, Bourdette D, et al. Partial MHC class II constructs inhibit MIF/CD74 binding and downstream effects. Eur J Immunol. 2013;43(5):1309–21.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  66. Meza-Romero R, Benedek G, Yu X, Mooney JL, Dahan R, Duvshani N, et al. HLA-DRalpha1 constructs block CD74 expression and MIF effects in experimental autoimmune encephalomyelitis. J Immunol. 2014;192(9):4164–73.

    Article  CAS  PubMed  Google Scholar 

  67. Benedek G, Meza-Romero R, Jordan K, Keenlyside L, Offner H, Vandenbark AA. HLA-DRalpha1-mMOG-35-55 treatment of experimental autoimmune encephalomyelitis reduces CNS inflammation, enhances M2 macrophage frequency, and promotes neuroprotection. J Neuroinflamm. 2015;12:123.

    Article  Google Scholar 

  68. Ji N, Kovalovsky A, Fingerle-Rowson G, Guentzel MN, Forsthuber TG. Macrophage migration inhibitory factor promotes resistance to glucocorticoid treatment in EAE. Neurol Neuroimmunol Neuroinflamm. 2015;2(5): e139.

    Article  PubMed  PubMed Central  Google Scholar 

  69. Denkinger CM, Denkinger M, Kort JJ, Metz C, Forsthuber TG. In vivo blockade of macrophage migration inhibitory factor ameliorates acute experimental autoimmune encephalomyelitis by impairing the homing of encephalitogenic T cells to the central nervous system. J Immunol. 2003;170(3):1274–82.

    Article  CAS  PubMed  Google Scholar 

  70. Kithcart AP, Cox GM, Sielecki T, Short A, Pruitt J, Papenfuss T, et al. A small-molecule inhibitor of macrophage migration inhibitory factor for the treatment of inflammatory disease. FASEB J. 2010;24(11):4459–66.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  71. Cho Y, Crichlow GV, Vermeire JJ, Leng L, Du X, Hodsdon ME, et al. Allosteric inhibition of macrophage migration inhibitory factor revealed by ibudilast. Proc Natl Acad Sci U S A. 2010;107(25):11313–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  72. Schwenkgrub J, Zaremba M, Mirowska-Guzel D, Kurkowska-Jastrzebska I. Ibudilast: a non-selective phosphodiesterase inhibitor in brain disorders. Postepy Hig Med Dosw (Online). 2017;71:137–48.

    Article  PubMed  Google Scholar 

  73. Oyama R, Yamamoto H, Titani K. Glutamine synthetase, hemoglobin alpha-chain, and macrophage migration inhibitory factor binding to amyloid beta-protein: their identification in rat brain by a novel affinity chromatography and in Alzheimer’s disease brain by immunoprecipitation. Biochim Biophys Acta. 2000;1479(1–2):91–102.

    Article  CAS  PubMed  Google Scholar 

  74. Lashuel HA, Aljabari B, Sigurdsson EM, Metz CN, Leng L, Callaway DJ, et al. Amyloid fibril formation by macrophage migration inhibitory factor. Biochem Biophys Res Commun. 2005;338(2):973–80.

    Article  CAS  PubMed  Google Scholar 

  75. Bryan KJ, Zhu X, Harris PL, Perry G, Castellani RJ, Smith MA, et al. Expression of CD74 is increased in neurofibrillary tangles in Alzheimer’s disease. Mol Neurodegener. 2008;3:13.

    Article  PubMed  PubMed Central  Google Scholar 

  76. Yoshiyama Y, Arai K, Oki T, Hattori T. Expression of invariant chain and pro-cathepsin L in Alzheimer’s brain. Neurosci Lett. 2000;290(2):125–8.

    Article  CAS  PubMed  Google Scholar 

  77. Bacher M, Deuster O, Aljabari B, Egensperger R, Neff F, Jessen F, et al. The role of macrophage migration inhibitory factor in Alzheimer’s disease. Mol Med. 2010;16(3–4):116–21.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  78. Carlred L, Michno W, Kaya I, Sjovall P, Syvanen S, Hanrieder J. Probing amyloid-beta pathology in transgenic Alzheimer’s disease (tgArcSwe) mice using MALDI imaging mass spectrometry. J Neurochem. 2016;138(3):469–78.

    Article  CAS  PubMed  Google Scholar 

  79. Li SQ, Yu Y, Han JZ, Wang D, Liu J, Qian F, et al. Deficiency of macrophage migration inhibitory factor attenuates tau hyperphosphorylation in mouse models of Alzheimer’s disease. J Neuroinflamm. 2015;12:177.

    Article  Google Scholar 

  80. Liang CJ, Li JH, Zhang Z, Zhang JY, Liu SQ, Yang J. Suppression of MIF-induced neuronal apoptosis may underlie the therapeutic effects of effective components of Fufang Danshen in the treatment of Alzheimer’s disease. Acta Pharmacol Sin. 2018;39(9):1421–38.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  81. Al-Abed Y, VanPatten S. MIF as a disease target: ISO-1 as a proof-of-concept therapeutic. Future Med Chem. 2011;3(1):45–63.

    Article  CAS  PubMed  Google Scholar 

  82. Nasiri E, Sankowski R, Dietrich H, Oikonomidi A, Huerta PT, Popp J, et al. Key role of MIF-related neuroinflammation in neurodegeneration and cognitive impairment in Alzheimer’s disease. Mol Med. 2020;26(1):34.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  83. Vitek MP, Bhattacharya K, Glendening JM, Stopa E, Vlassara H, Bucala R, et al. Advanced glycation end products contribute to amyloidosis in Alzheimer disease. Proc Natl Acad Sci U S A. 1994;91(11):4766–70.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  84. Sato T, Shimogaito N, Wu X, Kikuchi S, Yamagishi S, Takeuchi M. Toxic advanced glycation end products (TAGE) theory in Alzheimer’s disease. Am J Alzheimers Dis Other Demen. 2006;21(3):197–208.

    Article  PubMed  Google Scholar 

  85. Yu M, Zang D, Xu Y, Meng J, Qian S. Protective effect of ISO-1 against advanced glycation end product aggravation of PC12 cell injury induced by Abeta1-40. Mol Med Rep. 2019;20(3):2135–42.

    CAS  PubMed  PubMed Central  Google Scholar 

  86. Kassaar O, Pereira Morais M, Xu S, Adam EL, Chamberlain RC, Jenkins B, et al. Macrophage migration inhibitory factor is subjected to glucose modification and oxidation in Alzheimer’s Disease. Sci Rep. 2017;7:42874.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  87. Popp J, Bacher M, Kolsch H, Noelker C, Deuster O, Dodel R, et al. Macrophage migration inhibitory factor in mild cognitive impairment and Alzheimer’s disease. J Psychiatr Res. 2009;43(8):749–53.

    Article  PubMed  Google Scholar 

  88. Oikonomidi A, Tautvydaite D, Gholamrezaee MM, Henry H, Bacher M, Popp J. Macrophage migration inhibitory factor is associated with biomarkers of Alzheimer’s disease pathology and predicts cognitive decline in mild cognitive impairment and mild dementia. J Alzheimers Dis. 2017;60(1):273–81.

    Article  CAS  PubMed  Google Scholar 

  89. Zhang S, Zhao J, Zhang Y, Zhang Y, Cai F, Wang L, et al. Upregulation of MIF as a defense mechanism and a biomarker of Alzheimer’s disease. Alzheimers Res Ther. 2019;11(1):54.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  90. Bae SH, Yoo MR, Kim YY, Hong IK, Kim MH, Lee SH, et al. Brain-derived neurotrophic factor mediates macrophage migration inhibitory factor to protect neurons against oxygen-glucose deprivation. Neural Regen Res. 2020;15(8):1483–9.

    Article  PubMed  PubMed Central  Google Scholar 

  91. Moon HY, Kim SH, Yang YR, Song P, Yu HS, Park HG, et al. Macrophage migration inhibitory factor mediates the antidepressant actions of voluntary exercise. Proc Natl Acad Sci U S A. 2012;109(32):13094–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  92. Wesseling P, Capper D. WHO 2016 Classification of gliomas. Neuropathol Appl Neurobiol. 2018;44(2):139–50.

    Article  CAS  PubMed  Google Scholar 

  93. Wenger A, Ferreyra Vega S, Kling T, Bontell TO, Jakola AS, Caren H. Intratumor DNA methylation heterogeneity in glioblastoma: implications for DNA methylation-based classification. Neuro Oncol. 2019;21(5):616–27.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  94. Melin BS, Barnholtz-Sloan JS, Wrensch MR, Johansen C, Il’yasova D, Kinnersley B, et al. Genome-wide association study of glioma subtypes identifies specific differences in genetic susceptibility to glioblastoma and non-glioblastoma tumors. Nat Genet. 2017;49(5):789–94.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  95. Nobre CC, de Araujo JM, Fernandes TA, Cobucci RN, Lanza DC, Andrade VS, et al. Macrophage migration inhibitory factor (MIF): biological activities and relation with cancer. Pathol Oncol Res. 2017;23(2):235–44.

    Article  CAS  PubMed  Google Scholar 

  96. Bach JP, Deuster O, Balzer-Geldsetzer M, Meyer B, Dodel R, Bacher M. The role of macrophage inhibitory factor in tumorigenesis and central nervous system tumors. Cancer. 2009;115(10):2031–40.

    Article  CAS  PubMed  Google Scholar 

  97. Bucala R, Donnelly SC. Macrophage migration inhibitory factor: a probable link between inflammation and cancer. Immunity. 2007;26(3):281–5.

    Article  CAS  PubMed  Google Scholar 

  98. Kindt N, Journe F, Laurent G, Saussez S. Involvement of macrophage migration inhibitory factor in cancer and novel therapeutic targets. Oncol Lett. 2016;12(4):2247–53.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  99. Simpson KD, Templeton DJ, Cross JV. Macrophage migration inhibitory factor promotes tumor growth and metastasis by inducing myeloid-derived suppressor cells in the tumor microenvironment. J Immunol. 2012;189(12):5533–40.

    Article  CAS  PubMed  Google Scholar 

  100. Apte RS, Sinha D, Mayhew E, Wistow GJ, Niederkorn JY. Cutting edge: role of macrophage migration inhibitory factor in inhibiting NK cell activity and preserving immune privilege. J Immunol. 1998;160(12):5693–6.

    Article  CAS  PubMed  Google Scholar 

  101. Yan X, Orentas RJ, Johnson BD. Tumor-derived macrophage migration inhibitory factor (MIF) inhibits T lymphocyte activation. Cytokine. 2006;33(4):188–98.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  102. Ghoochani A, Schwarz MA, Yakubov E, Engelhorn T, Doerfler A, Buchfelder M, et al. MIF-CD74 signaling impedes microglial M1 polarization and facilitates brain tumorigenesis. Oncogene. 2016;35(48):6246–61.

    Article  CAS  PubMed  Google Scholar 

  103. Alban TJ, Bayik D, Otvos B, Rabljenovic A, Leng L, Jia-Shiun L, et al. Glioblastoma myeloid-derived suppressor cell subsets express differential macrophage migration inhibitory factor receptor profiles that can be targeted to reduce immune suppression. Front Immunol. 2020;11:1191.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  104. Mittelbronn M, Platten M, Zeiner P, Dombrowski Y, Frank B, Zachskorn C, et al. Macrophage migration inhibitory factor (MIF) expression in human malignant gliomas contributes to immune escape and tumour progression. Acta Neuropathol. 2011;122(3):353–65.

    Article  CAS  PubMed  Google Scholar 

  105. Zeiner PS, Preusse C, Blank AE, Zachskorn C, Baumgarten P, Caspary L, et al. MIF receptor CD74 is restricted to microglia/macrophages, associated with a M1-polarized immune milieu and prolonged patient survival in gliomas. Brain Pathol. 2015;25(4):491–504.

    Article  CAS  PubMed  Google Scholar 

  106. Wang D, Luo L, Chen W, Chen LZ, Zeng WT, Li W, et al. Significance of the vascular endothelial growth factor and the macrophage migration inhibitory factor in the progression of hepatocellular carcinoma. Oncol Rep. 2014;31(3):1199–204.

    Article  CAS  PubMed  Google Scholar 

  107. Chang KP, Lin SJ, Liu SC, Yi JS, Chien KY, Chi LM, et al. Low-molecular-mass secretome profiling identifies HMGA2 and MIF as prognostic biomarkers for oral cavity squamous cell carcinoma. Sci Rep. 2015;5:11689.

    Article  PubMed  PubMed Central  Google Scholar 

  108. Vera PL, Meyer-Siegler KL. Association between macrophage migration inhibitory factor promoter region polymorphism (-173 G/C) and cancer: a meta-analysis. BMC Res Notes. 2011;4:395.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  109. De Souza MB, Curioni OA, Kanda JL, De Carvalho MB. Serum and salivary macrophage migration inhibitory factor in patients with oral squamous cell carcinoma. Oncol Lett. 2014;8(5):2267–75.

    Article  PubMed  Google Scholar 

  110. Bacher M, Schrader J, Thompson N, Kuschela K, Gemsa D, Waeber G, et al. Up-regulation of macrophage migration inhibitory factor gene and protein expression in glial tumor cells during hypoxic and hypoglycemic stress indicates a critical role for angiogenesis in glioblastoma multiforme. Am J Pathol. 2003;162(1):11–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  111. Wang XB, Tian XY, Li Y, Li B, Li Z. Elevated expression of macrophage migration inhibitory factor correlates with tumor recurrence and poor prognosis of patients with gliomas. J Neurooncol. 2012;106(1):43–51.

    Article  PubMed  Google Scholar 

  112. Baron N, Deuster O, Noelker C, Stuer C, Strik H, Schaller C, et al. Role of macrophage migration inhibitory factor in primary glioblastoma multiforme cells. J Neurosci Res. 2011;89(5):711–7.

    Article  CAS  PubMed  Google Scholar 

  113. Baugh JA, Gantier M, Li L, Byrne A, Buckley A, Donnelly SC. Dual regulation of macrophage migration inhibitory factor (MIF) expression in hypoxia by CREB and HIF-1. Biochem Biophys Res Commun. 2006;347(4):895–903.

    Article  CAS  PubMed  Google Scholar 

  114. Guo X, Xu S, Gao X, Wang J, Xue H, Chen Z, et al. Macrophage migration inhibitory factor promotes vasculogenic mimicry formation induced by hypoxia via CXCR4/AKT/EMT pathway in human glioblastoma cells. Oncotarget. 2017;8(46):80358–72.

    Article  PubMed  PubMed Central  Google Scholar 

  115. Bacher M, Meinhardt A, Lan HY, Dhabhar FS, Mu W, Metz CN, et al. MIF expression in the rat brain: implications for neuronal function. Mol Med. 1998;4(4):217–30.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  116. Wei J, Gabrusiewicz K, Heimberger A. The controversial role of microglia in malignant gliomas. Clin Dev Immunol. 2013;2013: 285246.

    Article  PubMed  PubMed Central  Google Scholar 

  117. Gieryng A, Pszczolkowska D, Walentynowicz KA, Rajan WD, Kaminska B. Immune microenvironment of gliomas. Lab Invest. 2017;97(5):498–518.

    Article  CAS  PubMed  Google Scholar 

  118. Chen Z, Feng X, Herting CJ, Garcia VA, Nie K, Pong WW, et al. Cellular and molecular identity of tumor-associated macrophages in glioblastoma. Can Res. 2017;77(9):2266–78.

    Article  CAS  Google Scholar 

  119. Otvos B, Silver DJ, Mulkearns-Hubert EE, Alvarado AG, Turaga SM, Sorensen MD, et al. Cancer stem cell-secreted macrophage migration inhibitory factor stimulates myeloid derived suppressor cell function and facilitates glioblastoma immune evasion. Stem Cells. 2016;34(8):2026–39.

    Article  CAS  PubMed  Google Scholar 

  120. Kumar R, de Mooij T, Peterson TE, Kaptzan T, Johnson AJ, Daniels DJ, et al. Modulating glioma-mediated myeloid-derived suppressor cell development with sulforaphane. PLoS ONE. 2017;12(6): e0179012.

    Article  PubMed  PubMed Central  Google Scholar 

  121. Alban TJ, Alvarado AG, Sorensen MD, Bayik D, Volovetz J, Serbinowski E, Mulkearns-Hubert EE, Sinyuk M, Hale JS, Onzi GR, McGraw M, Huang P, Grabowski MM, Wathen CA, Ahluwalia MS, Radivoyevitch T, Kornblum HI, Kristensen BW, Vogelbaum MA, Lathia JD. Global immune fingerprinting in glioblastoma patient peripheral blood reveals immune-suppression signatures associated with prognosis. JCI Insight. 2018;3(21):e122264. https://doi.org/10.1172/jci.insight.122264.

  122. Maltby S, Khazaie K, McNagny KM. Mast cells in tumor growth: angiogenesis, tissue remodelling and immune-modulation. Biochim Biophys Acta. 2009;1796(1):19–26.

    CAS  PubMed  PubMed Central  Google Scholar 

  123. Polajeva J, Bergstrom T, Edqvist PH, Lundequist A, Sjosten A, Nilsson G, et al. Glioma-derived macrophage migration inhibitory factor (MIF) promotes mast cell recruitment in a STAT5-dependent manner. Mol Oncol. 2014;8(1):50–8.

    Article  CAS  PubMed  Google Scholar 

  124. Ohta S, Kawakami Y, Okano H. MIF: functions in brain and glioblastoma. Oncotarget. 2017;8(29):46706–7.

    Article  PubMed  PubMed Central  Google Scholar 

  125. Wang Z, Xue Y, Wang P, Zhu J, Ma J. MiR-608 inhibits the migration and invasion of glioma stem cells by targeting macrophage migration inhibitory factor. Oncol Rep. 2016;35(5):2733–42.

    Article  CAS  PubMed  Google Scholar 

  126. Fukaya R, Ohta S, Yaguchi T, Matsuzaki Y, Sugihara E, Okano H, et al. MIF maintains the tumorigenic capacity of brain tumor-initiating cells by directly inhibiting p53. Cancer Res. 2016;76(9):2813–23.

    Article  CAS  PubMed  Google Scholar 

  127. Jankauskas SS, Wong DWL, Bucala R, Djudjaj S, Boor P. Evolving complexity of MIF signaling. Cell Signal. 2019;57:76–88.

    Article  CAS  PubMed  Google Scholar 

  128. Shi X, Leng L, Wang T, Wang W, Du X, Li J, et al. CD44 is the signaling component of the macrophage migration inhibitory factor-CD74 receptor complex. Immunity. 2006;25(4):595–606.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  129. McCubrey JA, Steelman LS, Chappell WH, Abrams SL, Wong EW, Chang F, et al. Roles of the Raf/MEK/ERK pathway in cell growth, malignant transformation and drug resistance. Biochim Biophys Acta. 2007;1773(8):1263–84.

    Article  CAS  PubMed  Google Scholar 

  130. Manning BD, Toker A. AKT/PKB signaling: navigating the network. Cell. 2017;169(3):381–405.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  131. Lue H, Thiele M, Franz J, Dahl E, Speckgens S, Leng L, et al. Macrophage migration inhibitory factor (MIF) promotes cell survival by activation of the Akt pathway and role for CSN5/JAB1 in the control of autocrine MIF activity. Oncogene. 2007;26(35):5046–59.

    Article  CAS  PubMed  Google Scholar 

  132. Fingerle-Rowson G, Petrenko O, Metz CN, Forsthuber TG, Mitchell R, Huss R, et al. The p53-dependent effects of macrophage migration inhibitory factor revealed by gene targeting. Proc Natl Acad Sci U S A. 2003;100(16):9354–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  133. Tanese K, Hashimoto Y, Berkova Z, Wang Y, Samaniego F, Lee JE, et al. Cell surface CD74-MIF interactions drive melanoma survival in response to interferon-gamma. J Invest Dermatol. 2015;135(11):2775–84.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  134. Merk M, Zierow S, Leng L, Das R, Du X, Schulte W, 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):E577–85.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  135. Kobold S, Merk M, Hofer L, Peters P, Bucala R, Endres S. The macrophage migration inhibitory factor (MIF)-homologue D-dopachrome tautomerase is a therapeutic target in a murine melanoma model. Oncotarget. 2014;5(1):103–7.

    Article  PubMed  Google Scholar 

  136. Brock SE, Rendon BE, Yaddanapudi K, Mitchell RA. Negative regulation of AMP-activated protein kinase (AMPK) activity by macrophage migration inhibitory factor (MIF) family members in non-small cell lung carcinomas. J Biol Chem. 2012;287(45):37917–25.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  137. Coleman AM, Rendon BE, Zhao M, Qian MW, Bucala R, Xin D, et al. Cooperative regulation of non-small cell lung carcinoma angiogenic potential by macrophage migration inhibitory factor and its homolog, D-dopachrome tautomerase. J Immunol. 2008;181(4):2330–7.

    Article  CAS  PubMed  Google Scholar 

  138. Guo D, Guo J, Yao J, Jiang K, Hu J, Wang B, et al. D-dopachrome tautomerase is over-expressed in pancreatic ductal adenocarcinoma and acts cooperatively with macrophage migration inhibitory factor to promote cancer growth. Int J Cancer. 2016;139(9):2056–67.

    Article  CAS  PubMed  Google Scholar 

  139. Winner M, Meier J, Zierow S, Rendon BE, Crichlow GV, Riggs R, et al. A novel, macrophage migration inhibitory factor suicide substrate inhibits motility and growth of lung cancer cells. Cancer Res. 2008;68(18):7253–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  140. Lee SH, Kwon HJ, Park S, Kim CI, Ryu H, Kim SS, et al. Macrophage migration inhibitory factor (MIF) inhibitor 4-IPP downregulates stemness phenotype and mesenchymal trans-differentiation after irradiation in glioblastoma multiforme. PLoS ONE. 2021;16(9): e0257375.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  141. Pasupuleti V, Du W, Gupta Y, Yeh IJ, Montano M, Magi-Galuzzi C, et al. Dysregulated D-dopachrome tautomerase, a hypoxia-inducible factor-dependent gene, cooperates with macrophage migration inhibitory factor in renal tumorigenesis. J Biol Chem. 2014;289(6):3713–23.

    Article  CAS  PubMed  Google Scholar 

  142. Mangano K, Mazzon E, Basile MS, Di Marco R, Bramanti P, Mammana S, et al. Pathogenic role for macrophage migration inhibitory factor in glioblastoma and its targeting with specific inhibitors as novel tailored therapeutic approach. Oncotarget. 2018;9(25):17951–70.

    Article  PubMed  PubMed Central  Google Scholar 

  143. Karim R, Palazzo C, Evrard B, Piel G. Nanocarriers for the treatment of glioblastoma multiforme: current state-of-the-art. J Control Release. 2016;227:23–37.

    Article  CAS  PubMed  Google Scholar 

  144. Glaser T, Han I, Wu L, Zeng X. Targeted nanotechnology in glioblastoma multiforme. Front Pharmacol. 2017;8:166.

    Article  PubMed  PubMed Central  Google Scholar 

  145. Schrader J, Deuster O, Rinn B, Schulz M, Kautz A, Dodel R, et al. Restoration of contact inhibition in human glioblastoma cell lines after MIF knockdown. BMC Cancer. 2009;9:464.

    Article  PubMed  PubMed Central  Google Scholar 

  146. Piette C, Deprez M, Roger T, Noel A, Foidart JM, Munaut C. The dexamethasone-induced inhibition of proliferation, migration, and invasion in glioma cell lines is antagonized by macrophage migration inhibitory factor (MIF) and can be enhanced by specific MIF inhibitors. J Biol Chem. 2009;284(47):32483–92.

    Article  PubMed  PubMed Central  Google Scholar 

  147. Kitange GJ, Carlson BL, Schroeder MA, Decker PA, Morlan BW, Wu W, et al. Expression of CD74 in high grade gliomas: a potential role in temozolomide resistance. J Neurooncol. 2010;100(2):177–86.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  148. Ha W, Sevim-Nalkiran H, Zaman AM, Matsuda K, Khasraw M, Nowak AK, et al. Ibudilast sensitizes glioblastoma to temozolomide by targeting macrophage migration inhibitory factor (MIF). Sci Rep. 2019;9(1):2905.

    Article  PubMed  PubMed Central  Google Scholar 

  149. Berkova Z, Tao RH, Samaniego F. Milatuzumab—a promising new immunotherapeutic agent. Expert Opin Investig Drugs. 2010;19(1):141–9.

    Article  CAS  PubMed  Google Scholar 

  150. Gore Y, Starlets D, Maharshak N, Becker-Herman S, Kaneyuki U, Leng L, et al. Macrophage migration inhibitory factor induces B cell survival by activation of a CD74-CD44 receptor complex. J Biol Chem. 2008;283(5):2784–92.

    Article  CAS  PubMed  Google Scholar 

  151. Wallace DJ, Figueras F, Wegener WA, Goldenberg DM. Experience with milatuzumab, an anti-CD74 antibody against immunomodulatory macrophage migration inhibitory factor (MIF) receptor, for systemic lupus erythematosus (SLE). Ann Rheum Dis. 2021;80(7):954–5.

    Article  PubMed  Google Scholar 

  152. Mahalingam D, Patel MR, Sachdev JC, Hart LL, Halama N, Ramanathan RK, et al. Phase I study of imalumab (BAX69), a fully human recombinant antioxidized macrophage migration inhibitory factor antibody in advanced solid tumours. Br J Clin Pharmacol. 2020;86(9):1836–48.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  153. Pantouris G, Syed MA, Fan C, Rajasekaran D, Cho TY, Rosenberg EM Jr, et al. An analysis of MIF structural features that control functional activation of CD74. Chem Biol. 2015;22(9):1197–205.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  154. Fingerle-Rowson G, Kaleswarapu DR, Schlander C, Kabgani N, Brocks T, Reinart N, 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(7):1922–32.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  155. Orita M, Yamamoto S, Katayama N, Fujita S. Macrophage migration inhibitory factor and the discovery of tautomerase inhibitors. Curr Pharm Des. 2002;8(14):1297–317.

    Article  CAS  PubMed  Google Scholar 

  156. Lubetsky JB, Dios A, Han J, Aljabari B, Ruzsicska B, Mitchell R, 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):24976–82.

    Article  CAS  PubMed  Google Scholar 

  157. Cournia Z, Leng L, Gandavadi S, Du X, Bucala R, Jorgensen WL. Discovery of human macrophage migration inhibitory factor (MIF)-CD74 antagonists via virtual screening. J Med Chem. 2009;52(2):416–24.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  158. Hare AA, Leng L, Gandavadi S, Du X, Cournia Z, Bucala R, et al. Optimization of N-benzyl-benzoxazol-2-ones as receptor antagonists of macrophage migration inhibitory factor (MIF). Bioorg Med Chem Lett. 2010;20(19):5811–4.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  159. Leng L, Chen L, Fan J, Greven D, Arjona A, Du X, et al. A small-molecule macrophage migration inhibitory factor antagonist protects against glomerulonephritis in lupus-prone NZB/NZW F1 and MRL/lpr mice. J Immunol. 2011;186(1):527–38.

    Article  CAS  PubMed  Google Scholar 

  160. Nakamura A, Zeng F, Nakamura S, Reid KT, Gracey E, Lim M, et al. Macrophage migration inhibitory factor drives pathology in a mouse model of spondyloarthritis and is associated with human disease. Sci Transl Med. 2021;13(616):eabg1210.

    Article  CAS  PubMed  Google Scholar 

  161. Fox RJ, Coffey CS, Conwit R, Cudkowicz ME, Gleason T, Goodman A, et al. Phase 2 trial of ibudilast in progressive multiple sclerosis. N Engl J Med. 2018;379(9):846–55.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  162. Tilstam PV, Pantouris G, Corman M, Andreoli M, Mahboubi K, Davis G, et al. A selective small-molecule inhibitor of macrophage migration inhibitory factor-2 (MIF-2), a MIF cytokine superfamily member, inhibits MIF-2 biological activity. J Biol Chem. 2019;294(49):18522–31.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  163. Xiao Z, Osipyan A, Song S, Chen D, Schut RA, van Merkerk R, et al. Thieno[2,3-d]pyrimidine-2,4(1H,3H)-dione derivative inhibits d-dopachrome tautomerase activity and suppresses the proliferation of non-small cell lung cancer cells. J Med Chem. 2022;65(3):2059–77.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  164. 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):4961–71.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  165. Meza-Romero R, Benedek G, Jordan K, Leng L, Pantouris G, Lolis E, 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.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  166. Burrows GG, Chang JW, Bachinger HP, Bourdette DN, Offner H, Vandenbark AA. Design, engineering and production of functional single-chain T cell receptor ligands. Protein Eng. 1999;12(9):771–8.

    Article  CAS  PubMed  Google Scholar 

  167. Burrows GG. Systemic immunomodulation of autoimmune disease using MHC-derived recombinant TCR ligands. Curr Drug Targets Inflamm Allergy. 2005;4(2):185–93.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  168. Offner H, Sinha S, Wang C, Burrows GG, Vandenbark AA. Recombinant T cell receptor ligands: immunomodulatory, neuroprotective and neuroregenerative effects suggest application as therapy for multiple sclerosis. Rev Neurosci. 2008;19(4–5):327–39.

    CAS  PubMed  PubMed Central  Google Scholar 

  169. Vandenbark AA, Meza-Romero R, Benedek G, Andrew S, Huan J, Chou YK, 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.

    Article  CAS  PubMed  Google Scholar 

  170. Vandenbark AA, Rich C, Mooney J, Zamora A, Wang C, Huan J, et al. Recombinant TCR ligand induces tolerance to myelin oligodendrocyte glycoprotein 35–55 peptide and reverses clinical and histological signs of chronic experimental autoimmune encephalomyelitis in HLA-DR2 transgenic mice. J Immunol. 2003;171(1):127–33.

    Article  CAS  PubMed  Google Scholar 

  171. Yadav V, Bourdette DN, Bowen JD, Lynch SG, Mattson D, Preiningerova J, et al. Recombinant T-cell receptor ligand (RTL) for treatment of multiple sclerosis: a double-blind, placebo-controlled, phase 1, dose-escalation study. Autoimmune Dis. 2012;2012: 954739.

    PubMed  PubMed Central  Google Scholar 

  172. 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.

    Article  CAS  PubMed  Google Scholar 

  173. Vandenbark AA, Meza-Romero R, Benedek G, Offner H. A novel neurotherapeutic for multiple sclerosis, ischemic injury, methamphetamine addiction, and traumatic brain injury. J Neuroinflamm. 2019;16(1):14.

    Article  Google Scholar 

  174. Vandenbark AA, Meza-Romero R, Wiedrick J, Gerstner G, Headrick A, Kent G, et al. Brief report: enhanced DRalpha1-mMOG-35-55 treatment of severe EAE in MIF-1-deficient male mice. Cell Immunol. 2021;370: 104439.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  175. Vandenbark AA, Meza-Romero R, Wiedrick J, Gerstner G, Seifert H, Kent G, et al. “Near Cure” treatment of severe acute EAE in MIF-1-deficient female and male mice with a bifunctional MHCII-derived molecular construct. Cell Immunol. 2022;378: 104561.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  176. Meza-Romero R, Benedek G, Gerstner G, Kent G, Nguyen H, Offner H, et al. Increased CD74 binding and EAE treatment efficacy of a modified DRalpha1 molecular construct. Metab Brain Dis. 2019;34(1):153–64.

    Article  CAS  PubMed  Google Scholar 

Download references

Funding

Funding for this work was provided in part by: National Institute of Arthritis and Musculoskeletal and Skin Diseases (AR078334), Rheumatology Research Foundation, G. Harold and Leila Y. Mathers Charitable Foundation (RB); National Institute of Allergy and Infectious Diseases (AI148409) (HO); Biomedical Laboratory Research and Development, VA Office of Research and Development (BX000226, BX005112, and BX004209), and National Institute of Allergy and Infectious Diseases (AI122574) (AAV).

Author information

Authors and Affiliations

Authors

Contributions

A.M, A.A.V., H.O., R.B. and G.B. wrote the manuscript text. S.M. and G.B. created the Figures. All authors reviewed and edited the manuscript.

Corresponding authors

Correspondence to Agata Matejuk or Arthur A. Vandenbark.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Competing interests

The authors declare the following financial interests/personal relationships which may be considered as potential competing interests: Drs. Vandenbark and Offner, and OHSU, have a significant financial interest in Artielle ImmunoTherapeutics, Inc., a company that may have a commercial interest in the results of this research and technology. This potential conflict of interest has been reviewed and managed by the OHSU and VA Portland Health Care System Conflict of Interest in Research Committees.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Matejuk, A., Benedek, G., Bucala, R. et al. MIF contribution to progressive brain diseases. J Neuroinflammation 21, 8 (2024). https://doi.org/10.1186/s12974-023-02993-6

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s12974-023-02993-6

Keywords