CXCL10, also known as interferon gamma-induced protein 10 (IP-10), is a small cytokine belonging to the CXC chemokine family that functions as a chemoattractant for immune cells during inflammatory responses.[1] Encoded by the CXCL10gene located on humanchromosome 4q21, it is a 98-amino-acid precursor protein processed into a mature 10-kDa secreted form featuring a characteristic chemokine structure with a three-stranded β-sheet and an overlying α-helix stabilized by disulfide bonds.[1] Primarily induced by interferon gamma (IFN-γ) and sometimes lipopolysaccharide (LPS), CXCL10 is expressed by diverse cell types including monocytes, endothelial cells, fibroblasts, keratinocytes, and epithelial cells in response to immune stimuli.[2]CXCL10 exerts its effects mainly by binding to the G protein-coupled receptorCXCR3, which exists in splice variants (CXCR3A and CXCR3B) and is predominantly expressed on activated Th1 lymphocytes, natural killer (NK) cells, and dendritic cells.[1] This interaction promotes chemotaxis, facilitating the directed migration of these cells to sites of infection, inflammation, or tissue damage, while also modulating angiogenesis inhibition and inducing apoptosis in certain cell types.[2] Additionally, CXCL10 can bind to the receptor CCR3 at lower affinity, potentially acting as an antagonist and influencing eosinophil recruitment in allergic contexts.[3]Beyond its role in acute immune defense against viral and bacterial pathogens, CXCL10 contributes to chronic inflammatory processes and is implicated in numerous diseases.[4] In autoimmunity, elevated CXCL10 levels drive T cell infiltration in conditions such as rheumatoid arthritis, multiple sclerosis, and type 1 diabetes; in cancer, it exhibits dual effects by enhancing anti-tumor immunity through NK and T cell recruitment while also promoting tumor progression via angiogenesis modulation in some contexts.[5] Its involvement in transplant rejection and fibrotic diseases further underscores the CXCL10-CXCR3 axis as a potential therapeutic target for modulating pathological inflammation.[2]
Molecular Characteristics
Gene
The CXCL10 gene, officially known as C-X-C motif chemokine ligand 10, was first cloned and characterized in 1985 from U937 human monocytic cells treated with recombinant interferon-gamma (IFN-γ), where it was identified as an early-response gene induced by this cytokine.[6] This discovery highlighted its role in interferon-responsive transcription, with the full-length cDNA sequence revealing homology to platelet-derived proteins.In humans, the CXCL10 gene is located on the long arm of chromosome 4 at position 4q21.1, spanning approximately 2.4 kb from nucleotide 76,021,118 to 76,023,497 (GRCh38.p14 assembly).[7] It resides within a compact CXC chemokinegene cluster on this chromosomal region, alongside the closely related genes CXCL9 (MIG) and CXCL11 (I-TAC), which share structural and functional similarities as IFN-inducible chemokines.[8] The gene comprises four exons interrupted by three introns, with the coding sequence distributed across these exons to produce a 98-amino-acid precursor protein that undergoes signal peptide cleavage to yield the mature form.[9][7]Transcription of the CXCL10 gene is primarily regulated by its promoter, which includes multiple interferon-stimulated response elements (ISREs) that bind interferon regulatory factors (IRFs) in response to IFN-γ stimulation, enabling rapid and robust induction during immune activation.[10] This architectural feature ensures tight control over expression in various cell types under inflammatory conditions.The CXCL10 gene exhibits strong evolutionary conservation across mammals, with orthologs present in species such as the mouse (Cxcl10, located on chromosome 5), where the coding regions display approximately 67% amino acid sequence identity, preserving key functional motifs.[11][12][13] This conservation underscores the gene's fundamental role in innate and adaptive immunity across vertebrates.
Protein Structure
The CXCL10 protein is initially synthesized as a precursor polypeptide of 98 amino acids, with a calculated molecular weight of approximately 10.9 kDa.[9] Following proteolytic cleavage of the N-terminal signal peptide (residues 1–21), the mature secreted form consists of 77 amino acids (residues 22–98) and exhibits a molecular weight of about 8.7 kDa.[14]The primary amino acid sequence of mature CXCL10 includes the defining CXC motif, characterized by two conserved cysteine residues separated by a single amino acid (typically X representing any residue), which facilitates proper folding.[3] Notably, it lacks the Glu-Leu-Arg (ELR) tripeptide motif immediately preceding the CXC sequence, a feature that differentiates it from ELR-positive CXC chemokines known for their pro-angiogenic properties and instead aligns it with anti-angiogenic variants.[15]In terms of secondary and tertiary structure, CXCL10 adopts the canonical chemokine fold observed in the CXC subfamily, featuring three antiparallel β-strands forming a β-sheet that is overlaid by a C-terminal α-helix, along with a flexible N-terminal unstructured loop essential for receptor engagement.[3] The protein predominantly exists as a monomer in solution but can form dimers through β-sheet interactions between monomers, with crystal structures revealing higher-order oligomeric assemblies such as tetramers under certain conditions.[16]Post-translational modifications of CXCL10 are limited but critical for stability; the core structure is maintained by two conserved intramolecular disulfide bonds linking Cys9 to Cys36 and Cys11 to Cys53 (mature numbering).[3] Additionally, CXCL10 undergoes C-terminal truncations as a natural post-translational modification, contributing to structural and functional heterogeneity.[17] Although potential N-glycosylation sites are present (e.g., at Asn62), the protein is primarily non-glycosylated in its native form, preserving its compact size and activity.[18]Physicochemical properties of CXCL10 include a basic isoelectric point (pI ≈ 8.7), contributing to its positive charge at physiological pH and enhancing solubility in aqueous environments without aggregation.[14]
Expression and Regulation
Cellular Sources
CXCL10 is primarily produced by immune cells such as monocytes, macrophages, and dendritic cells, as well as non-immune cells including endothelial cells, fibroblasts, and keratinocytes under proinflammatory conditions.[1] Monocytes and macrophages represent key leukocyte sources, with enhanced expression observed in these cell types during activation. Dendritic cells contribute to CXCL10 secretion in response to inflammatory stimuli, while endothelial cells and fibroblasts produce it in vascular and connective tissues.[1]Keratinocytes, particularly in the skin, serve as a prominent epithelial source.[1]In terms of tissue distribution, CXCL10 exhibits low basal expression levels in most healthy adult organs, reflecting its role as an inducible chemokine.[19] It is prominently upregulated in inflamed tissues, including the lung, liver, skin, and brain, where local inflammation drives production by resident cells.[1] For instance, in the lung and liver, expression increases during inflammatory responses, while in the skin and brain, it localizes to epithelial and glial cells under stress.[19][1]Pathogen-induced expression of CXCL10 is notably upregulated in response to viral infections, such as SARS-CoV-2, particularly in epithelial cells of the respiratory tract.[20] In the central nervous system, neurons also express elevated levels during viral infections, contributing to local immune responses.[21]In the tumor microenvironment (TME) of solid tumors, CXCL10 is produced by cancer cells and stromal cells, including fibroblasts and immune infiltrates, in malignancies like melanoma and breast cancer.[22][23] Stromal cells in the TME, such as cancer-associated fibroblasts, often amplify CXCL10 secretion, influencing the local cellular milieu.[24]Developmentally, CXCL10 shows minimal expression in healthy adult tissues but higher levels during fetal development, where it supports immune priming in structures like the placenta and airway smooth muscle cells.[25][26] This pattern aids in establishing early immune surveillance at maternal-fetal interfaces.[26]
Regulatory Mechanisms
The expression of CXCL10 is primarily regulated at the transcriptional level through induction by interferon-gamma (IFN-γ), which activates the JAK-STAT pathway, leading to the binding of STAT1 homodimers to gamma-activated site (GAS) elements and the STAT1-IRF1 complex to interferon-stimulated response elements (ISRE) in the CXCL10 promoter. Additionally, interferon regulatory factor 1 (IRF1) transcriptionally up-regulates CXCL10 expression.[27][28][29] Secondary transcriptional activation occurs via tumor necrosis factor-alpha (TNF-α) and interleukin-1 beta (IL-1β), which stimulate NF-κB translocation and binding to NF-κB sites in the promoter, often synergizing with IFN-γ for enhanced expression.[30][31]Epigenetic modifications further modulate CXCL10 promoter accessibility, particularly in immune cells, where histone acetylation facilitates transcriptional activation by increasing chromatin openness, while DNA methylation at CpG islands represses expression by compacting chromatin structure.[32][33] For instance, inhibition of histone deacetylases or methyltransferases like EZH2 and G9a elevates histone acetylation at the CXCL10 promoter, promoting gene activation in contexts such as idiopathic pulmonary fibrosis.[32]Post-transcriptional control influences CXCL10 mRNA stability, with the RNA-binding protein HuR binding to AU-rich elements in the 3' untranslated region to prevent degradation and enhance longevity, thereby amplifying protein output during inflammatory responses.[34] Conversely, microRNAs act as negative regulators by targeting the CXCL10 transcript, reducing its stability and translation in a feedback mechanism that tempers excessive inflammation.At the protein level, proteolytic processing by dipeptidyl peptidase IV (DPP4, also known as CD26) cleaves the N-terminal dipeptide of mature CXCL10, yielding a truncated form (CXCL10(3-77)) that exhibits reduced chemotactic activity and can function as a competitive antagonist at CXCR3 receptors.[35][36] This truncation alters CXCL10's biological potency, shifting it from an agonist to an inhibitory variant in certain inflammatory milieus.Feedback loops involving CXCL10 and its receptor CXCR3 provide autocrine and paracrine regulation, where in some contexts, CXCR3 signaling on producing cells triggers CD26-mediated degradation of CXCL10, establishing a negative feedback that limits sustained production and prevents overactivation of immune responses.[37][1]
Biological Functions
Chemotactic Activity
CXCL10 primarily directs the migration of immune cells through concentration gradients formed at sites of inflammation, targeting activated T cells (including Th1 and CD8+ subsets), natural killer (NK) cells, eosinophils, and monocytes.[1][38] These cells express the receptor CXCR3, to which CXCL10 binds to initiate directional movement.[39]The chemotactic mechanism involves both haptotactic migration, where CXCL10 bound to glycosaminoglycans on endothelial and epithelial surfaces guides cells along substrates, and chemokinetic effects that increase random motility at higher concentrations.[1] This process is facilitated by integrin activation on migrating cells, enabling adhesion and traversal of endothelial barriers.[40] CXCL10 effectively induces these responses at low nanomolar concentrations.[41] The steepness of the CXCL10 gradient further influences cell speed and directionality, with steeper gradients promoting faster, more persistent migration toward the source.[42]In tissue-specific contexts, CXCL10 recruits these immune cells to inflamed sites such as atherosclerotic plaques, where it draws Th1 cells and monocytes to exacerbate or modulate plaque progression,[43] and to viral-infected epithelia, facilitating rapid infiltration during infections like those caused by SARS-CoV-2.[44] Beyond immune cells, CXCL10 exhibits limited attraction of endothelial cells, primarily modulating angiogenesis through inhibitory effects that reduce proliferation and tube formation.[1]
Immune Modulation
CXCL10 plays a pivotal role in T cell polarization by promoting Th1 responses through enhanced production of interferon-gamma (IFN-γ) and inhibition of Th2 and Th17 differentiation. It fosters a positive feedback loop with IFN-γ, where CXCL10 recruits Th1 cells to sites of inflammation, amplifying their activation and cytokine secretion, while suppressing Th2-associated cytokines such as IL-4 and IL-13. This skewing toward Th1 dominance is evident in infectious models, where CXCL10 enhances T cell responsiveness to IL-12, further boosting IFN-γ output and limiting alternative effector pathways.[45]In addition to T cell effects, CXCL10 activates natural killer (NK) cells and dendritic cells (DCs), elevating cytotoxicity and antigen presentation. CXCL10 recruits CXCR3-expressing NK cells to infected or inflamed tissues, increasing IFN-γ secretion and perforin/granzyme-mediated killing of target cells. Similarly, for DCs, CXCL10 promotes their maturation and migration, enhancing MHC class II expression and IL-12 production to prime Th1-biased adaptive responses.[46]CXCL10 contributes to anti-viral immunity by amplifying type I IFN responses and restricting viral replication within infected cells. It is upregulated by IFN-α/β, creating a feedbackmechanism that recruits effector cells and directly inhibits viral entry and propagation, as seen in infections like coxsackievirus and respiratory syncytial virus. This modulation strengthens innate antiviral barriers before adaptive immunity fully engages.[46]In contexts of immune tolerance, CXCL10 recruits regulatory T cells (Tregs) expressing CXCR3, helping to dampen excessive inflammation and maintain homeostasis.[47] By attracting these suppressive cells to inflamed sites, CXCL10 can mitigate overactive responses, particularly in chronic settings. However, CXCL10 exhibits dual effects in autoimmunity; while it may limit overall disease in some models, astroglial-derived CXCL10 exacerbates tissue damage in experimental autoimmune encephalomyelitis (EAE) by enhancing perivascular CD4+ T cell accumulation and acute demyelination.[48]
Receptor Interactions
CXCR3 Binding
CXCR3 is a seven-transmembrane G protein-coupled receptor (GPCR) predominantly expressed on activated T lymphocytes, natural killer cells, and other immune effectors, playing a key role in directing leukocyte trafficking during inflammation. It exists in three main isoforms generated by alternative splicing: CXCR3-A, the canonical form; CXCR3-B, which has an extended N-terminal domain; and CXCR3-alt, characterized by a truncated C-terminal domain due to exon skipping. These isoforms differ in their N- and/or C-terminal regions: CXCR3-A and CXCR3-B primarily in the N-terminus, with CXCR3-B having an extension, and CXCR3-alt featuring a truncated C-terminus, which influence ligand responsiveness and downstream effects.[49][50][51]CXCL10 exhibits high-affinity binding to CXCR3, with a dissociation constant (Kd) of approximately 0.2–0.3 nM, enabling efficient receptor engagement under physiological conditions. This interaction is mediated by the N-terminal domain of CXCL10, which inserts into the orthosteric binding pocket of CXCR3, forming key contacts with transmembrane (TM) helices, including hydrophobic interactions with residues in TM3 (e.g., Phe131, Phe135), TM6 (e.g., Tyr271), and TM7 (e.g., Ser304, Tyr308), as inferred from structural analogies with related ligands. The proximal N-terminus of CXCR3 is also essential for stable binding of CXCL10, involving tyrosine sulfation that enhances ligand recognition.[52][53][54]In terms of ligand specificity, CXCL10 selectively binds CXCR3 with high affinity, sharing this receptor with its fellow ELR-negative CXC chemokines CXCL9 and CXCL11, which collectively form the primary agonist set for CXCR3 activation; CXCL10 shows no significant affinity for other chemokine receptors such as CXCR4, but exhibits low-affinity binding to CCR3, potentially acting as an antagonist and influencing eosinophil recruitment in allergic contexts. Among these, CXCL11 displays the highest affinity, followed by CXCL10 with intermediate potency, and CXCL9 with the lowest.[3][53][3]Structural insights into the CXCL10–CXCR3 interaction derive from cryo-EM structures of CXCR3 bound to CXCL11 (at 3.0 Å resolution), which reveal a conserved binding mode where the chemokine's N-terminal residues penetrate the receptor's helical bundle, inducing conformational changes for activation; molecular dynamics simulations support a similar engagement for CXCL10. Additionally, crystal structures of CXCL10 demonstrate its ability to form dimers and tetramers via β-sheet interfaces, potentially modulating receptor dimerization and enhancing signaling efficiency, though monomeric forms predominate in solution.[53][55]The isoforms exhibit distinct functional profiles upon CXCL10 binding: CXCR3-A promotes chemotactic migration of immune cells, while CXCR3-B triggers anti-angiogenic responses in endothelial cells, highlighting isoform-specific contributions to CXCL10's pleiotropic effects. Binding of CXCL10 to these isoforms initiates G protein-mediated signaling for CXCR3-A, influencing cellular motility and vascular regulation.[56][49]
Other Receptors
In addition to CXCR3, CXCL10 binds with low affinity to the chemokine receptor CCR3, primarily expressed on eosinophils and Th2 cells. This interaction, with a Kd in the micromolar range, functions as an antagonist, inhibiting CCR3-mediated responses to eotaxin (CCL11) and thereby modulating allergic inflammation and eosinophil recruitment.[3]
Signaling Pathways
Upon binding to its receptor CXCR3, CXCL10 initiates intracellular signaling primarily through G-protein-coupled mechanisms. The receptor couples predominantly to Gαi proteins for the CXCR3-A isoform, which inhibit adenylate cyclase activity and thereby reduce intracellular cyclic AMP (cAMP) levels, contributing to chemotaxis and cell activation in immune cells.[57] Recent studies indicate that CXCR3-B does not efficiently couple to G proteins such as Gαi or Gαs, but instead engages alternative pathways like β-arrestin recruitment to mediate its anti-proliferative and anti-angiogenic effects.[51]Downstream of G-protein activation in CXCR3-A, several effector pathways are engaged to promote cellular responses. The PI3K-Akt pathway is activated, enhancing cell survival and migration through phosphorylation of Akt, which is critical for leukocyte recruitment.[58] Similarly, the MAPK/ERK cascade is stimulated, driving proliferation and gene expression changes via ERK1/2 phosphorylation.[58] Additionally, phospholipase C β (PLCβ) activation leads to inositol trisphosphate (IP3) production and subsequent calcium mobilization from intracellular stores, facilitating cytoskeletal rearrangements essential for directed motility.[58]CXCL10-CXCR3 signaling integrates with interferon-γ (IFN-γ) pathways, where receptor engagement promotes STAT1 phosphorylation, amplifying Th1-biased immune responses in synergy with IFN-γ-induced effects.[59] Signal termination occurs via β-arrestin recruitment to the phosphorylated receptor, which uncouples G-proteins, promotes clathrin-mediated internalization, and desensitizes the pathway to prevent overstimulation.[60]In non-immune cells, such as endothelial or epithelial cells, CXCL10-CXCR3 activation can engage alternative routes, including p38 MAPK signaling, which mediates stress responses and growth inhibition, particularly through the CXCR3-B isoform.[51]
Clinical Relevance
Role in Diseases
CXCL10 exhibits a dual role in diseasepathogenesis, acting as both a protector against infections through immune cell recruitment and a contributor to excessive inflammation in chronic conditions, thereby influencing outcomes in infectious, autoimmune, neoplastic, neurodegenerative, and cardiovascular disorders.[61]In infectious diseases, CXCL10 enhances viral clearance by promoting the migration and activation of CXCR3-expressing leukocytes, such as T cells and natural killer cells, which help control pathogen spread in infections like SARS-CoV-2 and HIV.[62] For instance, in SARS-CoV-2 infection, CXCL10 upregulation supports antiviral defenses but can lead to immunopathology, including cytokine storms that exacerbate acute respiratory distress syndrome.[62] Similarly, in HIV, elevated CXCL10 levels drive immune activation and T cell trafficking to infected sites, aiding containment while contributing to chronic inflammation and disease progression.[62]In autoimmune disorders, CXCL10 drives pathogenesis by facilitating the infiltration of Th1-polarized T cells and other leukocytes into target tissues, amplifying inflammation and tissue damage.[38] In rheumatoid arthritis, increased synovial CXCL10 expression recruits CXCR3+ T cells and induces RANKL-mediated bone erosion, perpetuating joint destruction.[38] In multiple sclerosis, CXCL10 promotes T cell migration across the blood-brain barrier, leading to demyelination and neuronal injury in the central nervous system.[38] Likewise, in type 1 diabetes, elevated CXCL10 in pancreatic islets attracts autoreactive T cells, accelerating β-cell destruction and hyperglycemia onset.[38]In cancer, CXCL10 displays context-dependent effects, either bolstering anti-tumor immunity or fostering tumor progression.[63] In immunogenic tumors like melanoma, it recruits CD8+ T cells and NK cells via CXCR3, activating JAK/STAT pathways to enhance Th1 responses and improve outcomes with PD-1 inhibitors.[63] Conversely, in breast cancer, particularly triple-negative subtypes, CXCL10 promotes angiogenesis and metastasis by stimulating PI3K/Akt signaling in tumor cells and recruiting immunosuppressive myeloid-derived suppressor cells to the microenvironment.[63]In neuroinflammatory conditions, CXCL10 exacerbates neurodegeneration through microglial activation and immune cell influx into the brain.[64] In Alzheimer's disease, elevated CXCL10 in cerebrospinal fluid co-localizes with amyloid-β plaques, impairing microglial phagocytosis and promoting astrocytic inflammation, as evidenced by reduced plaque burden in CXCR3-deficient models.[64] In Parkinson's disease, CXCL10 upregulation in response to dopaminergic neuron loss activates microglia and correlates with cognitive decline, potentially disrupting the blood-brain barrier to allow peripheral immune infiltration.[64]In cardiovascular diseases, CXCL10 contributes to atherosclerosis by attracting monocytes to arterial plaques, intensifying local inflammation and lesion instability.[65] Expressed by endothelial cells and smooth muscle cells under shear stress, it binds CXCR3 on monocytes to promote their adhesion and infiltration, as shown in reduced plaque sizes in CXCL10-deficient mouse models.[65] Elevated circulating CXCL10 levels thus serve as a marker of plaque progression in human coronary artery disease.[65]
Biomarkers
CXCL10 serves as a key biomarker for detecting and monitoring neuroinflammation and various inflammatory conditions through its measurement in biological fluids and tissues. Levels of CXCL10 protein are routinely quantified using enzyme-linked immunosorbent assay (ELISA) in serum, plasma, and cerebrospinal fluid (CSF), providing sensitive detection down to the picogram per milliliter range. Quantitative polymerase chain reaction (qPCR) is employed to evaluate CXCL10 mRNA expression in tissue biopsies, offering insights into local production during disease states. For example, in HTLV-1-associated myelopathy, a CSF cutoff value exceeding 110 pg/mL has demonstrated high diagnostic accuracy for neuroinflammation, with 97% sensitivity and 96% specificity.[66][67][68][69]Elevated CXCL10 levels exhibit diagnostic utility across acute infections, autoimmune disorders, and malignancies. In severe COVID-19 cases, serum CXCL10 concentrations at hospital admission robustly predict disease progression and outcomes, often rising significantly in patients requiring intensive care. For autoimmune flares, such as relapses in relapsing-remitting multiple sclerosis (RRMS), increased CSF and serum CXCL10 correlates with disease activity and short-term risk of progression, serving as an indicator of active inflammation. In cancers like papillary thyroid carcinoma, higher tissue and circulating CXCL10 expression distinguishes malignant from benign lesions and reflects immune infiltration.[70][71][72]Prognostically, elevated CXCL10 levels forecast adverse outcomes in several conditions. High CSF CXCL10 in people living with HIV is associated with HIV-associated neurocognitive disorder (HAND), particularly in antiretroviral therapy-naïve individuals, where it signals greater cognitive impairment severity and poorer neurological prognosis. In multisystem inflammatory syndrome in children (MIS-C), plasma CXCL10 strongly correlates with left ventricular dysfunction, enabling early identification of cardiac complications. For kidney transplant recipients, urinary CXCL10 above established thresholds predicts acute rejection episodes and long-term graft dysfunction, with levels correlating to inflammatory burden and eGFR decline.[73][74][75]To enhance specificity, CXCL10 is frequently combined with related chemokines CXCL9 and CXCL11, improving predictive accuracy for treatment responses. In triple-negative breast cancer (TNBC) patients receiving pembrolizumabimmunotherapy, elevated baseline levels of CXCL9, CXCL10, and CXCL11 indicate a favorable immune-enriched tumor microenvironment and longer survival post-treatment. This trio of biomarkers outperforms individual measures in stratifying responders versus non-responders.Recent 2025 validations underscore CXCL10's clinical relevance in emerging contexts. In systemic chronic active Epstein-Barr virus (sCAEBV) disease, plasma CXCL10 levels reflect disease activity and treatment efficacy, positioning it as a non-invasive prognostic tool for monitoring progression and response to interventions. Similarly, in hepatocellular carcinoma (HCC), serum and tissue CXCL10 expression serves as an independent prognostic marker, with higher levels linked to worse overall survival and potential for guiding targeted therapies.[76]
Therapeutic Potential
CXCL10 modulation offers therapeutic potential in inflammatory and neoplastic diseases through antagonists that inhibit excessive signaling and agonists that enhance immune recruitment. In autoimmune conditions, blocking the CXCL10/CXCR3 axis reduces pathogenic T-cell infiltration and inflammation. For instance, the CXCR3 antagonist AMG487 has demonstrated efficacy in preclinical models of rheumatoid arthritis by suppressing joint inflammation and shifting the Th17/Treg balance, and it advanced to clinical trials for psoriasis and rheumatoid arthritis, though development was discontinued due to limited efficacy. Similarly, neutralizing antibodies such as MDX-1100 (eldelumab), a fully human anti-CXCL10 monoclonal antibody, were evaluated in a phase II trial for rheumatoid arthritis, demonstrating significant improvements in American College of Rheumatology (ACR) 20 response rates (54% vs. 17% for placebo; P=0.0024) when combined with methotrexate, though further development was discontinued. Another anti-CXCL10 antibody, NI-0801, has been investigated for its potential in blocking CXCL10-mediated inflammation in autoimmune settings.In cancer and antiviral contexts, strategies to augment CXCL10 activity aim to bolster antitumor and antiviral immunity by promoting effector T-cell and NK-cell trafficking. IFN-γ adjuvants upregulate CXCL10 expression to enhance immune responses in viral infections, as seen in studies where IFN-γ-induced CXCL10 production correlates with improved T-cell infiltration and viral clearance. For oncology, gene therapy approaches like AAV-CXCL10 vectors or oncolytic adenoviruses armed with CXCL10 have shown promise in preclinical glioblastoma models by increasing tumor-infiltrating lymphocytes and synergizing with radiotherapy or checkpoint inhibitors to improve survival. Ongoing studies explore CXCL10 upregulation in glioblastomaimmunotherapy, where macrophage-derived CXCL10 enhances responses to PD-1 blockade, and phase I/II trials are evaluating combinations that boost CXCL10 to overcome immunosuppressive microenvironments. DPP4 inhibitors, such as sitagliptin, preserve full-length agonist CXCL10 by preventing its truncation to an inactive form, thereby enhancing T-cell recruitment in tumors and improving outcomes in preclinical cancer models when combined with immunotherapy.Emerging therapies address CXCL10's processing and dual roles to optimize efficacy. DPP4-resistant CXCL10-Fc fusions, developed in 2025, resist enzymatic cleavage to maintain full agonistic activity and show potent antitumor effects in mouse models by sustaining IFN-γ and IL-2 production, positioning them as candidates for next-generation cancer immunotherapies resistant to enzymatic degradation.[77] Bispecific antibodies targeting CXCL10 alongside other cytokines, like TNF-α/CXCL10 constructs, ameliorate inflammation in arthritis models by dual blockade, with potential extensions to neuroinflammatory conditions where CXCL10/CXCR3 drives microglial activation. Phase II trials for psoriasis have tested CXCR3 blockade, including related inhibitors, to curb CXCL10-driven skin inflammation, while glioblastoma studies investigate CXCL10 enhancers in combination regimens. Challenges include CXCL10's context-dependent effects, where excessive agonism may promote pro-tumor angiogenesis in some cancers, necessitating biomarker-guided selection to balance anti-inflammatory and immunostimulatory outcomes without exacerbating pathology.