Next Article in Journal
Intranasally Administered Insulin as Neuromodulating Factor and Medication in Treatment of Neuropsychiatric Disorders—Current Findings from Clinical Trials
Previous Article in Journal
The Impact of Providing Pharmaceutical Care on Work Satisfaction of Pharmacists in Poland—A Preliminary Study
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Deciphering Drug Repurposing Strategies: Antiviral Properties of Candidate Agents Against the Mpox Virus

by
Aganze Gloire-Aimé Mushebenge
* and
David Ditaba Mphuthi
Department of Health Studies, College of Human Sciences, Muckleneuk Campus, University of South Africa, Pretoria 0027, South Africa
*
Author to whom correspondence should be addressed.
Sci. Pharm. 2025, 93(4), 51; https://doi.org/10.3390/scipharm93040051
Submission received: 26 August 2025 / Revised: 13 October 2025 / Accepted: 14 October 2025 / Published: 17 October 2025

Abstract

Monkeypox (Mpox) has re-emerged as a global public health threat, with recent outbreaks linked to novel mutations that enhance viral transmissibility and immune evasion. The Mpox virus (MPXV), a double-stranded deoxyribonucleic acid (DNA) orthopoxvirus, shares high structural and enzymatic similarity with the variola virus, underscoring the need for urgent therapeutic interventions. While conventional antiviral development is time-intensive and costly, drug repurposing offers a rapid and cost-effective strategy by leveraging the established safety and pharmacological profiles of existing medications. This is a narrative integrative review synthesizing published evidence on drug repurposing strategies against MPXV. To address these issues, this review explores MPXV molecular targets critical for genome replication, transcription, and viral assembly, highlighting how the Food and Drug Administration (FDA)-approved antivirals (cidofovir, tecovirimat), antibiotics (minocycline, nitroxoline), antimalarials (atovaquone, mefloquine), immunomodulators (infliximab, adalimumab), and chemotherapeutics (doxorubicin) have demonstrated inhibitory activity against the virus using computational or experimental approaches. This review further evaluates advances in computational methodologies that have accelerated the identification of host-directed and viral-directed therapeutic candidates. Nonetheless, translational challenges persist, including pharmacokinetic limitations, toxicity concerns, and the limited efficacy of current antivirals such as tecovirimat in severe Mpox cases. Future research should integrate computational predictions with high-throughput screening, organ-on-chip technologies, and clinical pipelines, while using real-time genomic surveillance to track viral evolution. These strategies establish a scalable and sustainable framework for the MPXV drug discovery.

1. Introduction

The re-emergence of Mpox, caused by the Mpox virus (MPXV), has escalated from an endemic disease in Central and West Africa to a global health concern [1,2,3]. With a wide geographic dispersion in more than 116 nations and new forms of transmission, such as sexual contact, the outbreak in 2022 brought it to the public’s attention [4]. The cessation of smallpox vaccination programs increased susceptibility by creating an immunity gap in the community. Mpox has historically been associated with zoonotic spillover from animal reservoirs, such as rats [5,6]. Environmental change, human behaviour, and globalization drive the re-emergence of Mpox [7]. This underscores the need for rapid and effective countermeasures.
Because of the advent of virulent and genetically varied clades such as clade I, Mpox poses distinct hurdles in diagnosis, treatment, and containment, despite its clinical similarity to smallpox. The global health threat posed by Mpox led the World Health Organization (WHO) to declare it a Public Health Emergency of International Concern (PHEIC), while inequitable healthcare access, disinformation, and stigma continue to delay treatment and create major public health barriers (https://www.who.int/news-room/fact-sheets/detail/Mpox, accessed on 4 October 2025) [8]. In addition, most outbreaks arise from viruses without approved medication. Previously, virus-specific drugs followed the traditional discovery approach, but the re-emergence of Mpox underscores the need for broad-spectrum antivirals.
These challenges require specialized therapies to address the evolving epidemiological landscape. Although vaccination remains crucial, current smallpox vaccines such as JYNNEOS offer limited protection against Mpox and face challenges in coverage, distribution, and public acceptance in endemic regions [9,10]. Furthermore, the rapid evolution of MPXV calls for creating polyvalent vaccinations that target several circulating clades [9]. Novel approaches to improve vaccination tactics have been made possible by developments in nanotechnology-based delivery systems, which have demonstrated promise in enhancing vaccine durability and bioavailability [11]. However, the emphasis needs to change to therapeutic options because immunization might not be enough to stop the pandemic.
One crucial tactic for meeting the pressing need for efficient Mpox therapies is drug repurposing (DR). De novo drug discovery is more expensive and time-consuming than using current pharmacological resources [12]. Finding possible antiviral drugs that target the MPXV replication and immune evasion pathways has been made possible in large part by computational methods, including molecular docking and dynamics simulations [13]. The efficacy of repurposed drugs such as tecovirimat (TPOXX®), cidofovir, and brincidofovir varies, while host-targeted therapies such as mycophenolate mofetil (MMF) and IMP-1088 are under investigation for their synergistic effects with direct-acting antivirals [14,15,16].
Relevant comprehensive literature was identified through structured searches in PubMed, Scopus, ScienceDirect, Web of Science, Google Scholar and preprints, among others, and publisher archives covering publications from 2006 to October 2025. Keyword searches combined ‘Mpox virus’, ‘monkeypox’, ‘drug repurposing’, ‘antiviral’, and ‘in silico docking.’ Studies were screened for relevance to MPXV antiviral activity, computational predictions, or mechanistic insights into viral inhibition. Review articles, molecular docking analyses, and experimental studies were all considered for inclusion, while commentaries and methodological papers were omitted. This is a narrative integrative review that highlights the milestones, developments, and prospects of repurposed medications in the fight against Mpox, with particular emphasis on integrating preclinical and clinical findings into feasible treatment strategies. This underscores the importance of cutting-edge technologies, especially computational approaches, in improving therapeutic outcomes. Viral and host protein targets have been evaluated with diverse drugs in silico, in vitro, and in vivo, with several showing promising activity for potential MPXV repurposing. However, Mpox eradication cannot rely solely on the conventional drug discovery pipeline, despite the remarkable advances in computational and experimental methods in recent decades. DR has emerged as a successful strategy for identifying candidate antivirals against MPXV, a member of the Orthopoxvirus genus. This review further advocates a coordinated global response via a One Health framework integrating human, animal, and environmental health. Evidence underscores the need for interdisciplinary collaboration to accelerate antiviral development, mitigate future outbreaks, and strengthen global health resilience.

2. Mpox Virus Biology and Pathogenesis

2.1. Genomic Characteristics and the Viral Lifecycle

MPXV, a zoonotic orthopoxvirus, has drawn significant attention due to its rapid global spread and designation as a PHEIC [7,17]. The conserved and specialized genes that make up the viral genome allow MPXV to infect a variety of species, withstand host immune responses, and continue to spread from person to person [18]. The deletion of functional OPG041 (K3L) and OPG065 (E3L) genes distinguishes MPXV from other orthopoxviruses and reflects its strategies to evade host antiviral defenses. These modifications, which include decreased the double-stranded ribonucleic acid (dsRNA) production and altered interferon (IFN) responses, highlight the virus’s capacity to infect humans and small rodents [19]. The evolution of MPXV toward improved human adaptation is highlighted by the formation of Clade I and Clade II MPXV lineages, each of which is distinguished by unique virulence factors and transmission dynamics [13]. Continuous mutation and selection further shape the virus’s pathogenic capacity, driven by genomic reduction and codon optimization via the human apolipoprotein B mRNA editing enzyme, catalytic polypeptide-like (APOBEC) editing [13,20].
Ankyrin-containing and Bcl-2-like proteins, which are essential for immune evasion and host range, form part of MPXV’s genomic toolkit and are closely linked to its lifecycle and pathogenesis [13]. Comparative studies have shown that MPXV contains distinct virulence genes, including OPG174 (A44L), which facilitates persistent infections by enabling immunosuppression through the synthesis of glucocorticoids [19]. There are still significant gaps in the knowledge of MPXV biology that need to be filled to fully comprehend the molecular mechanisms behind viral replication and host interactions. For the development of antiviral medications, computational methods have found conserved viral targets such as DNA polymerase and thymidine kinase [13,21]. However, mutations in therapeutic targets highlight the need for new treatment approaches that focus on low-mutation viral areas and raise concerns about drug resistance [22]. In-depth functional research and ongoing genetic surveillance are crucial for creating potent antiviral medications, anticipating future outbreaks, and reducing the harm MPXV poses to the world health.

2.2. Key Molecular Targets for Therapeutic Intervention

Finding and confirming molecular targets within the MPXV that may be viable candidates for antiviral development is an essential objective in this effort. Using genomic and subtractive proteomics, researchers identified the highly druggable core proteins A20R, I7L, Top1B, and VETFS conserved across MPXV strains and non-homologous to human proteins [23]. Potential inhibitors have been identified using high-throughput virtual screening of already available medications [13,23]. Molecular dynamics (MD) simulations have confirmed these inhibitors’ interactions. These discoveries accelerate treatment development by enabling the repurposing of existing compounds against MPXV, as shown in Figure 1 [24].
Other important viral proteins, including DNA polymerase, Topoisomerase 1, D9 decapping enzyme, and thymidylate kinase, have also been investigated as potential therapeutic targets [25]. Functional and structural studies have enhanced our understanding of the roles these proteins play in immune evasion and viral replication. Efforts have mapped MPXV–host protein interactions to identify key viral proteins, such as OPG054, OPG084, and OPG190, as therapeutic targets [26]. Moreover, broad-spectrum antiviral drugs targeted host proteins implicated in immunological pathways, including the most interacting MPXV-interacting proteins [26]. The combination of experimental data and computational tools emphasizes how crucial a systems biology approach is to solve the problems caused by the MPXV.
The rapid genomic evolution of MPXV and the emergence of mutant strains resistant to existing treatments underscore the urgent need for alternative antiviral strategies. Advances in artificial intelligence and bioinformatics have facilitated the identification of immunogenic antigens and druggable targets, enabling the development of novel small-molecule inhibitors and multi-epitope vaccines [27]. Focusing on key viral and host biological targets, coupled with an enhanced understanding of virus–host interactions, will enable the development of more effective therapies. Supported by genomic surveillance, these initiatives are essential to mitigate the impact of MPXV outbreaks and strengthen preparedness for future public health threats.

2.3. Host–Pathogen Interactions

MPXV exhibits complex host–pathogen interactions that influence disease severity and progression (https://www.who.int/news-room/fact-sheets/detail/Mpox, accessed on 4 October 2025). The virus rapidly replicates in keratinocytes, fibroblasts, and antigen-presenting cells following infection via mucosal surfaces or skin breaks [22,28]. Lesions are dispersed based on lymphatic propagation, resulting in systemic viremia that affects vital organs such as the liver, spleen, and lungs. This permits systemic dispersion through both hematogenous and lymphatic routes [29,30]. With CD4+ and CD8+ T cells driving the removal of infected macrophages and the generation of cytokines such as IFN-γ and TNF, cellular immune responses are essential for viral control [31]. Nevertheless, MPXV has evolved defenses against immunological reactions, such as inhibiting T-cell activation [13]. Studies on human skin organoids revealed productive keratinocyte replication and multiple phases of intracellular virus assembly, highlighting the virus’s affinity for skin epithelial cells and the potential for severe localized or systemic symptoms influenced by host and viral factors [32].
The pathophysiology of MPXV has been further clarified by non-human primate models, which show that inoculation routes such as aerosol or intravenous exposure affect the course of the disease [33]. Aerosolized MPXV closely resembles human Mpox pathophysiology in that it produces severe bronchopneumonia and systemic dissemination [13]. Important genes that improve viral transmissibility and immune evasion have been identified through molecular research, including OPG027 and D1L [13]. Computational studies of host–pathogen interactions have identified key host proteins and signalling pathways, including AKT, STAT3, and NF-κB, as potential therapeutic targets in MPXV infection [34]. There is promise for repurposed treatments because promising FDA-approved drugs have demonstrated potential in targeting these pathways [34]. While antiviral therapies target viral envelope proteins and DNA polymerases, their clinical efficacy in real-world settings has been inconsistent, especially against Clade I strains with severe disease manifestations [35]. Furthermore, the MPXV genome shows potential for rapid adaptive mutations, particularly under antiviral pressure, which could compromise long-term efficacy [13]. In addition, MPXV encodes immunomodulatory proteins that allow it to evade host immune responses, limiting the therapeutic window of antivirals [13].

3. The Rationale for Drug Repurposing in Mpox Management

3.1. Advantages of Repurposed Drugs in Addressing MPXV

High MPXV mutability and the emergence of drug-resistant strains hinder the development of traditional antiviral drugs [24]. Repurposing drugs has become a crucial strategy in the fight against new viral illnesses. This approach, which is faster and inexpensive than traditional drug development, involves identifying new therapeutic uses for already approved medications. Advances in computational techniques have greatly facilitated this process by enabling the analysis of gene expression patterns, host–pathogen interactions, and protein–protein interaction (PPI) networks [36]. PPIs during MPXV infections may be predicted with excellent accuracy using ensemble feature-based deep learning models [37]. From FDA-approved databases, this method, which is supplemented by molecular docking and dynamics simulations, has found possible therapeutic targets and repurposed medications such as cannabidiol and fostamatinib [37]. These results highlight how DR works to combat the pressing demand for antiviral treatments against Mpox and related infectious illnesses.
Drug repurposing leverages approved drugs with established safety profiles, reducing the time needed for clinical application. A literature review identified 160 human-host proteins associated with MPXV and related viruses, highlighting 15 key targets via network-based drug discovery. Approved drugs such as baricitinib and adalimumab have emerged as promising repurposing candidates [38]. These outcomes shed light on the molecular processes of MPXV and demonstrate how repurposed medications may be used to address the dual problems of treatment resistance and the virus’s evolving nature.
In addition to improving Mpox treatment, combining computational and experimental methods in drug repurposing establishes a standard approach for tackling emerging viruses. Novel treatments have been made possible by insights from molecular diagnostics and therapeutic target identification, such as VP39 and thymidine kinase proteins [13]. They have also strengthened the arsenal against Mpox by developing nanomedicine, using herbal therapies, and creating contemporary diagnostic platforms such as the clustered regularly interspaced short palindromic repeats (CRISPR)-based techniques [39]. These initiatives underscore the critical role of DR in public health, enabling rapid and effective responses to emerging viral threats. By integrating advanced technologies and leveraging existing resources, drug repurposing provides a robust framework to mitigate the global impact of viral epidemics.

3.2. Historical Success of Drug Repurposing in Viral Outbreaks

The historical effectiveness of therapeutic repurposing in controlling viral outbreaks provides a promising way to combat the growing threat of Mpox. Because poxviruses share evolutionarily conserved proteins and functional traits, they allow effective cross-species treatments. [33]. The smallpox vaccine, which was created for a similar orthopoxvirus, showed 85% effectiveness in preventing Mpox infection [40]. Baricitinib, initially an anti-rheumatoid arthritis agent, was successfully repurposed for the coronavirus disease of 2019 (COVID-19) treatment, reducing mortality by 38% in hospitalized patients. Remdesivir, initially developed for Ebola, showed moderate success against the Severe acute respiratory syndrome coronavirus 2 but with variable clinical outcomes [41]. Favipiravir, another repurposed antiviral for influenza, displayed inconsistent efficacy in the COVID-19 trials. Similarly, tecovirimat, originally developed to combat smallpox, has been used to treat Mpox cases in Europe, demonstrating the value of repurposing already approved antiviral medications [42,43]. Thus, the DR strategy minimizes the time and expense involved in developing new therapies while highlighting the benefits of repurposing medications that have demonstrated efficacy against the genus Orthopoxvirus. However, repurposing success rates remain low, with only ~5% of antiviral candidates advancing to clinical approval [44]. Factors such as differences in viral replication mechanisms, host–pathogen interactions, and pharmacokinetics often hinder successful translation.
Computational and experimental approaches targeting viral and host proteins have identified promising candidates, including brincidofovir and cidofovir, with the potential to reduce Mpox infections [24,45]. The use of medication repurposing for Mpox demonstrates both its promise and the necessity of investigating other therapeutic approaches, considering the shortcomings of the available treatments. Repurposed medications, including tecovirimat and nucleoside analogues, serve as first-line defenses, despite the WHO not recommending mass vaccination and the lack of antivirals specifically approved for Mpox [24]. Developments in structural biology have helped us to rationally select and optimize nucleoside analogs by shedding light on the distinctive characteristics of viral DNA polymerases. These focused strategies highlight the wider significance of understanding viral protein structures and functions, enabling the development of effective and selective antiviral treatments [46]. DR is a crucial and flexible tactic to counteract Mpox and other new viral threats, given its ongoing development and potential for cross-species transmission.

3.3. Criteria for Selecting Candidate Drugs

Given the absence of specific antivirals or vaccines and the rapid global spread to 82 non-endemic countries, selecting suitable therapeutic options has become a crucial task. This challenge stems from several factors: declining population immunity after smallpox eradication, facilitating global susceptibility; international travel and social gatherings, enabling undetected transmission; and underestimation of asymptomatic or subclinical cases, hindering early detection and intervention [47,48,49]. Computational techniques such as subtractive proteomics and genomics have been used to identify possible therapeutic targets within the MPXV proteome [23]. Four highly druggable targets were identified from 69 conserved proteins found in 125 publicly accessible MPXV genomes [23]. Three frequent inhibitors, batefenterol, burixafor, and eluxadoline, that showed considerable binding affinities were found by high-throughput virtual screening of 5893 licensed and experimental medications [23]. MD simulations confirmed stable and favourable binding patterns for these inhibitors, identifying them as strong candidates for repurposing. This approach highlights the value of computational methods in prioritizing compounds for experimental validation. Advanced computational tools further facilitate the prediction of new inhibitors targeting key MPXV proteins, such as the cysteine proteinase, leading to the discovery of molecules such as CHEMBL32926 and CHEMBL4861364 with high binding affinity and drug-like properties [46]. Strong protein-ligand interactions were observed by researchers using ensemble docking and MD simulations, which were corroborated by advantageous binding free energy predictions [50]. The therapeutic landscape was further expanded by the discovery of analogues of these drugs with comparable characteristics that were anticipated to have potent inhibitory efficacy [46,50].

4. Computational Approaches in Drug Repurposing

4.1. Molecular Docking and Dynamics Simulations

Molecular docking predicts the preferred orientation of a ligand bound to a target macromolecule, such as a viral protein, and estimates the binding affinity. It provides a cost-effective primary screening method [51,52]. However, docking relies on rigid receptor structures and may not fully account for protein flexibility or solvation effects, which are critical in biological systems. MD simulations complement docking by modeling atomic-level movements over time, allowing exploration of binding stability and conformational changes under near-physiological conditions [52,53]. Despite their strengths, MD simulations are computationally intensive and depend on the accuracy of the applied force fields. The process depicted in Figure 2 below emerged as a revolutionary strategy in tackling the worldwide issue. In silico, drug design follows two main routes. Route B encompasses pharmacophore development and validation (B1), database exploration for structurally similar compounds (B2), the absorption; distribution; metabolism; excretion; and toxicity (ADMET) and the blood-brain barrier (BBB) filtering to select candidates with favourable pharmacokinetic properties (B3), high-throughput virtual screening (B4), protein structure preparation (B5), and molecular docking (B6) to predict binding modes and affinities. Route A employs the quantitative structure-activity relationship (QSAR) modeling (A1–A2) to establish quantitative relationships between chemical structures and biological activity, followed by molecular docking (A3) and MD simulations with molecular mechanics generalized born surface area (MM-GBSA) to assess the binding stability and energetics. Together, these integrated workflows provide an efficient and comprehensive strategy to prioritize drug candidates for experimental validation [54].
In silico docking, MD simulations, AI-driven structure prediction, high-throughput screening, Cryo-electron microscopy (Cryo-EM), phenotypic assays, and multi-omics integration contribute uniquely to the MPXV therapeutic pipeline (Table 1). High-throughput virtual screening has identified potential inhibitors targeting viral proteins essential for replication and survival. Antibiotics derived from tetracyclines, such as tigecycline and eravacycline, showed substantial binding affinities to DNA-dependent RNA polymerase (DdRp) [56]. MD simulations confirmed this, showing persistent protein-ligand interactions. Similarly, substances such as minocycline and omadacycline efficiently targeted viral proteinases, highlighting the possibility of repurposing antibiotics [56]. Principal component analysis and binding free energy computations supported the stability and specificity of these inhibitors, underscoring their potential as therapeutics [56]. Screening of FDA-approved drugs and plant-derived metabolites identified promising candidates, including DB16335, punicalagin, and nilotinib, after prior studies used structural insights to target MPXV proteins such as thymidylate kinase and the E8 envelope protein [57,58,59]. The studies further confirmed the compounds’ favorable docking scores, dynamic stability, and binding affinities using molecular mechanics Poisson–Boltzmann surface area (MM/PBSA) calculations and MD simulations.
The range of computational DR broaden the computational drug-repurposing efforts for MPXV by integrating genomic data, applying subtractive proteomics, and leveraging artificial intelligence [23]. Database mining identified conserved viral proteins, including A20R, I7L, and Top1B, as druggable targets. High-throughput virtual screening of approved drugs yielded potential multi-target inhibitors, such as burixafor, batefenterol, and eluxadoline, with binding affinities confirmed by MD simulations [23]. FDA-approved drugs, such as elvitegravir, have been prioritized by advanced AI-based platforms such as DeepRepurpose, which employ rigorous computational workflows incorporating homology modelling and metadynamics [61]. Beyond small molecules, plant-derived compounds such as pseudohypericin and amentoflavone exhibit dynamic stability and strong binding, highlighting the utility of bioinformatics for exploring antiviral plant metabolites [13,69].

4.2. Virtual Screening for Antiviral Candidates

Virtual screening, an advanced computational approach that leverages publicly available MPXV genomes and molecular docking techniques, has proven essential for identifying potential antiviral candidates [13,58,70]. Ligand-based and structure-based virtual screening emphasize their potential to address the pressing demand for MPXV treatment. Targeting viral proteins such as A6R (DNA-dependent RNA polymerase), E8 (envelope attachment factor), and thymidylate kinase has identified promising repurposed drugs. Nilotinib, conivaptan, and ponatinib showed favourable binding energies (~−7.5 kcal/mol) against A6R [71], while punicalagin outperformed maraviroc in docking to E8 (−9.1 vs. −7.8 kcal/mol), supported by 100 ns MD simulations and MM-PBSA binding energy calculations [58]. Small-molecule scaffolds such as tipranavir, cefiderocol, and dolutegravir emerged as potential inhibitors of thymidylate kinase and the D9 decapping enzyme [72], while novel heterocyclic compounds such as EH4 exhibited strong binding scores in docking and MD simulations [73]. Furthermore, several computer models have assessed how tecovirimat substitutes interact with viral proteins, including the p37 receptor, showing a high potential for viral inhibition [15,24]. Novel inhibitors that target viral replication pathways and capsid proteins essential to the MPXV lifecycle have been discovered, with potential potency exceeding that of tecovirimat [24,74]. These results highlight the need to integrate MD, genomics, and structure-based design to speed up the identification of antivirals. In addition, integrative platforms such as the Mpox Virus Docking Server offer a convenient and collaborative way to conduct research on target-ligand interactions [75].
DR processes are further enhanced by AI-driven protein modeling and pharmacophore ligand-based approaches, bridging the gap between in silico predictions and experimental validation [12]. These computational pipelines, combined with free-energy calculations and normal mode analysis, prioritize candidates for phenotypic validation and in vivo testing, thus streamlining Mpox drug discovery.

4.3. Phenotypical Drug Repurposing Approaches

Phenotypic DR for MPXV therapeutics leverages assays that identify compounds capable of inhibiting any stage of the viral lifecycle, including entry, uncoating, replication, assembly, or egress or modulating host pathways critical for infection, without requiring prior structural target information [34]. Recent high-throughput cytopathic effect (CPE) and foci-reduction screens in permissive cell lines (Vero E6 and BSC-40) and primary human dermal fibroblasts have identified diverse active scaffolds from FDA-approved drug libraries, including nucleotide polymerase inhibitors (cidofovir, brincidofovir), VP37 egress inhibitors (tecovirimat; nanomolar EC50), antiparasitic (nitazoxanide), antimalarials (mefloquine), and antibiotics with unexpected antiviral activity (nitroxoline, minocycline) [68,76]. Translational workflows use cytotoxicity screens to calculate selectivity indices, orthogonal assays (plaque/foci reduction, qPCR, viral yield), time-of-addition studies to map intervention points, and in vivo testing in CAST/EiJ mice or primates to assess PK/PD and safety [77,78]. Tecovirimat’s clinical use has validated this pipeline while revealing rare F13L/VP37 mutations in immunocompromised patients, highlighting the need for genomic surveillance and resistance monitoring [79]. Phenotypic repurposing captures host-directed and multi-mechanistic antivirals, enabling first-in-class discoveries, but is prone to false positives from cytotoxicity or plasma-protein binding, and many hits fail due to poor PK/PD or toxicity at therapeutic doses [80]. Current best practices integrate phenotypic screening with rapid mechanism deconvolution (chemical genetics, resistance selection, proteomics), in silico ADME/Tox modelling, and combination-matrix assays to identify synergistic drug pairs, raising the genetic barrier to resistance as shown in the human immunodeficiency virus (HIV) and Hepatitis C Virus models [81]. Hybrid pipelines that combine high-content phenotypic assays with computational docking, MD, and AI-driven network pharmacology provide a pragmatic route from the bench to bedside in MPXV drug discovery.

4.4. Systems Biology and Network Pharmacology in Drug Discovery

Systems biology and network pharmacology offer critical insights into the complex dynamics of host–pathogen interactions and therapeutic targeting. In a recent study, 39 major hub proteins were identified as possible therapeutic targets for the human MPXV using a network-based approach that created human-host protein interactomes [38,82]. Analysis revealed that 15 of these targets were already linked to existing drugs, indicating potential for DR [38]. Additionally, the functional enrichment of these hubs revealed pathways associated with viral replication and immune response, highlighting the effect of MPXV infection at the systems level. Baricitinib, infliximab, adalimumab, and etanercept emerged as promising repurposed candidates, while ZINC22060520 was identified as a novel molecule with potential antiviral activity via pharmacophore-based screening [38].
Network pharmacology speeds up the development of targeted therapies by providing a comprehensive understanding of the molecular interactions. Traditional medicine and computational methods have demonstrated potential in discovering new treatment routes to supplement this strategy. Using network pharmacology and molecular docking, research on the traditional Chinese medicinal Shengma-Gegen decoction showed promise against MPXV [83]. Protein–protein interaction analysis identified eight hub targets and 94 active components linked to key viral replication and inflammatory pathways, including TNF, IL-17, and MAPK signaling [38,83]. Similarly, computer modelling revealed that curcumin, a naturally occurring polyphenolic molecule, is a multi-target antiviral drug that targets host pathways including TNF and NF-κB as well as viral proteins [84]. By integrating multi-omics data, computational predictions, and experimental validation, these studies link systems biology with network pharmacology and advance drug discovery against emerging threats.

5. Repurposed Drug Candidates for Mpox Virus

5.1. Antiviral Agents with Potential Efficacy Against Mpox

Several bioactive chemicals that may be used to treat MPXV have been identified using in silico virtual screening, as depicted in the non-exhaustive Table 1 below. A virtual screen of 10,640 compounds identified tucatinib, bictegravir, dolutegravir, and glecaprevir as potential inhibitors of VP37, the MPXV F13 gene product [85]. A similar AlphaFold2-based study predicted the MPXV protein structures and identified trypan blue and catharanthine as candidates with strong binding affinities to multiple viral proteins [62].
DR efforts have also targeted conserved orthopoxvirus proteins involved in host cell interactions and viral DNA replication, alongside computational approaches [13,38,62]. During recent outbreaks, the anti-smallpox medication tecovirimat was repurposed to treat Mpox because it has already been demonstrated to block MPXV replication. Other drugs, including nucleoside/nucleotide analogs such as cidofovir and brincidofovir, have shown antiviral activity against MPXV by chain termination or inhibition of viral DNA polymerase, thereby blocking replication [86,87]. Furthermore, atovaquone and molnupiravir may be able to target the MPXV post-entry processes, according to a recent study on their activity. When combined with tecovirimat, mathematical models indicate that the virus will be significantly cleared [88]. This further highlights how combinatorial treatment can improve antiviral effectiveness. While several repurposed medications show promise, issues such as resistance, toxicity, and the need for more precisely targeted treatments remain. Nitroxoline, an antibiotic with antiviral activity, shows effectiveness against MPXV and retains efficacy against tecovirimat-resistant strains, making it a promising combination therapy option [11,88]. There is potential for the creation of new antivirals through the investigation of novel druggable targets (Table 2), such as those found through proteomics and protein–protein interaction (PPI) networks [37,38,82]. Medications such as fostamatinib, trilostane, and raloxifene suppress host and viral proteins essential for MPXV replication [26]. Combating MPXV and preparing for future outbreaks requires a combination of in silico predictions, experimental validation, and strategic DR.

5.2. Immunomodulatory Drugs for Symptom Management

Immunomodulatory drugs are at the forefront of the DR approach, especially when it comes to addressing the symptoms and immune responses associated with MPXV infection [98]. Tecovirimat and Brincidofovir have demonstrated limited efficacy in treating Mpox, highlighting the urgent need for novel therapeutic strategies [101]. Recent studies have identified potential candidates through diverse DR efforts targeting viral proteins such as envelope proteins, DNA polymerase, and thymidylate kinase [13]. These drugs have the potential to modulate the immune system effectively, providing an alternative method of managing the immune dysregulation seen in Mpox infections [13,24,34]. Immunomodulatory compounds with antiviral properties, such as polyphenols and isoquinoline alkaloids, may reduce disease severity, particularly in immunocompromised individuals [102].
The expanding research on immunomodulatory drugs for Mpox underscores the central role of the immune system in MPXV pathophysiology, as poxviruses such as the MPXV disrupt key mediators, including interleukins, interferons, and tumour necrosis factors [103]. This results in immunological suppression and an accelerated viral replication cycle. By focusing on essential viral enzymes such as thymidine kinase and DNA Topoisomerase I, polyphenolic drugs such as rosmarinic acid and resveratrol have been shown in recent research to suppress MPXV and other poxviruses [102]. By regulating immune responses and preventing viral replication, these bioactive ingredients have the potential to be used as supplemental medicines for symptom control [102,103]. Their anti-inflammatory and antioxidant properties may reduce infection-induced inflammation and improve outcomes, especially in immunocompromised patients prone to severe disease [102,103].
Recent computational and network-based approaches have identified 268 drug candidates, including bioactive compounds, with potential activity against Mpox [82,104]. These candidates were selected for their ability to interact with viral proteins involved in immune regulation [104]. Network medicine has enabled the discovery of immunomodulatory agents targeting key immune hubs, which may prevent MPXV-driven immune disruption [105]. Repurposing such agents could be critical not only for Mpox therapy but also for managing co-infections and associated immunological complications [24,98]. Researchers can accelerate Mpox treatment development by repurposing existing drugs, particularly for vulnerable populations such as individuals with HIV or other immunocompromising conditions.

5.3. Drugs Targeting Mpox-Specific Molecular Pathways

The major viral proteins and enzymes required for the virus reproduction and morphogenesis are the focus of medications that target Mpox-specific biological pathways [18]. Tecovirimat targets the p37 envelope protein, which is essential for the maturation of viruses and the creation of enveloped virions, such as intracellular enveloped virions (IEVs) [18]. By blocking virion release, this protein disrupts the viral life cycle and limits propagation. Additionally, drugs such as brincidofovir target orthopoxvirus DNA polymerase, a key enzyme required for replicating the double-stranded viral genome. However, genetic changes have generated resistance [98]. Methisazone inhibits viral RNA polymerases, which are among the other targets [96,106]. Current antiviral development focuses on targeting replication factors, including thymidine/thymidylate kinases, Abl- and Src-family tyrosine kinases, and key proteases encoded by I7L and G1L [107]. Cellular kinases and transcription factors drive the Mpox infection process, in addition to viral enzymes and proteins [18]. Multi-omics studies have identified numerous host proteins involved in immune evasion and viral replication, highlighting potential targets for therapeutic repurposing [108]. Mpox exploits transcription factors, including STAT3 and NF-κB, and kinases such as ERK1/2, CDK4, and GSK3B, to regulate key cellular functions [34]. FDA-approved kinase inhibitors and steroid hormone receptor agonists have been proposed as repurposing candidates [34]. By modulating host signaling pathways, including Wnt, AP-1, and IFN, these agents may simultaneously inhibit viral replication and enhance immune responses [34]. Interestingly, niclosamide activates antiviral pathways, presenting a potential therapeutic option for Mpox [34]. Researchers have also leveraged the Mpox–human interactome to identify potential therapeutic targets through a comprehensive network-based approach [37,82]. With considerable overlap in proteins linked to immune response pathways, this approach shows that Mpox preferentially targets hubs in the human protein network [70,103]. Repurposing existing drugs to target viral–host protein interactions shows promise, including agents against viral structural proteins A50R and D13L and kinase inhibitors that disrupt key replication pathways [109,110,111]. Understanding these interactions advances Mpox pathophysiology and guides the translation of candidates into preclinical and clinical evaluation.

6. Preclinical and Clinical Evaluation of Repurposed Drugs

6.1. In Vitro and In Vivo Studies on Drug Efficacy

Developing effective therapies for emerging infections such as MPXV requires preclinical and clinical evaluation of repurposed drugs. Candidates include agents effective against other Orthopoxvirus members [112]. TPOXX®, originally developed for smallpox, remains highly effective against circulating MPXV strains in vitro and in vivo, including CAST/EiJ mouse models [18]. It inhibits viral replication at nanomolar concentrations and significantly reduces tissue viral titers, supporting its use in the current outbreak [113]. DNA replication inhibitors cidofovir and its prodrug brincidofovir also show broad-spectrum activity against orthopoxviruses in multiple animal models [105]. Mefloquine, a long-acting antimalarial with known neuropsychiatric and cardiac risks, demonstrated antiviral activity against MPXV clade Ib in foci-reduction assays (IC50 ≈ 0.51–5.2 µM) and docking studies predict engagement with polymerase and envelope-associated proteins, indicating both direct and host-mediated effects [114]. Minocycline, a tetracycline antibiotic (t½ ~11–22 h) with anti-inflammatory properties, emerged from docking and network-pharmacology analyses as a binder to MPXV methyltransferases and host kinases, potentially modulating NF-κB and microglial activation [60]. Whereas mefloquine has corroborating in vitro efficacy, minocycline’s activity remains largely computational and requires phenotypic validation in plaque/foci reduction, cytotoxicity, and time-of-addition assays [89,115]. Mefloquine’s translational use is limited by its safety profile, long persistence, and drug–drug interaction potential. Minocycline carries pregnancy contraindications and interactions with divalent cations that could restrict use [116]. These medications have a potential for DR in MPXV treatment, which is further supported by their effectiveness in animal models. Even while preclinical research has shown promise, practical applications must carefully consider constraints such as the changing nature of MPXV strains and possible antiviral resistance.
Furthermore, nucleoside/nucleotide analogues (cidofovir, brincidofovir) exhibit broad orthopoxvirus activity in vitro and in animal models but carry dose-limiting host toxicities (notably cidofovir-associated nephrotoxicity and brincidofovir-associated transaminase elevations) that constrain clinical use (FDA Tembexa prescribing information; product labels) [117]. High-throughput in vitro screening and in silico docking/MD pipelines identified repurposed candidates, including nitroxoline, niclosamide, mefloquine, atovaquone, and several kinase inhibitors. These compounds show high predicted binding affinities to MPXV targets (VP39 methyltransferase, DNA polymerase, and host kinases) and exhibit antiviral activity in plaque and foci-reduction assays [114,118]. Tecovirimat represents a promising treatment, reducing viral load and improving patient outcomes, as shown in early clinical studies and compassionate use data [119,120,121]. Remdesivir, molnupiravir, and ribavirin also exhibit antiviral activity against MPXV, though their efficacy varies, and further studies are needed to assess their full therapeutic potential [98]. Importantly, MD simulations combined with consensus docking have strengthened confidence in ligand–target interactions by reducing false positives, while phenotypic assays (live-virus inhibition, EC50 determinations) remain essential, as several candidates with favorable docking scores fail to reach therapeutic levels in cell culture or animal models due to pharmacokinetic limitations (poor bioavailability, high protein binding) or safety concerns [122]. Thus, integrating drug repurposing with in vitro, in vivo, and clinical evaluations is crucial to mitigate MPXV impact.

6.2. Insights from Case Studies and Small-Scale Clinical Trials

Antivirals initially developed for smallpox, including tecovirimat, cidofovir, and brincidofovir, demonstrate activity against MPXV and are under evaluation for treating Mpox, particularly in immunocompromised individuals such as people with HIV (https://www.cdc.gov/monkeypox/hcp/clinical-care/immunocompromised-people.html, accessed on 4 October 2025) [15,123]. Tecovirimat is the best-characterized agent, showing robust nanomolar in vitro inhibition and protection in animal models, though clinical outcomes remain heterogeneous [18,124]. Clinical trials and case reports suggest it can reduce viral replication and lesion formation, supported by multicenter registries indicating higher systemic symptoms and lesion incidence in solid organ transplant recipients [31,119,125]. However, many small-scale trials suffer from design limitations and inadequate sample sizes, introducing potential type II errors and underestimating drug efficacy (https://www.who.int/emergencies/disease-outbreak-news/item/2024-DON522, accessed on 4 October 2025) [126]. Consequently, there is growing consensus that larger, well-designed studies are required to accurately assess the therapeutic potential of antivirals such as tecovirimat and smallpox vaccines such as JYNNEOS [127].

6.3. Challenges in Scaling Drug Trials for Emerging Viruses

Scaling medication trials for new viruses raises several pressing issues, especially in preclinical and clinical assessments. Emerging infectious diseases, often vector-borne or zoonotic, remain a major public health threat, exacerbated by environmental changes and global trade [128]. The COVID-19 pandemic demonstrated how rapidly novel viral threats can arise, necessitating swift and coordinated efforts in drug research. The rapid development and deployment of treatments within two years revealed key barriers, including the scarcity of high-quality chemical probes targeting poorly characterized viral proteins. Future medication discovery depends on addressing these issues, particularly for newly emerging viruses such as MPXV, whose genetic drift and changing viral dynamics make diagnostics and treatment approaches more difficult [129]. Enhancing the screening of potential antiviral candidates is a critical priority. This requires leveraging omics technologies and computational power to identify and characterize prospective therapies, while emphasizing rational drug design strategies [13,108].
Large-scale clinical trials for emerging viruses face challenges from diagnostic limitations, variable disease presentations, and complexities in drug evaluation [130,131]. Because of its clinical overlap with other illnesses such as varicella, the diagnosis of MPXV is challenging [18,112]. Low inter-rater reliability plagues accurate lesion evaluation, which is crucial for assessing patient outcomes in clinical trials, particularly in areas with limited resources and diagnostic competence [132]. Furthermore, considering the quickly evolving epidemiology, determining important research goals is crucial, as demonstrated by the global comeback of Mpox in 2022. Successful clinical trials and interventions must ensure equity and inclusivity in developing therapies and vaccines, overcoming logistical and diagnostic obstacles while emphasizing innovative research and public health preparedness against emerging viral threats [125].

7. Limitations and Challenges in Repurposing Drugs for Mpox

In silico models accelerate target identification but face challenges related to data bias, oversimplification of biological complexity, and limited predictive power for dynamic protein-ligand interactions [133]. Furthermore, computational models cannot fully predict pharmacokinetics, off-target effects, or immune responses, necessitating empirical validation.

7.1. Drug–Drug Interactions and Safety Concerns

The network proximity technique, which measures the separation between drug targets on the human interactome and host proteins that Mpox targets, is a significant barrier to finding possible medications [134]. This approach requires further experimental validation, as it cannot distinguish between activation and inhibition or account for dynamic network changes during infection [134,135]. Furthermore, medications found with this static technique may not be effective at every stage of Mpox infection. Applying diverse computational techniques, such as graph convolutional networks and network diffusion algorithms, can provide complementary insights for drug repurposing strategies [37,136]. Access to shelved medications and their trial data is hampered by intellectual property laws and a lack of data openness, which further exacerbates the shortcomings of existing strategies [137,138]. DR still necessitates a substantial financial commitment, estimated to be between hundreds of millions and billions of dollars, while avoiding the expenses of preclinical research [139].
Another significant issue is drug–drug interactions, especially for patients who also have co-infections such as HIV [140,141]. Significant pharmacokinetic variations in tecovirimat exposure between those with and without HIV were found in recent research, including 14 male Mpox patients who were hospitalized. Compared with HIV-negative individuals, people with HIV receiving antiretroviral therapy (ART) showed reduced plasma tecovirimat levels, with minimum and maximum concentrations decreased by 42% and 39%, respectively [141,142]. These variations did not substantially affect clinical results, and all measured drug levels remained above the 90% inhibitory concentration [142]. The study’s limitations, including small sample size, single-centre design, and variability in ART regimen, restrict the generalizability of its findings. Variations in food intake and the absence of free plasma tecovirimat concentration measurements further complicate the understanding of pharmacokinetic interactions [40,142]. These findings underscore the need for extensive clinical trials and real-world data to optimize dosing, prevent resistance, and ensure the safe, effective use of repurposed Mpox therapies.

7.2. Addressing Obstacles and Viral Resistance to Repurposed Agents

Although MPXV, a DNA virus, mutates more slowly than RNA viruses, recent outbreaks show accelerated microevolution, likely from APOBEC3-induced mutations. TPOXX® clinical data reveal heterogeneous outcomes and F13L resistance mutations during prolonged therapy in immunocompromised patients [143,144]. According to reports, the EC50 values of resistant variants are 85–230 times higher than those of wild-type isolates [99,145]. Furthermore, tecovirimat’s limited efficacy against clade I Mpox and its inability to fully suppress viral shedding underscore the urgent need for alternative therapeutic strategies [134]. Other repurposed drugs, such as cidofovir and brincidofovir, have serious adverse effects that prevent them from being used as supplementary therapies [15,146,147,148]. These include renal problems and liver toxicity, respectively [15,24]. Additionally, the logistical and legal obstacles related to the use of these medications make it more difficult to guarantee fair access, particularly in areas with limited resources [149,150].
Combination therapy that targets multiple viral and host pathways offers a promising strategy to address these challenges [14,24]. This necessitates a paradigm shift toward combination therapies pairing tecovirimat with brincidofovir or immunomodulatory agents targeting host pathways (JAK-STAT inhibitors). In preclinical investigations, tecovirimat has shown synergistic efficacy against Mpox when combined with MMF or N-myristoyltransferase (NMT) inhibitors such as IMP-1088 [14]. By raising the genetic barrier to mutations, these combinations increase antiviral effectiveness and reduce the chance of resistance [18,134,146,150]. Clinical studies of NMT inhibitors such as PCLX-001, now in phase II, demonstrate the promise of host-targeted therapies [14]. Combination treatments also reduce the required dosages of individual drugs, lowering associated adverse effects [18,112,134,146]. Implementing such techniques requires robust clinical validation and a determined commitment to expanding access to cutting-edge medicines in both endemic and non-endemic areas.

7.3. Regulatory and Ethical Considerations

Repurposing medications for Mpox faces notable ethical and regulatory challenges. Despite the urgency highlighted by the ongoing multi-country outbreak, gaps remain in understanding transmission dynamics, community-specific effects, and optimal treatment strategies [98,151]. Ensuring the safety and efficacy of vaccines and therapeutics requires robust regulatory oversight and health policy reform. Currently, tecovirimat is accessible for Mpox only under expanded access or compassionate use protocols, necessitating physician registration, informed consent, and continuous data reporting [152]. The WHO recommends using antivirals such as tecovirimat, despite unproven human efficacy, only within clinical trials or under the Monitored Emergency Use of Unregistered Interventions (MEURI) framework, which requires social value, scientific validity, regulatory oversight, informed consent, and contribution to knowledge (https://www.who.int/teams/health-care-readiness/clinical-management-of-monkeypox, accessed on 4 October 2025). Moreover, international regulatory bodies, including the International Coalition of Medicines Regulatory Authorities and the European Medicines Agency, emphasize that coordinated global collaboration, harmonized data collection, and large-scale trials are vital for accelerated but evidence-generating approval pathways for Mpox therapies (https://www.ema.europa.eu/en/news/fostering-regulatory-collaboration-improve-access-mpox-medicines, accessed on 4 October 2025).
ACAM2000, MVA-BN, and LC16 are the three licensed Mpox vaccines; however, further research must clarify their safety and cross-protection profiles [153]. Vulnerable groups, including children, expectant mothers, and those with impaired immune systems, are disproportionately affected by vaccine shortages and uneven supply chains, which further impede fair distribution (https://press.un.org/en/2021/sc14438.doc.htm, accessed on 4 October 2025). Regulatory obstacles such as safety investigations, clinical trials, and authorization procedures delay access and worsen preexisting health inequities [154]. Addressing these ethical and regulatory challenges requires international collaboration to advance vaccine development, strengthen research, and ensure equitable distribution.

8. Future Directions in Mpox Therapeutics

8.1. Integrating Artificial Intelligence and Machine Learning in Drug Repurposing

Mpox therapeutics’ future is situated at the nexus of computational DR, machine learning, and artificial intelligence. There is a gap in targeted therapeutics for Mpox, and integrating artificial intelligence and machine learning (AI/ML) in drug development has shown great promise in discovering possible therapeutic molecules [24,94]. Recent advances, such as identifying PPIs during MPXV infection with ensemble-based deep learning algorithms, have enabled the discovery of promising therapeutic targets [37,155]. Research has used molecular docking and dynamic simulation analysis to identify FDA-approved medications such as fostamatinib, cannabidiol, and nicotinamide adenine dinucleotide + hydrogen as viable possibilities [37]. These computational methods reduce side effects and resistance associated with existing drugs such as Tecovirimat, providing a rapid and cost-effective option to repurpose approved medications. Furthermore, ML-powered algorithms, such as support vector machines and convolutional neural networks, have demonstrated high precision in diagnostic applications, using imaging and clinical data for early detection; a crucial aspect of outbreak control [156,157].
Although there are still obstacles to overcome, further research into AI/ML technology is likely to transform the development of Mpox treatments. Persistent challenges that call for cooperative international efforts to standardize data collecting and sharing include data unpredictability and the requirement for sizable, high-quality databases [134,158]. Thorough clinical studies of repurposed therapies, supported by equitable funding and infrastructure, are essential to address drug resistance and safety concerns. AI-driven, data-informed approaches enable the development of effective antivirals and establish a robust framework for tackling emerging infectious diseases.

8.2. Advancing Combination Therapies and Technologies for Enhanced Efficacy

Future directions in Mpox therapeutics emphasize the urgent need to develop combination therapies, integrating insights from virus–host interactions, viral evolution, and innovative treatment strategies. Identifying effective treatment targets requires understanding the MPXV lifecycle and immune evasion strategies, such as its use of the endoplasmic reticulum as replication factories [112]. The discovery of host components exploited by MPXV through high-throughput screening now enables deeper investigation into viral replication and immune evasion pathways [13]. The emergence of novel lineages, such as IIb C.1, underscores the need for genomic monitoring to track recombination events that drive genetic diversity and immune evasion [13,159,160]. The development of specific antiviral treatments while reducing resistance can be guided by cooperative research into immunomodulatory genes, which influence virus fitness and host immune responses.
The future of Mpox therapy is being shaped by cutting-edge biotechnological techniques, including treatments based on nanotechnology. Through dose-dependent inhibitory effects, nanoparticles such as silver nanoparticles have demonstrated potential in lowering MPXV infectivity [161]. By improving the physicochemical characteristics of antivirals at the nanoscale, these treatments can provide more precise administration and better pharmacokinetic profiles [161,162]. Furthermore, accurate medication design with fewer side effects and resistance is made possible by bioinformatics-driven biomarker-based therapeutics [163]. Incorporating computational models into drug development pipelines can optimize treatment strategies by predicting resistance mutations and enhancing antiviral efficacy. Early diagnosis and outbreak control require advanced diagnostics and expanded genomic surveillance [164,165].

8.3. Establishing Global Networks for Collaborative Research

Prospects for Mpox treatments necessitate the creation of international networks to support cooperative research, especially considering the 2024 outbreak that has affected 15 African countries, leading to more than 6,201 confirmed cases and 32 fatalities in just eight months (https://cdn.who.int/media/docs/default-source/documents/emergencies/20240922_mpox_external-sitrep_-37.pdf?download=true&sfvrsn=1c5db9d1_1, accessed on 4 October 2025). The need for fair cooperation between high- and low-income nations is highlighted by notable research gaps in epidemiology, genetic monitoring, and localized health systems. In the past, the United States has spearheaded global research initiatives, collaborating with endemic areas such as the Democratic Republic of Congo (https://www.who.int/emergencies/disease-outbreak-news/item/2024-DON522, accessed on 4 October 2025) [166]. The prominence of African researchers in leadership roles underscores the need to support indigenous scientists. Strengthening such networks, as shown during COVID-19, can accelerate vaccine delivery, enhance real-time data sharing, and improve genomic surveillance of viral evolution.
The Africa Centers for Disease Control and Prevention (CDC) and the WHO are leading the continent’s response, which includes a $0.6 billion Mpox Continental Preparedness and Response Plan that prioritizes strong health infrastructure, equitable resource distribution, and regional leadership (https://africacdc.org/download/mpox-continental-preparedness-and-response-plan-for-africa/, accessed on 4 October 2025). Reducing reliance on outside sources requires investments in domestic production capacities for medications, vaccines, and diagnostics. Although recent advances, such as the WHO prequalification of a new Mpox vaccine in 2024, show promise, they require improved accessibility and scalable production. Equipping African scientists with the tools to implement context-specific, sustainable solutions will build long-term resilience and enable the continent to manage public health crises more effectively.

9. Conclusions

Recent advances in repurposing existing drugs against the Mpox virus (MPXV) underscore the urgent need for innovative yet pragmatic therapeutic strategies against this re-emerging zoonotic threat. The high genetic similarity between MPXV and variola virus, along with MPXV’s rapid adaptation via mutations in immune evasion and transmissibility genes, underscores the need to leverage safe, well-characterized therapeutics. Promising candidates such as tecovirimat, brincidofovir, nitroxoline, and deoxyuridine analogues, alongside host-directed immunomodulators such as baricitinib and infliximab, illustrate the therapeutic potential of repurposing pipelines. Integration of computational approaches, including molecular docking, pharmacophore modelling, and AI-enhanced structural predictions, has accelerated target identification and candidate prioritization, complementing experimental validation and offering affordable alternatives to conventional antiviral discovery. These developments establish a tiered drug discovery framework, where computational screening, in vitro and in vivo validation, and controlled clinical trials work in synergy to rapidly expand therapeutic options for Mpox.
Despite these advances, significant translational and implementation challenges remain. Repurposed drugs must overcome pharmacokinetic limitations, tissue penetration barriers in systemic and cutaneous disease, and variable efficacy across MPXV clades, as seen in clinical trials such as PALM007. Bridging these gaps demands integrating preclinical-to-clinical pipelines that monitor resistance, test host-directed and combination regimens, and incorporate ADME/Tox profiling early. Future Mpox preparedness should therefore focus on expanding computational-experimental integration, employing organ-on-chip systems, CRISPR-based functional genomics, and Cryo-EM structural validation to enhance translational accuracy. Parallel innovations in vaccines, diagnostics, and equitable access strategies, particularly in under-resourced endemic regions, will be essential to global resilience. Ultimately, successful Mpox DR depends on coordinated international collaboration, harmonized regulatory frameworks, and robust One Health strategies that simultaneously target human, animal, and environmental reservoirs of infection.

Author Contributions

Conceptualization—A.G.-A.M.; writing—original draft preparation, A.G.-A.M.; writing—review and editing, A.G.-A.M. and D.D.M.; supervision, D.D.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

Authors acknowledge Samuel Chima Ugbaja, Tambwe Willy Muzumbukilwa, Mukanda Gedeon Kadima and Nonjabulo Ntombikhona Magwaza for reading and correcting the final version of the manuscript.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Marzo, R.R. Bridging borders: The collective fight against monkeypox outbreaks. In The Scientific Basis of Mpox (Monkeypox); Elsevier: Amsterdam, The Netherlands, 2025; pp. 75–100. [Google Scholar]
  2. Atceken, N.; Bayaki, I.; Can, B.; Yigci, D.; Tasoglu, S. Mpox disease, diagnosis, and point of care platforms. Bioeng. Transl. Med. 2025, 10, e10733. [Google Scholar] [CrossRef]
  3. Sander, W.E. Global One Health and Infectious Diseases: An Interdisciplinary Practitioner’s Guide; CRC Press: Boca Raton, FL, USA, 2025. [Google Scholar]
  4. Scarpa, F.; Azzena, I.; Ciccozzi, A.; Branda, F.; Locci, C.; Perra, M.; Pascale, N.; Romano, C.; Ceccarelli, G.; Terrazzano, G.; et al. Update of the genetic variability of monkeypox virus clade IIb Lineage B.1. Microorganisms 2024, 12, 1874. [Google Scholar] [CrossRef]
  5. Ogunleye, S.C.; Akinsulie, O.C.; Aborode, A.T.; Olorunshola, M.M.; Gbore, D.; Oladoye, M.; Adesola, R.O.; Gbadegoye, J.O.; Olatoye, B.J.; Lawal, M.A.; et al. The re-emergence and transmission of Monkeypox virus in Nigeria the role of one health. Front. Public Health 2024, 11, 1334238. [Google Scholar] [CrossRef]
  6. Mushebenge, A.G.-A.; Mphuthi, D.D. Emerging Insights into Monkeypox: Clinical Features, Epidemiology, Molecular Insights, and Advancements in Management. BioMed 2025, 5, 21. [Google Scholar] [CrossRef]
  7. Cabanillas, B.; Murdaca, G.; Guemari, A.; Torres, M.J.; Azkur, A.K.; Aksoy, E.; Vitte, J.; Fernández-Santamaria, R.; Karavelia, A.; Castagnoli, R.; et al. Monkeypox 2024 outbreak: Fifty essential questions and answers. Allergy 2024, 79, 3285–3309. [Google Scholar] [CrossRef]
  8. Protopapas, K.; Dimopoulou, D.; Kalesis, N.; Akinosoglou, K.; Moschopoulos, C.D. Mpox and Lessons Learned in the Light of the Recent Outbreak: A Narrative Review. Viruses 2024, 16, 1620. [Google Scholar] [CrossRef]
  9. Shafaati, M.; Forghani, S.; Davoudi, A.S.; Samiee, R.; Mohammadi, K.; Akbarpour, S.; Seifi, A.; Salehi, M.; Zare, M. Current advances and challenges in mpox vaccine development a global landscape. Ther. Adv. Vaccines Immunother. 2025, 13, 25151355251314339. [Google Scholar] [CrossRef] [PubMed]
  10. Halwani, M.A. Endemic cross-transmission of Mpox within health-care settings: A comprehensive review. Hail J. Health Sci. 2024, 6, 9–18. [Google Scholar]
  11. Ganesan, A.; Arunagiri, T.; Mani, S.; Kumaran, V.R.; Sk, G.; Elumalai, S.; Kannaiah, K.P.; Chanduluru, H.K. Mpox treatment evolution past milestones, present advances, and future directions. Naunyn-Schmiedeberg’s Arch. Pharmacol. 2024, 398, 1057–1080. [Google Scholar] [CrossRef]
  12. Zhang, W.-Y.; Zheng, X.-L.; Coghi, P.S.; Chen, J.-H.; Dong, B.-J.; Fan, X.-X. Revolutionizing adjuvant development: Harnessing AI for next-generation cancer vaccines. Front. Immunol. 2024, 15, 1438030. [Google Scholar] [CrossRef]
  13. Cambaza, E.M. A review of the molecular understanding of the Mpox virus (MPXV) genomics, immune evasion, and therapeutic targets. Zoonotic Dis. 2025, 5, 3. [Google Scholar] [CrossRef]
  14. Witwit, H.; Cubitt, B.; Khafaji, R.; Castro, E.M.; Goicoechea, M.; Lorenzo, M.M.; Blasco, R.; Martinez-Sobrido, L.; de la Torre, J.C. Repurposing Drugs for Synergistic Combination Therapies to Counteract Monkeypox Virus Tecovirimat Resistance. Viruses 2025, 17, 92. [Google Scholar] [CrossRef]
  15. Shamim, M.A.; Satapathy, P.; Padhi, B.K.; Veeramachaneni, S.D.; Akhtar, N.; Pradhan, A.; Agrawal, A.; Dwivedi, P.; Mohanty, A.; Pradhan, K.B.; et al. Pharmacological treatment and vaccines in monkeypox virus: A narrative review and bibliometric analysis. Front. Pharmacol. 2023, 14, 1149909. [Google Scholar] [CrossRef]
  16. De Clercq, E.; Jiang, Y.; Li, G. Therapeutic strategies for human poxvirus infections: Monkeypox (mpox), smallpox, molluscipox, and orf. Travel Med. Infect. Dis. 2023, 52, 102528. [Google Scholar] [CrossRef]
  17. Nederlof, R.A.; Sainmaa, S.; Wissink-Argilaga, N.; Koo, B.-S.; Bakker, J. Preventative Vaccination of Nonhuman Primates. J. Zool. Bot. Gard. 2025, 6, 8. [Google Scholar] [CrossRef]
  18. Jiang, L.; Xu, A.; Guan, L.; Tang, Y.; Chai, G.; Feng, J.; Wu, Y.; Li, M.; Zhang, C.; Liu, X.; et al. A review of Mpox: Biological characteristics, epidemiology, clinical features, diagnosis, treatment, and prevention strategies. Exploration 2024, 5, 20230112. [Google Scholar] [CrossRef] [PubMed]
  19. Young, B.; Seifert, S.N.; Lawson, C.; Koehler, H. Exploring the genomic basis of Mpox virus-host transmission and pathogenesis. MSphere 2024, 9, e0057624. [Google Scholar] [CrossRef] [PubMed]
  20. Li, Y.; Wang, L.; Chen, S. An overview of the progress made in research into the Mpox virus. Med. Res. Rev. 2025, 45, 788–812. [Google Scholar] [CrossRef]
  21. Eslamkhah, S.; Aslan, E.S.; Yavas, C.; Akcalı, N.; Batur, L.K.; Abuaisha, A.; Yildirim, E.E.; Solak, M.; White, K.N. Mpox virus (MPXV) comprehensive analysis of pandemic risks, pathophysiology, treatments, and mRNA vaccine development. Naunyn-Schmiedeberg’s Arch. Pharmacol. 2025, 398, 6143–6163. [Google Scholar] [CrossRef] [PubMed]
  22. Chaudhary, V.; Lucky, L.; Sable, H.; Bhalla, N. Interdisciplinary Approach to Monkeypox Prevention: Integrating Nanobiosensors, Nanovaccines, Artificial Intelligence, Visual Arts, and Social Sciences. Small Struct. 2025, 6, 2400647. [Google Scholar] [CrossRef]
  23. Sahu, A.; Gaur, M.; Mahanandia, N.C.; Subudhi, E.; Swain, R.P.; Subudhi, B.B. Identification of core therapeutic targets for Monkeypox virus and repurposing potential of drugs against them: An in silico approach. Comput. Biol. Med. 2023, 161, 106971. [Google Scholar] [CrossRef] [PubMed]
  24. Lu, J.; Xing, H.; Wang, C.; Tang, M.; Wu, C.; Ye, F.; Yin, L.; Yang, Y.; Tan, W.; Shen, L. Mpox (formerly monkeypox) pathogenesis, prevention and treatment. Signal Transduct. Target. Ther. 2023, 8, 458. [Google Scholar] [CrossRef]
  25. Mandal, M.; Mandal, S. Metadynamics approach elucidating the free energy landscape and binding modes of vomicine with virulence factor towards anti-mpox drug discovery. Chem. Phys. Lett. 2024, 846, 141355. [Google Scholar] [CrossRef]
  26. Zhu, Z.; Sun, Q.; Xu, Y.; Niu, Y.; Yang, F.; Feng, S. Novel drug targets for monkeypox: From viral to host proteins. Infect. Med. 2025, 4, 100165. [Google Scholar] [CrossRef]
  27. Tan, C.; Zhu, F.; Pan, P.; Wu, A.; Li, C. Development of multi-epitope vaccines against the monkeypox virus based on envelope proteins using immunoinformatics approaches. Front. Immunol. 2023, 14, 1112816. [Google Scholar] [CrossRef]
  28. Rathi, D.; Sahu, L.; Minj, A.; Satapathy, T. Post Invasional Host Interaction Consequences of Monkey Pox Virus and Its Advanced Treatment: A Mechanism Based Review. Curr. Res. Altern Complement Integr. Med. 2024, 1, 26–33. [Google Scholar]
  29. Malik, S.; Ahmed, A.; Ahsan, O.; Muhammad, K.; Waheed, Y. Monkeypox Virus: A Comprehensive Overview of Viral Pathology, Immune Response, and Antiviral Strategies. Vaccines 2023, 11, 1345. [Google Scholar] [CrossRef] [PubMed]
  30. Meyer zu Natrup, C.; Clever, S.; Schünemann, L.M.; Tuchel, T.; Ohrnberger, S.; Volz, A. Strong and early monkeypox virus-specific immunity associated with mild disease after intradermal clade-IIb-infection in CAST/EiJ-mice. Nat. Commun. 2025, 16, 1729. [Google Scholar] [CrossRef]
  31. He, S.; Zhao, J.; Chen, J.; Liang, J.; Hu, X.; Zhang, X.; Zeng, H.; Sun, G. Urogenital Manifestations in Mpox (Monkeypox) Infection: A Comprehensive Review of Epidemiology, Pathogenesis, and Therapeutic Approaches. Infect. Drug Resist. 2025, 18, 209–226. [Google Scholar] [CrossRef]
  32. Brown, B.; Fricke, I.; Imarogbe, C.; González, A.A.P.; Batista, O.A.; Mensah, P.; Chacon-Cruz, E. Immunopathogenesis of Orthopoxviridae: Insights into immunology from smallpox to monkeypox (mpox). Explor. Immunol. 2023, 3, 525–553. [Google Scholar] [CrossRef]
  33. Ganesan, A.; Arunagiri, T.; Mani, S.; Kumaran, V.R.; Kannaiah, K.P.; Chanduluru, H.K. From pox to protection: Understanding Monkeypox pathophysiology and immune resilience. Trop. Med. Health 2025, 53, 33. [Google Scholar] [CrossRef] [PubMed]
  34. Imani, S.; Aminnezhad, S.; Alikarami, M.; Abedi, Z.; Mosleh, I.S.; Maghsoudloo, M.; Taheri, Z. Exploration of drug repurposing for Mpox outbreaks targeting gene signatures and host-pathogen interactions. Sci. Rep. 2024, 14, 29436. [Google Scholar] [CrossRef]
  35. Adler, H.; Gould, S.; Hine, P.; Snell, L.B.; Wong, W.; Houlihan, C.F.; Osborne, J.C.; Rampling, T.; Beadsworth, M.B.; Duncan, C.J.; et al. Clinical features and management of human monkeypox: A retrospective observational study in the UK. Lancet Infect. Dis. 2022, 22, 1153–1162, Correction in Lancet Infect. Dis. 2022, 22, E177. [Google Scholar] [CrossRef] [PubMed]
  36. Hua, Y.; Dai, X.; Xu, Y.; Xing, G.; Liu, H.; Lu, T.; Chen, Y.; Zhang, Y. Drug repositioning: Progress and challenges in drug discovery for various diseases. Eur. J. Med. Chem. 2022, 234, 114239. [Google Scholar] [CrossRef]
  37. Paul, D.; Saha, S.; Basu, S.; Chakraborti, T. Computational analysis of pathogen-host interactome for fast and low-risk in-silico drug repurposing in emerging viral threats like Mpox. Sci. Rep. 2024, 14, 18736. [Google Scholar] [CrossRef]
  38. Rabaan, A.A.; Alfaresi, M.; Alrasheed, H.A.; Al Kaabi, N.A.; Abduljabbar, W.A.; Al Fares, M.A.; Al-Subaie, M.F.; Alissa, M. Network-Based Drug Repurposing and Genomic Analysis to Unveil Potential Therapeutics for Monkeypox Virus. Chem. Biodivers. 2024, 21, e202400895. [Google Scholar] [CrossRef]
  39. Ahamed, M.A.; Politza, A.J.; Liu, T.; Khalid, M.A.U.; Zhang, H.; Guan, W. CRISPR-based strategies for sample-to-answer monkeypox detection: Current status and emerging opportunities. Nanotechnology 2024, 36, 042001. [Google Scholar] [CrossRef]
  40. Hou, W.; Wu, N.; Liu, Y.; Tang, Y.; Quan, Q.; Luo, Y.; Jin, C. Mpox: Global epidemic situation and countermeasures. Virulence 2025, 16, 2457958. [Google Scholar] [CrossRef]
  41. Kalil, A.C.; Patterson, T.F.; Mehta, A.K.; Tomashek, K.M.; Wolfe, C.R.; Ghazaryan, V.; Marconi, V.C.; Ruiz-Palacios, G.M.; Hsieh, L.; Kline, S.; et al. Baricitinib plus remdesivir for hospitalized adults with Covid-19. N. Engl. J. Med. 2021, 384, 795–807. [Google Scholar] [CrossRef]
  42. Rejinold, N.S.; Jin, G.; Choy, J. Harnessing Nanohybridized Niclosamide for Precision Mpox Therapeutics. Adv. Healthc. Mater. 2025, 14, 2404818. [Google Scholar] [CrossRef] [PubMed]
  43. Yadesa Tesema, M. Mitigation approaches of monkeypox outbreaks in developing countries. Discov. Viruses 2025, 2, 4. [Google Scholar] [CrossRef]
  44. Pushpakom, S.; Iorio, F.; Eyers, P.A.; Escott, K.J.; Hopper, S.; Wells, A.; Doig, A.; Guilliams, T.; Latimer, J.; Mcnamee, C.; et al. Drug repurposing: Progress, challenges and recommendations. Nat. Rev. Drug Discov. 2019, 18, 41–58. [Google Scholar] [CrossRef]
  45. Yousaf, M.A.; Michel, M.; Khan, A.T.-A.; Noreen, M.; Bano, S. Repurposing doxycycline for the inhibition of monkeypox virus DNA polymerase: A comprehensive computational study. Silico Pharmacol. 2025, 13, 27. [Google Scholar] [CrossRef] [PubMed]
  46. Shannon, A.; Canard, B. Nucleotide analogues and mpox: Repurposing the repurposable. Antivir. Res. 2024, 234, 106057. [Google Scholar] [CrossRef]
  47. Bunge, E.M.; Hoet, B.; Chen, L.; Lienert, F.; Weidenthaler, H.; Baer, L.R.; Steffen, R. The changing epidemiology of human monkeypox—A potential threat? A systematic review. PLoS Neglected Trop. Dis. 2022, 16, e0010141. [Google Scholar] [CrossRef]
  48. Kozlov, M. Monkeypox in Africa: The science the world ignored. Nature 2022, 607, 17–18. [Google Scholar] [CrossRef]
  49. Girometti, N.; Byrne, R.; Bracchi, M.; Heskin, J.; McOwan, A.; Tittle, V.; Gedela, K.; Scott, C.; Patel, S.; Gohil, J.; et al. Demographic and clinical characteristics of confirmed human monkeypox virus cases in individuals attending a sexual health centre in London, UK: An observational analysis. Lancet Infect. Dis. 2022, 22, 1321–1328. [Google Scholar] [CrossRef] [PubMed]
  50. Rabaan, A.A.; Halwani, M.A.; Alshehri, A.A.; Al-Subaie, M.F.; Almansour, Z.H.; AlShehail, B.M.; Alotaibi, N.; Khamis, F.; Al Kaabi, N.A.; Alsomali, G.; et al. Bioprospecting of Meliaceae family phytomolecules for the treatment of monkeypox virus infection: A QSAR modeling and MD simulation approach. J. Biomol. Struct. Dyn. 2025, 43, 2277–2299. [Google Scholar] [CrossRef]
  51. Meng, X.-Y.; Zhang, H.-X.; Mezei, M.; Cui, M. Molecular docking: A powerful approach for structure-based drug discovery. Curr. Comput. Aided Drug Des. 2011, 7, 146–157. [Google Scholar] [CrossRef] [PubMed]
  52. Mushebenge, A.G.; Ugbaja, S.C.; Mtambo, S.E.; Ntombela, T.; Metu, J.I.; Babayemi, O.; Chima, J.I.; Appiah-Kubi, P.; Odugbemi, A.I.; Ntuli, M.L.; et al. Unveiling the Inhibitory Potentials of Peptidomimetic Azanitriles and Pyridyl Esters towards SARS-CoV-2 Main Protease: A Molecular Modelling Investigation. Molecules 2023, 28, 2641. [Google Scholar] [CrossRef]
  53. Hollingsworth, S.A.; Dror, R.O. Molecular dynamics simulation for all. Neuron 2018, 99, 1129–1143. [Google Scholar] [CrossRef]
  54. Ugbaja, S.C.; Mushebenge, A.G.-A.; Kumalo, H.; Ngcobo, M.; Gqaleni, N. Potential Benefits of In Silico Methods: A Promising Alternative in Natural Compound’s Drug Discovery and Repurposing for HBV Therapy. Pharmaceuticals 2025, 18, 419. [Google Scholar] [CrossRef]
  55. Mushebenge, A.G.-A.; Ugbaja, S.C.; Magwaza, N.N.; Mbatha, N.A.; Muzumbukilwa, T.W.; Kadima, M.G.; Tata, F.Y.; Nxumalo, M.B.; Manimani, R.G.; Ndage, N.; et al. Mechanistic Insights into the Mutational Landscape of the Main Protease/3CLPro and Its Impact on Long-Term COVID-19/SARS-CoV-2 Management. Future Pharmacol. 2024, 4, 825–852. [Google Scholar] [CrossRef]
  56. Alandijany, T.A.; El-Daly, M.M.; Tolah, A.M.; Bajrai, L.H.; Khateb, A.M.; Kumar, G.S.; Dubey, A.; Dwivedi, V.D.; Azhar, E.I. A multi-targeted computational drug discovery approach for repurposing tetracyclines against monkeypox virus. Sci. Rep. 2023, 13, 14570. [Google Scholar] [CrossRef] [PubMed]
  57. Ajmal, A.; Mahmood, A.; Hayat, C.; Hakami, M.A.; Alotaibi, B.S.; Umair, M.; Abdalla, A.N.; Li, P.; He, P.; Wadood, A.; et al. Computer-assisted drug repurposing for thymidylate kinase drug target in monkeypox virus. Front. Cell. Infect. Microbiol. 2023, 13, 1159389. [Google Scholar] [CrossRef] [PubMed]
  58. Das, R.; Bhattarai, A.; Karn, R.; Tamang, B. Computational investigations of potential inhibitors of monkeypox virus envelope protein E8 through molecular docking and molecular dynamics simulations. Sci. Rep. 2024, 14, 19585. [Google Scholar] [CrossRef]
  59. Dhapola, R.; Kumari, S.; Sharma, P.; KumarKushawaha, P.; HariKrishnaReddy, D. Update on monkeypox virus infection: Focusing current treatment and prevention approaches. Fundam. Clin. Pharmacol. 2024, 38, 465–478. [Google Scholar] [CrossRef] [PubMed]
  60. Maurya, V.K.; Kumar, S.; Maurya, S.; Ansari, S.; Paweska, J.T.; Abdel-Moneim, A.S.; Saxena, S.K. Structure-based drug designing for potential antiviral activity of selected natural product against Monkeypox (Mpox) virus and its host targets. VirusDisease 2024, 35, 589–608. [Google Scholar] [CrossRef]
  61. Patel, C.N.; Mall, R.; Bensmail, H. AI-driven drug repurposing and binding pose meta dynamics identifies novel targets for monkeypox virus. J. Infect. Public Health 2023, 16, 799–807. [Google Scholar] [CrossRef]
  62. Ezat, A.A.; Abduljalil, J.M.; Elghareib, A.M.; Samir, A.; Elfiky, A.A. The discovery of novel antivirals for the treatment of mpox: Is drug repurposing the answer? Expert Opin. Drug Discov. 2023, 18, 551–561. [Google Scholar] [CrossRef]
  63. Hong, Y.; Huang, B.; Zhang, J.; Peng, C.; Kong, W.; Tan, W.; Li, S. Molecular architecture of monkeypox mature virus. Cell Discov. 2024, 10, 108. [Google Scholar] [CrossRef] [PubMed]
  64. Ivanisevic, T.; Sewduth, R.N. Multi-Omics Integration for the Design of Novel Therapies and the Identification of Novel Biomarkers. Proteomes 2023, 11, 34. [Google Scholar] [CrossRef] [PubMed]
  65. Perdomo-Quinteiro, P.; Wolstencroft, K.; Roos, M.; Queralt-Rosinach, N. Knowledge graphs and explainable ai for drug repurposing on rare diseases. bioRxiv 2024. bioRxiv:10.17.618804. [Google Scholar] [CrossRef]
  66. El-kenawy, E.-S.M. Leveraging Advanced Machine Learning for Pioneering Monkeypox Diagnosis: A New Paradigm in Infectious Disease Detection. Metaheuristic Optim. Rev. 2024, 1, 1–16. [Google Scholar]
  67. Pan, Q.; Li, P. Human skin organoids are valid models of mpox virus infection. Nature 2023, 8, 1950–1951. [Google Scholar]
  68. Rosa, R.B.; Ferreira de Castro, E.; Vieira da Silva, M.; Paiva Ferreira, D.C.; Jardim, A.C.G.; Santos, I.A.; Marinho, M.D.S.; Ferreira França, F.B.; Pena, L.J. In vitro and in vivo models for monkeypox. iScience 2023, 26, 105702. [Google Scholar] [CrossRef]
  69. Hossain, F.M.A.; Bappy, N.I.; Robin, T.B.; Ahmad, I.; Patel, H.; Jahan, N.; Rabbi, G.R.; Roy, A.; Chowdhury, W.; Ahmed, N.; et al. A review on computational studies and bioinformatics analysis of potential drugs against monkeypox virus. J. Biomol. Struct. Dyn. 2024, 42, 6091–6107. [Google Scholar] [CrossRef]
  70. Loganathan, T.; Fletcher, J.; Abraham, P.; Kannangai, R.; Chakraborty, C.; El Allali, A.; Alsamman, A.M.; Zayed, H.; C, G.P.D. Expression analysis and mapping of Viral-Host Protein interactions of Poxviridae suggests a lead candidate molecule targeting Mpox. BMC Infect. Dis. 2024, 24, 483. [Google Scholar] [CrossRef]
  71. Abdizadeh, T. Identification of novel potential inhibitors of monkeypox virus thymidine kinase using molecular docking, molecular dynamics simulation and MM/PBSA methods. Mol. Divers. 2024, 28, 2513–2546. [Google Scholar] [CrossRef]
  72. Sahoo, A.K.; Augusthian, P.D.; Muralitharan, I.; Vivek-Ananth, R.P.; Kumar, K.; Kumar, G.; Ranganathan, G.; Samal, A. In silico identification of potential inhibitors of vital monkeypox virus proteins from FDA approved drugs. Mol. Divers. 2023, 27, 2169–2184. [Google Scholar] [CrossRef]
  73. Kaur, M.; Sharma, A.; Kaur, H.; Singh, M.; Devi, B.; Raj, A.R.N.; Sood, V.; Pandey, A.; Gartia, J.; Kumar, R.; et al. Screening of potential inhibitors against structural proteins from Monkeypox and related viruses of Poxviridae family via docking and molecular dynamics simulation. J. Biomol. Struct. Dyn. 2024, 42, 10978–10993. [Google Scholar] [CrossRef]
  74. Duchoslav, V.; Boura, E. Structure of monkeypox virus poxin implications for drug design. Arch. Virol. 2023, 168, 192. [Google Scholar] [CrossRef]
  75. Arasu, M.; Vijayaragavan, P.; Purushothaman, S.; Rathi, M.; Al-Dhabi, N.A.; Gopalakrishnan, V.; Choi, K.C.; Ilavenil, S. Molecular docking of monkeypox (mpox) virus proteinase with FDA approved lead molecules. J. Infect. Public Health 2023, 16, 784–791. [Google Scholar] [CrossRef]
  76. Harrison, L.B.; Bergeron, G.; Cadieux, G.; Charest, H.; Fafard, J.; Levade, I.; Blais, A.C.; Huchet, E.; Trottier, B.; Vlad, D.; et al. Monkeypox in Montréal: Epidemiology, phylogenomics, and public health response to a large North American outbreak. Ann. Intern. Med. 2023, 176, 67–76. [Google Scholar] [CrossRef] [PubMed]
  77. Liu, W.; Zhang, E.; Li, W.; Lv, R.; Lin, Y.; Xu, Y.; Li, J.; Lai, Y.; Jiang, Y.; Lin, S.; et al. Advances and challenges of mpox detection technology. Biosaf. Health 2024, 6, 260–269. [Google Scholar] [CrossRef] [PubMed]
  78. Barnard, D.L.; Day, C.W.; Bailey, K.; Heiner, M.; Montgomery, R.; Lauridsen, L.; Winslow, S.; Hoopes, J.; Li, J.K.-K.; Lee, J.; et al. Enhancement of the infectivity of SARS-CoV in BALB/c mice by IMP dehydrogenase inhibitors, including ribavirin. Antivir. Res. 2006, 71, 53–63. [Google Scholar] [CrossRef]
  79. Chenchula, S.; Atal, S.; Ghanta, M.K.; Uppugunduri, C.R.; Karunakaran, S.; Amerneni, K.C.; Sarma, P.; Prakash, S.; Amerneni, L.S.; Padmavathi, R.; et al. Emerging variants of Mpox virus and tecovirimat resistance: Genomic insights and implications for treatment strategies. Virology 2025, 608, 110532. [Google Scholar] [CrossRef] [PubMed]
  80. Rietdijk, J.; Tampere, M.; Pettke, A.; Georgiev, P.; Lapins, M.; Warpman-Berglund, U.; Spjuth, O.; Puumalainen, M.-R.; Carreras-Puigvert, J. A phenomics approach for antiviral drug discovery. BMC Biol. 2021, 19, 156. [Google Scholar] [CrossRef]
  81. Vincent, F.; Nueda, A.; Lee, J.; Schenone, M.; Prunotto, M.; Mercola, M. Phenotypic drug discovery: Recent successes, lessons learned and new directions. Nat. Rev. Drug Discov. 2022, 21, 899–914, Erratum in Nat. Rev. Drug Discov. 2022, 21, 541. [Google Scholar] [CrossRef]
  82. Tang, K.; Sun, Q.; Zeng, J.; Tang, J.; Cheng, P.; Qiu, Z.; Long, H.; Chen, Y.; Zhang, C.; Wei, J.; et al. Network-based approach for drug repurposing against mpox. Int. J. Biol. Macromol. 2024, 270, 132468. [Google Scholar] [CrossRef]
  83. Dai, L.; Zhang, G.; Wan, X. Network Pharmacology and Molecular Docking Analysis Reveal Insights into the Molecular Mechanism of Shengma-Gegen Decoction on Monkeypox. Pathogens 2022, 11, 1342. [Google Scholar] [CrossRef]
  84. Ho, W.-Y.; Shen, Z.-H.; Chen, Y.; Chen, T.-H.; Lu, X.; Fu, Y.-S. Therapeutic implications of quercetin and its derived-products in COVID-19 protection and prophylactic. Heliyon 2024, 10, e30080. [Google Scholar] [CrossRef] [PubMed]
  85. Chen, Z.R.; Lv, Q.-L.; Peng, H.-W.; Liu, X.-Y.; Hu, W.-L.; Hu, J.-F. Drug screening against F13 protein, the target of tecovirimat, as potential therapies for monkeypox virus. J. Infect. 2023, 86, 154–225. [Google Scholar] [CrossRef]
  86. Kumari, S.; Aneja, S.; Biswal, B.K. Approved and experimental antivirals for mpox. In The Scientific Basis of Mpox (Monkeypox); Elsevier: Amsterdam, The Netherlands, 2025; pp. 401–420. [Google Scholar]
  87. Siegrist, E.A.; Sassine, J. Antivirals with activity against mpox: A clinically oriented review. Clin. Infect. Dis. 2023, 76, 155–164. [Google Scholar] [CrossRef] [PubMed]
  88. Akazawa, D.; Ohashi, H.; Hishiki, T.; Morita, T.; Iwanami, S.; Kim, K.S.; Jeong, Y.D.; Park, E.-S.; Kataoka, M.; Shionoya, K.; et al. Potential anti-mpox virus activity of atovaquone, mefloquine, and molnupiravir, and their potential use as treatments. J. Infect. Dis. 2023, 228, 591–603. [Google Scholar] [CrossRef]
  89. Winkler, D.A. Computational repurposing of drugs for viral diseases and current and future pandemics. J. Math. Chem. 2024, 62, 2844–2879. [Google Scholar] [CrossRef]
  90. Wang, J.; Mahmud, S.A.; Chen, A.; Li, K.; Tan, H.; Joyce, R. An overview of antivirals against monkeypox virus and other orthopoxviruses. J. Med. Chem. 2023, 66, 4468–4490. [Google Scholar] [CrossRef]
  91. Filho, E.B.; Neto, R.F.O.; Santos, V.C.; Ferreira, A.L.S. Deep Reinforcement Learning and Structure-Based Approaches in the de novo Design of a New Potential Inhibitor of F13 Protein from Monkeypox Virus. J. Braz. Chem. Soc. 2025, 36, e-20250022. [Google Scholar]
  92. Shah, B.M.; Modi, P. Breaking Barriers: Current Advances and Future Directions in Mpox Therapy. Curr. Drug Targets 2024, 25, 62–76. [Google Scholar] [CrossRef]
  93. Dutt, M.; Kumar, A.; Rout, M.; Dehury, B.; Martinez, G.; Ndishimye, P.; Kelvin, A.A.; Kelvin, D.J. Drug repurposing for Mpox: Discovery of small molecules as potential inhibitors against DNA-dependent RNA polymerase using molecular modeling approach. J. Cell. Biochem. 2023, 124, 701–715. [Google Scholar] [CrossRef]
  94. Grajales, D.B.; Kar, S. Exploring Monkeypox: Prospects for therapeutics through computational-aided drug discovery. Mol. Divers. 2023, 28, 3497–3521. [Google Scholar] [CrossRef] [PubMed]
  95. Tiwari, H.; Ilyas, A.; Upadhyay, S.; Rai, P.K.; Borkotoky, S. Computational screening of FDA-approved and natural compounds against Mpox Dual specificity protein phosphatase (H1). Indian J. Biochem. Biophys. 2025, 62, 398–405. [Google Scholar] [CrossRef]
  96. Malik, S.; Ahmad, T.; Ahsan, O.; Muhammad, K.; Waheed, Y. Recent Developments in Mpox Prevention and Treatment Options. Vaccines 2023, 11, 500. [Google Scholar] [CrossRef]
  97. Bojkova, D.; Zöller, N.; Tietgen, M.; Steinhorst, K.; Bechtel, M.; Rothenburger, T.; Kandler, J.D.; Schneider, J.; Corman, V.M.; Ciesek, S.; et al. Repurposing of the antibiotic nitroxoline for the treatment of mpox. J. Med. Virol. 2023, 95, e28652. [Google Scholar] [CrossRef]
  98. Kumar, S.; Guruparan, D.; Karuppanan, K.; Kumar, K.J.S. Comprehensive Insights into Monkeypox (mpox): Recent Advances in Epidemiology, Diagnostic Approaches and Therapeutic Strategies. Pathogens 2025, 14, 1. [Google Scholar] [CrossRef]
  99. Vernuccio, R.; León, A.M.; Poojari, C.S.; Buchrieser, J.; Selverian, C.; Jaleta, Y.; Meola, A.; Guivel-Benhassine, F.; Porrot, F.; Haouz, A.; et al. Mechanisms of tecovirimat antiviral activity and poxvirus resistance. Res. Sq. 2024. preprint. [Google Scholar]
  100. Hashemi, M.; Zabihian, A.; Hajsaeedi, M.; Hooshmand, M. Antivirals for monkeypox virus: Proposing an effective machine/deep learning framework. PloS ONE 2024, 19, e0299342. [Google Scholar] [CrossRef]
  101. Rizk, Y.; Lippi, G.; Henry, B.M.; Notarte, K.I.; Rizk, J.G. Update on mpox management: Epidemiology, vaccines and therapeutics, and regulatory changes. Drugs 2025, 85, 1–9. [Google Scholar] [CrossRef]
  102. Ali, S.I.; Salama, A. Natural Immunomodulatory Agents as a Complementary Therapy for Poxviruses. Poxviruses 2024, 2024, 337–354. [Google Scholar]
  103. Fang, D.; Liu, Y.; Dou, D.; Su, B. The unique immune evasion mechanisms of the mpox virus and their implication for developing new vaccines and immunotherapies. Virol. Sin. 2024, 39, 709–718. [Google Scholar] [CrossRef]
  104. Lian, Z.H.; Yang, C.-H.; Qiu, Y.; Ge, X.-Y. Evolutionary Analysis and Antiviral Drug Prediction of Mpox Virus. Microorganisms 2024, 12, 2239. [Google Scholar] [CrossRef]
  105. Lee, S.Z. Subcellular Investigation of Poxvirus Coinfections and the Impact on the Evolution of Cidofovir Resistance. Master’s Thesis, University of Alberta, Edmonton, AB, Canada, 2024. [Google Scholar]
  106. Lee, W.; Kim, Y.-J.; Lee, S.J.; Ahn, D.-G.; Kim, S.-J. Current Status of Epidemiology, Diagnosis, Therapeutics, and Vaccines for the Re-Emerging Human Monkeypox Virus. J. Microbiol. Biotechnol. 2023, 33, 981–991. [Google Scholar] [CrossRef]
  107. Prichard, M.N.; Kern, E.R. Orthopoxvirus targets for the development of new antiviral agents. Antivir. Res. 2012, 94, 111–125. [Google Scholar] [CrossRef]
  108. Huang, Y.; Bergant, V.; Grass, V.; Emslander, Q.; Hamad, M.S.; Hubel, P.; Mergner, J.; Piras, A.; Krey, K.; Henrici, A.; et al. Multi-omics characterization of the monkeypox virus infection. Nat. Commun. 2024, 15, 6778. [Google Scholar] [CrossRef]
  109. Duan, H.; Shi, Q.; Yue, X.; Zhang, Z.; Liu, L.; Wang, Y.; Cao, Y.; Ou, Z.; Liang, L.; Hu, J.; et al. Identification of core therapeutic targets for Monkeypox virus and repurposing potential of drugs: A WEB prediction approach. PLoS ONE 2024, 19, e0303501. [Google Scholar] [CrossRef] [PubMed]
  110. Mohapatra, R.K.; Mahal, A.; Mohapatra, P.K.; Sarangi, A.K.; Mishra, S.; Alsuwat, M.A.; Alshehri, N.N.; Abdelkhalig, S.M.; Garout, M.; Aljeldah, M.; et al. Structure-based discovery of F. religiosa phytochemicals as potential inhibitors against Monkeypox (mpox) viral protein. J. Biosaf. Biosecur. 2024, 6, 157–169. [Google Scholar] [CrossRef]
  111. Lam, H.Y.I.; Guan, J.S.; Mu, Y. In silico repurposed drugs against monkeypox virus. Molecules 2022, 27, 5277. [Google Scholar] [CrossRef]
  112. Mohanto, S.; Faiyazuddin; Gholap, A.D.; Darshan, J.C.; Bhunia, A.; Subbaram, K.; Ahmed, M.G.; Nag, S.; Akhtar, M.S.; Bonilla-Aldana, D.K.; et al. Addressing the resurgence of global monkeypox (Mpox) through advanced drug delivery platforms. Travel Med. Infect. Dis. 2023, 56, 102636. [Google Scholar] [CrossRef] [PubMed]
  113. Warner, B.M.; Klassen, L.; Sloan, A.; Deschambault, Y.; Soule, G.; Banadyga, L.; Cao, J.; Strong, J.E.; Kobasa, D.; Safronetz, D. In vitro and in vivo efficacy of tecovirimat against a recently emerged 2022 monkeypox virus isolate. Sci. Transl. Med. 2022, 14, eade7646. [Google Scholar] [CrossRef] [PubMed]
  114. El Sharazly, B.M.; Ahmed, A.; Elsheikha, H.M.; Carter, W.G. An In Silico and In Vitro Assessment of the Neurotoxicity of Mefloquine. Biomedicines 2024, 12, 505. [Google Scholar] [CrossRef]
  115. Leela, S.L.; Srisawat, C.; Sreekanth, G.P.; Noisakran, S.; Yenchitsomanus, P.-T.; Limjindaporn, T. Drug repurposing of minocycline against dengue virus infection. Biochem. Biophys. Res. Commun. 2016, 478, 410–416. [Google Scholar] [CrossRef]
  116. González, R.; Hellgren, U.; Greenwood, B.; Menéndez, C. Mefloquine safety and tolerability in pregnancy: A systematic literature review. Malar. J. 2014, 13, 75. [Google Scholar] [CrossRef]
  117. Andrei, G.; Snoeck, R. Cidofovir Activity against Poxvirus Infections. Viruses 2010, 2, 2803–2830. [Google Scholar] [CrossRef]
  118. Manikyam, H.K.; Patil, S.B.; Hussain, N.; Vallal, R.E.E.; Sharma, S.; Patil, A.R. High-Throughput Insilico Drug Screen against Mpox Targeted Proteins in Comparison with Repurposed Antiviral Drugs against Natural Compounds. J. Pharm. Res. Int. 2024, 36, 41–52. [Google Scholar] [CrossRef]
  119. McLean, J.; Stoeckle, K.; Huang, S.; Berardi, J.; Gray, B.; Glesby, M.J.; Zucker, J. Tecovirimat Treatment of People With HIV During the 2022 Mpox Outbreak: A Retrospective Cohort Study. Ann. Intern. Med. 2023, 176, 642–648. [Google Scholar] [CrossRef]
  120. DeLaurentis, C.E.; Kiser, J.; Zucker, J. New Perspectives on Antimicrobial Agents: Tecovirimat for Treatment of Human Monkeypox Virus. Antimicrob. Agents Chemother. 2022, 66, e0122622. [Google Scholar] [CrossRef]
  121. Shabil, M.; Khatib, M.N.; Ballal, S.; Bansal, P.; Tomar, B.S.; Ashraf, A.; Kumar, M.R.; Sinha, A.; Rawat, P.; Gaidhane, A.M.; et al. Effectiveness of Tecovirimat in Mpox Cases: A Systematic Review of Current Evidence. J. Med. Virol. 2024, 96, e70122. [Google Scholar] [CrossRef] [PubMed]
  122. Aghajani, J.; Farnia, P.; Farnia, P.; Ghanavi, J.; Velayati, A.A. Molecular Dynamic Simulations and Molecular Docking as a Potential Way for Designed New Inhibitor Drug without Resistance. Tanaffos 2022, 21, 1–14. [Google Scholar] [PubMed]
  123. Prévost, J.; Sloan, A.; Deschambault, Y.; Tailor, N.; Tierney, K.; Azaransky, K.; Kammanadiminti, S.; Barker, D.; Kodihalli, S.; Safronetz, D. Treatment efficacy of cidofovir and brincidofovir against clade II Monkeypox virus isolates. Antivir. Res. 2024, 231, 105995. [Google Scholar] [CrossRef]
  124. Warner, N.C.; Shishido, A.; Fulco, P.P.; Sastry, S. Immune reconstitution inflammatory syndrome due to monkeypox in two patients with AIDS. AIDS 2023, 37, 1187–1188. [Google Scholar] [CrossRef] [PubMed]
  125. Singh, P.; Sridhar, S.B.; Shareef, J.; Talath, S.; Mohapatra, P.; Khatib, M.N.; Ballal, S.; Kaur, M.; Nathiya, D.; Sharma, S.; et al. The resurgence of monkeypox: Epidemiology, clinical features, and public health implications in the post-smallpox eradication era. New Microbes New Infect. 2024, 62, 101487. [Google Scholar] [CrossRef] [PubMed]
  126. Alorfi, N.M.; Alshehri, F.S.; Haseeb, A. Characterization of interventional clinical trials for monkeypox; systematic review of ClinicalTrials.gov database. Front. Public Health 2023, 11, 1144325. [Google Scholar] [CrossRef]
  127. Roper, R.L.; Garzino-Demo, A.; Del Rio, C.; Bréchot, C.; Gallo, R.; Hall, W.; Esparza, J.; Reitz, M.; Schinazi, R.F.; Parrington, M.; et al. Monkeypox (Mpox) requires continued surveillance, vaccines, therapeutics and mitigating strategies. Vaccine 2023, 41, 3171–3177. [Google Scholar] [CrossRef]
  128. Tajudeen, Y.A.; Oladipo, H.J.; Muili, A.O.; Ikebuaso, J.G. Monkeypox: A review of a zoonotic disease of global public health concern. Health Promot. Perspect. 2023, 13, 1–9. [Google Scholar] [CrossRef] [PubMed]
  129. Olawade, D.B.; Wada, O.Z.; Fidelis, S.C.; Oluwole, O.S.; Alisi, C.S.; Orimabuyaku, N.F.; David-Olawade, A.C. Strengthening Africa’s response to Mpox (monkeypox): Insights from historical outbreaks and the present global spread. Sci. One Health 2024, 3, 100085. [Google Scholar] [CrossRef] [PubMed]
  130. Garcia-Junior, M.A.; Andrade, B.S.; Guevara-Vega, M.; de Melo, I.S.; Cunha, T.M.; Jardim, A.C.G.; Sabino-Silva, R. Oral Infection, Oral Pathology and Salivary Diagnostics of Mpox Disease: Relevance in Dentistry and OMICs Perspectives. Int. J. Mol. Sci. 2023, 24, 14362. [Google Scholar] [CrossRef] [PubMed]
  131. Santos, G.R.d.S.; Ribeiro, C.J.N.; Santos Júnior, J.F.C.d.; Almeida, V.S.; Nascimento, R.d.C.D.; Barreto, N.M.P.V.; Sousa, A.R.d.; Bezerra-Santos, M.; Cepas, L.A.; Fernandes, A.P.M.; et al. Mpox Vaccine Hesitancy Among Brazilian Men Who Have Sex with Men: A National Cross-Sectional Study. Vaccines 2024, 12, 1229. [Google Scholar] [CrossRef]
  132. Bourner, J.; Garcia-Gallo, E.; Mbrenga, F.; Boum, Y.; Nakouné, E.; Paterson, A.; Jones, B.; Olliaro, P.; Rojek, A. Challenges in clinical diagnosis of Clade I Mpox: Highlighting the need for enhanced diagnostic approaches. PLoS Negl. Trop. Dis. 2024, 18, e0012087. [Google Scholar] [CrossRef]
  133. Jumper, J.; Evans, R.; Pritzel, A.; Green, T.; Figurnov, M.; Ronneberger, O.; Tunyasuvunakool, K.; Bates, R.; Žídek, A.; Potapenko, A.; et al. Highly accurate protein structure prediction with AlphaFold. Nature 2021, 596, 583–589. [Google Scholar] [CrossRef]
  134. Abdoul-Latif, F.M.; Ainane, A.; Mohamed, H.; Ali, A.M.; Aboubaker, I.H.; Jutur, P.P.; Ainane, T. Mpox Resurgence: A Multifaceted Analysis for Global Preparedness. Viruses 2024, 16, 1737. [Google Scholar] [CrossRef]
  135. Škvára, P. Structural Analysis of 2′-O-methyltransferase VP39 Derived from Mpox Virus. Ph.D. Thesis, Charles University, Prague, Czech Republic, 2024. [Google Scholar]
  136. Idisi, I.O.; Oshinubi, K.; Sewanu, V.B.; Yahaya, M.M.; Olagbami, O.S.; Edogbanya, H.O. Investigating Mpox Strain Dynamics Using Computational and Data-Driven Approaches. Viruses 2025, 17, 154. [Google Scholar] [CrossRef]
  137. Jung, Y.; Kwon, S. The Effects of Intellectual Property Rights on Access to Medicines and Catastrophic Expenditure. Int. J. Health Serv. 2015, 45, 507–529. [Google Scholar] [CrossRef]
  138. Tenni, B.; Moir, H.V.J.; Townsend, B.; Kilic, B.; Farrell, A.-M.; Keegel, T.; Gleeson, D. What is the impact of intellectual property rules on access to medicines? A systematic review. Glob. Health 2022, 18, 40. [Google Scholar] [CrossRef] [PubMed]
  139. Shahiwala, A. Advancing drug delivery research: Sustainable strategies for innovation and translation. Drug Deliv. Transl. Res. 2025, 15, 1513–1524. [Google Scholar] [CrossRef] [PubMed]
  140. Saldana, C.S.; Kelley, C.F.; Aldred, B.M.; Cantos, V.D. Mpox and HIV: A Narrative Review. Curr. HIV/AIDS Rep. 2023, 20, 261–269. [Google Scholar] [CrossRef]
  141. Liu, B.M.; Rakhmanina, N.Y.; Yang, Z.; Bukrinsky, M.I. Mpox (Monkeypox) Virus and Its Co-Infection with HIV, Sexually Transmitted Infections, or Bacterial Superinfections: Double Whammy or a New Prime Culprit? Viruses 2024, 16, 784. [Google Scholar] [CrossRef] [PubMed]
  142. Tempestilli, M.; Mondi, A.; D’AVolio, A.; Forini, O.; Pinnetti, C.; Mazzotta, V.; Gagliardini, R.; Beccacece, A.; De Nicolò, A.; Faccendini, P.; et al. Pharmacokinetics of tecovirimat in subjects with Mpox. Int. J. Antimicrob. Agents 2024, 63, 107068. [Google Scholar] [CrossRef]
  143. Hoy, S.M. Tecovirimat: First Global Approval. Drugs 2018, 78, 1377–1382. [Google Scholar] [CrossRef]
  144. Russo, A.T.; Grosenbach, D.W.; Chinsangaram, J.; Honeychurch, K.M.; Long, P.G.; Lovejoy, C.; Maiti, B.; Meara, I.; Hruby, D.E. An overview of tecovirimat for smallpox treatment and expanded anti-orthopoxvirus applications. Expert Rev. Anti Infect. Ther. 2021, 19, 331–344. [Google Scholar] [CrossRef]
  145. Smith, T.G.; Gigante, C.M.; Wynn, N.T.; Matheny, A.; Davidson, W.; Yang, Y.; Condori, R.E.; O’cOnnell, K.; Kovar, L.; Williams, T.L.; et al. Tecovirimat Resistance in Mpox Patients, United States, 2022–2023. Emerg. Infect. Dis. 2023, 29, 2426–2432. [Google Scholar] [CrossRef]
  146. Zucker, J.; Hazra, A.; Titanji, B.K. Mpox and HIV—Collision of two diseases. Curr. HIV/AIDS Rep. 2023, 20, 440–450. [Google Scholar] [CrossRef]
  147. Aden, D.; Zaheer, S.; Kumar, R.; Ranga, S. Monkeypox (Mpox) outbreak during COVID-19 pandemic—Past and the future. J. Med. Virol. 2023, 95, e28701. [Google Scholar] [CrossRef] [PubMed]
  148. Kanoujia, J.; Tarannum, S.; Kaurav, M.; Raina, N.; Jain, K.; Gupta, M. Insight into Recent Updates on Vaccines Development and Immunology of Monkeypox Infection. Curr. Treat. Options Infect. Dis. 2024, 16, 118–128. [Google Scholar] [CrossRef]
  149. Danladi, N.P.; Agboola, P.; Olaniyi, P.; Eze, S.; Oladapo, O.; Obiwulu, D.; Akano, O.S.; Adeola, O.A.; Olawale, K.; Adiatu, A.I.; et al. Challenges in Global Distribution and Equitable Access to Monkeypox Vaccines. Viruses 2024, 16, 1815. [Google Scholar] [CrossRef]
  150. Zinnah, M.A.; Uddin, B.; Hasan, T.; Das, S.; Khatun, F.; Hasan, H.; Udonsom, R.; Rahman, M.; Ashour, H.M. The Re-Emergence of Mpox: Old Illness. Mod. Challenges. Biomed. 2024, 12, 1457. [Google Scholar]
  151. El Dine, F.B.; Gebreal, A.; Samhouri, D.; Estifanos, H.; Kourampi, I.; Abdelrhem, H.; Mostafa, H.A.; Elshaar, A.G.; Suvvari, T.K.; Ghazy, R.M. Ethical considerations during Mpox Outbreak: A scoping review. BMC Med. Ethics 2024, 25, 79. [Google Scholar] [CrossRef]
  152. O’Laughlin, K. Clinical use of tecovirimat (Tpoxx) for treatment of monkeypox under an investigational new drug protocol—United States, May–August 2022. MMWR. Morb. Mortal. Wkly. Rep. 2022, 71, 1190–1195. [Google Scholar] [CrossRef] [PubMed]
  153. Prompetchara, E.; Ketloy, C.; Khawsang, C.; Palaga, T.; Ruxrungtham, K. Mpox global health emergency: Insights into the virus, immune responses, and advancements in vaccines PART II: Insights into the advancements in vaccines. Asian Pac. J. Allergy Immunol. 2024, 42, 191–206. [Google Scholar] [CrossRef]
  154. Musuka, G.; Moyo, E.; Tungwarara, N.; Mhango, M.; Pierre, G.; Saramba, E.; Iradukunda, P.G.; Dzinamarira, T. A critical review of mpox outbreaks, risk factors, and prevention efforts in Africa: Lessons learned and evolving practices. IJID Reg. 2024, 12, 100402. [Google Scholar] [CrossRef] [PubMed]
  155. Witzel, T.C.; Ghobrial, A.; Palich, R.; Charles, H.; Rodger, A.J.; Sabin, C.; Sparrowhawk, A.; Pool, E.R.; Prochazka, M.; Vivancos, R.; et al. Experiences of mpox illness and case management among cis and trans gay, bisexual and other men who have sex with men in England: A qualitative study. EclinicalMedicine 2024, 70, 102522. [Google Scholar] [CrossRef]
  156. Rampogu, S. A review on the use of machine learning techniques in monkeypox disease prediction. Sci. One Health 2023, 2, 100040. [Google Scholar] [CrossRef]
  157. Bamaqa, A.; Bahgat, W.M.; AbdulAzeem, Y.; Balaha, H.M.; Badawy, M.; Elhosseini, M.A. Early detection of monkeypox: Analysis and optimization of pretrained deep learning models using the Sparrow Search Algorithm. Results Eng. 2024, 24, 102985. [Google Scholar] [CrossRef]
  158. Rahim, F.O.; Fallah, M.; Jain, U.; Richardson, E.T.; Ndembi, N.; Ngongo, N.; Kaseya, J. Challenges and Ongoing Actions to Address the Mpox Emergency in Africa. Ann. Glob. Health 2024, 90, 68. [Google Scholar] [CrossRef] [PubMed]
  159. Forni, D.; Cagliani, R.; Molteni, C.; Clerici, M.; Sironi, M. Monkeypox virus: The changing facets of a zoonotic pathogen. Infect. Genet. Evol. 2022, 105, 105372. [Google Scholar] [CrossRef]
  160. Luna, N.; Ramírez, A.L.; Muñoz, M.; Ballesteros, N.; Patiño, L.H.; Castañeda, S.A.; Bonilla-Aldana, D.K.; Paniz-Mondolfi, A.; Ramírez, J.D. Phylogenomic analysis of the monkeypox virus (MPXV) 2022 outbreak: Emergence of a novel viral lineage? Travel Med. Infect. Dis. 2022, 49, 102402. [Google Scholar] [CrossRef]
  161. Lal, H.M.; Uthaman, A.; Thomas, S. Silver Nanoparticle as an Effective Antiviral Agent. In Polymer Nanocomposites Based on Silver Nanoparticles: Synthesis, Characterization and Applications; Lal, H.M., Maria, H.J., Thomas, S., Li, T., Eds.; Springer International Publishing: Cham, Switzerland, 2021; pp. 247–265. [Google Scholar]
  162. Mohamed, N.A.; Zupin, L.; Mazi, S.I.; Al-Khatib, H.A.; Crovella, S. Nanomedicine as a Potential Tool against Monkeypox. Vaccines 2023, 11, 428. [Google Scholar] [CrossRef]
  163. Debnath, J.P.; Hossen, K.; Sayed, S.B.; Khandaker, S.; Dev, P.C.; Sarker, S.; Hossain, T. Identification of potential biomarkers for 2022 Mpox virus infection a transcriptomic network analysis and machine learning approach. Sci. Rep. 2025, 15, 2922. [Google Scholar] [CrossRef]
  164. Chen, Y.; Zhao, R.; Hu, X.; Wang, X. The current status and future prospects of CRISPR-based detection of monkeypox virus: A review. Anal. Chim. Acta 2025, 1336, 343295. [Google Scholar] [CrossRef] [PubMed]
  165. Silva, S.; Kohl, A.; Pena, L.; Pardee, K. Clinical and laboratory diagnosis of monkeypox (mpox): Current status and future directions. iScience 2023, 26, 106759. [Google Scholar] [CrossRef]
  166. Klingelhöfer, D.; Braun, M.; Groneberg, D.A.; Brüggmann, D. Global mpox research in the light of the current outbreak: Demands, drivers, and obstacles. Emerg. Microbes Infect. 2023, 12, 2210696. [Google Scholar] [CrossRef]
Figure 1. Schematic representation of the life cycle of MPXV replication in hosts and potential targets for anti-MPXV drugs as adapted from [24] under Creative Commons Attribution 4.0 International License (CC BY 4.0). The MPXV undergoes a complete life cycle, beginning with host cell entry and concluding with excretion. Both extracellular enveloped virions (EEV) and intracellular mature virions (IMV) enter host cells through membrane fusion and endocytosis, using glycosaminoglycans (GAGs) as receptors. IMV particles travel along microtubules to the perinuclear replication factory, where the viral genome templates DNA replication. They then envelop the virions via the Golgi apparatus to form intracellular enveloped virions (IEVs), which they transport to the cell surface along actin filaments or microtubules. Structurally and functionally, IMVs and EEVs differ: IMVs possess a single membrane, making them more robust and resistant to external damage, yet more detectable by the immune system. In contrast, EEVs, with their double membrane, are more effective at spreading within the host. Cholesterol plays a critical role in viral entry by stabilizing lipid rafts on the host cell membrane. Disrupting these rafts using amphotericin B or cholesterol-lowering drugs such as statins and PCSK9 inhibitors can inhibit viral entry. In addition, marine sulfated polysaccharides, which mimic GAGs, can block viral attachment and entry.
Figure 1. Schematic representation of the life cycle of MPXV replication in hosts and potential targets for anti-MPXV drugs as adapted from [24] under Creative Commons Attribution 4.0 International License (CC BY 4.0). The MPXV undergoes a complete life cycle, beginning with host cell entry and concluding with excretion. Both extracellular enveloped virions (EEV) and intracellular mature virions (IMV) enter host cells through membrane fusion and endocytosis, using glycosaminoglycans (GAGs) as receptors. IMV particles travel along microtubules to the perinuclear replication factory, where the viral genome templates DNA replication. They then envelop the virions via the Golgi apparatus to form intracellular enveloped virions (IEVs), which they transport to the cell surface along actin filaments or microtubules. Structurally and functionally, IMVs and EEVs differ: IMVs possess a single membrane, making them more robust and resistant to external damage, yet more detectable by the immune system. In contrast, EEVs, with their double membrane, are more effective at spreading within the host. Cholesterol plays a critical role in viral entry by stabilizing lipid rafts on the host cell membrane. Disrupting these rafts using amphotericin B or cholesterol-lowering drugs such as statins and PCSK9 inhibitors can inhibit viral entry. In addition, marine sulfated polysaccharides, which mimic GAGs, can block viral attachment and entry.
Scipharm 93 00051 g001
Figure 2. Proposed translational pipeline of in silico drug design methods. Figure created by the authors as adapted from the cited sources [54,55].
Figure 2. Proposed translational pipeline of in silico drug design methods. Figure created by the authors as adapted from the cited sources [54,55].
Scipharm 93 00051 g002
Table 1. Comparative landscape of technological approaches used in Mpox (MPXV) therapeutic discovery.
Table 1. Comparative landscape of technological approaches used in Mpox (MPXV) therapeutic discovery.
Technological ApproachWhat It Does (Scope)/Typical InputsPrimary Readouts/OutputsKey Strengths/Principal LimitationsValidation/Next StepsExample MPXV ApplicationsBest Used for/Pipeline StageRefs
Structure-based in silico dockingPredicts ligand binding poses/affinities to viral or host targets/3D target structures (experimental or predicted), prepared ligand librariesDocking scores, predicted poses, interaction mapsFast, low cost; screens 104–106 compounds; hypothesis generation for SAR/Dependent for structure quality and protonation; ignores full dynamics/solvent; risk of false positivesRe-dock known ligands; MM/GBSA rescoring; move hits to MD, biophysics (isothermal titration calorimetry and surface plasmon resonance), and cell assaysDocking to MPXV enzymes (e.g., VP39 cap MTase, D1/D12 polymerase complex) and host JAK1; prioritization of nitroxoline, atovaquone, tilorone, ZINC leadsEarly hit identification and triage[60]
Molecular dynamics (MD) simulationsTests the stability and dynamics of protein–ligand complexes; refines docking/protein–ligand complexes, force fields, explicit solvent/ionsRoot Mean Square Deviation/Root Mean Square Fluctuation, H-bond occupancy, binding free energy (e.g., MM/PBSA), conformational ensemblesCaptures flexibility, water networks; filters docking artifacts; supports binding hypotheses/Compute-intensive; sensitive to parameters/timescale; not a direct measure of potencyOrthogonal biophysics; mutagenesis; enzymatic half-maximal inhibitory concentration (IC50/EC50)/Ki; cell-based EC50Stability of candidate inhibitors in MPXV VP39 pocket; MD of host-targeted the Janus kinase-signal transducer and activator of transcription (JAK/STAT) modulatorsPost-docking refinement; pre-experimental risk reduction[58]
AI-driven structure prediction (e.g., AlphaFold/ColabFold)Predicts protein 3D structures and complexes; maps pockets/amino-acid sequences, multiple sequence alignments, co-evolution data3D models with confidence metrics (predicted local distance difference test, predicted aligned error), interface predictionsEnables targets lacking structures; rapid; broad proteome coverage/Accuracy varies for flexible regions/complexes; requires experimental validationCryo-EM/X-ray/Nuclear Magnetic Resonance confirmation; benchmarking via known domainsModels for MPXV proteins without PDB structures to enable docking/MDTarget Enablement and Pocket Discovery[61]
High-throughput screening (HTS)Empirical activity screening across large libraries/cell-based infection assays or enzymatic assays; 103–106 compoundsHit rates, EC50/IC50, 50% cytotoxic concentration (CC50), selectivity index (SI)Direct activity readout; unbiased mechanism of action (MoA) discovery; scalable robotics/costly infrastructure; false positives; assay interference; needs robust BSL-2/3 modelsHit confirmation, counter-screens, MoA deconvolution, medicinal chemistryCell-based screens of FDA libraries against orthopox/MPXV surrogates; identification of DNA synthesis and egress blockersPrimary empirical discovery; lead finding[62]
Cryo-electron microscopy (Cryo-EM)Determines the near-native structures of large assemblies/Purified proteins/complexes/virions, vitrification3–5 Å (or better) maps; atomic models; ligand densityCaptures native states and complexes; ideal for large pox proteins/assemblies/high cost; expertise; not high throughputFunctional assays; docking/MD guided by EM maps; fragment campaignsStructural analysis of poxvirus polymerase and capping machinery; guide structure-based designStructure determination; hit-to-lead optimization[63]
Phenotypic assays (infection-based)Measure the antiviral effect in relevant biology without prior target/Live virus or pseudotyped systems; human cell lines/organoidsViral load reduction (quantitative polymerase chain reaction (qPCR)/plaques), CPE rescue, imaging, SIPhysiological relevance; reveals host-directed MoAs; detects polypharmacology/Target unknown; off-target/cytotoxicity risk; model dependenceTarget deconvolution (proteomics/CRISPR), pharmacokinetic pharmacodynamic (PK/PD), in vivoMPXV infection assays in primary keratinocytes and organoids; validation of tecovirimat alternativesHit validation, MoA discovery, prioritization[13,34]
Multi-omics integration (genomics/proteomics/transcriptomics/metabolomics)Maps virus–host networks; identifies targets/pathways and biomarkers/bulk/single-cell RNA-seq, proteomics, phospho-proteomics, metabolomicsDifferential pathways, network hubs, drug–gene signaturesSystems-level insight; reveals host targets and combinations/Complex analysis; batch effects; causal inference is hardNetwork pharmacology, CRISPR perturbations, small-molecule probingHost interactome proximity analyses suggesting JAK/STAT and NF-κB modulators; signatures guiding baricitinib/infliximab hypothesesTarget nomination; combination strategy design[64]
AI/ML-guided repurposing (knowledge graphs, signature matching)Prioritizes candidates using literature, omics, and chemistry graphs/Curated corpora, drug–target networks, and disease signaturesRanked candidates, mechanism hypotheses, polypharmacy suggestionsLeverages existing data; scalable; uncovers non-obvious links/Data bias; spurious correlations; requires wet-lab confirmationProspective validation in phenotypic assays and animal modelsPrioritized immunomodulators (such as JAK inhibitors and anti-TNF agents) and DNA metabolism inhibitors for MPXVHypothesis generation; portfolio triage[65,66]
Organoid/3D tissue modelsHuman-relevant platforms for efficacy/toxicity/skin/mucosal organoids, immune co-culturesViral replication kinetics, barrier integrity, cytokine profilingCloser to human physiology; detects tissue-specific effects/Throughput lower than 2D; cost; standardizationBridge to in vivo; PK/PD translation; safety pharmacologyMPXV replication and drug testing in skin organoids relevant to lesion tropismPreclinical validation; safety/efficacy translation[67]
Animal models (marmoset, prairie dog, and mouse with VACV/ectromelia models)In vivo efficacy and safety assessment/infected animals; candidate dosingSurvival, lesion burden, viral load, PK/PDIntegrates immunity/ADME; regulatory credibility/species differences; ethics; BSL-3; costDose optimization, tox studies; clinical trial designBenchmarking tecovirimat; testing host-directed combinationsLate preclinical go/no-go[68]
Table 2. Repurposed Drug Candidates for MPXV Therapy identified through in silico methods.
Table 2. Repurposed Drug Candidates for MPXV Therapy identified through in silico methods.
Drug NameOriginal IndicationTarget Protein in MPXVMechanism of ActionComputational Method UsedIn Vitro/In Vivo EvidencePotential BenefitsLimitations/ChallengesRefs
AdalimumabRheumatoid arthritisTNF-α (host)Reduces inflammation and viral pathogenesisNetwork-based approachComputational analysis suggests anti-inflammatory benefitWell-studied immune modulatorMay not directly inhibit viral replication[38,89]
AtovaquoneAntimalarialD13L capsid proteinInhibits viral assembly by targeting structural proteinsMolecular docking, MD simulationsIn vitro inhibition of poxvirus replicationFDA-approved, well-toleratedRequires further clinical validation[90]
BaricitinibRheumatoid arthritisJAK1/JAK2 (host pathway)Reduces hyperinflammatory responseNetwork-based pharmacologyIdentified via AI; suppresses MPXV-driven inflammationFDA-approved immunomodulatorIndirect antiviral activity; risk of immunosuppression[38]
BatefenterolChronic Obstructive Pulmonary Disease (COPD) (β-agonist)Unknown (Host target)Modulates immune and inflammatory responseAI-based drug screeningSuggested in silico as host-modulatory agentPotential host-targeted antiviral strategyNo experimental validation yet[13]
BatefenterolCOPDDNA-dependent RNA polymeraseInhibits viral transcription machineryMolecular docking, MD simulationsPotential activity suggested by computational studiesHigh specificity, existing safety profileLacks direct antiviral validation[13]
BrincidofovirCytomegalovirus, AdenovirusDNA Polymerase (D5R)Inhibits viral DNA polymerizationVirtual screening, dockingIn vivo studies show partial MPXV inhibitionOral bioavailability; lipid-modified for uptakeGastrointestinal toxicity; mixed clinical efficacy[62]
BurixaforStem cell mobilizerDNA-dependent RNA polymeraseInhibits viral transcriptionMolecular docking, MD simulationsPredicted inhibition in in silico modelsPotential for rapid repurposingNo in vivo validation yet[13]
CidofovirAntiviral (CMV)Viral DNA polymeraseInhibits viral DNA synthesisMolecular docking, MD simulationsEffective against MPXV in vitro and in vivoEstablished antiviral, broad-spectrumHigh nephrotoxicity risk[62,91]
Deoxyuridine AnalogsHerpesvirus infectionsDNA Polymerase (D5R)Inhibits viral DNA elongationPharmacophore modeling, MD simulationsStrong docking scores & molecular stabilityPotential for combination therapyNo in vivo validation yet[62]
DoxorubicinChemotherapyViral DNA polymeraseInhibits viral DNA replicationMolecular docking, MD simulationsShows inhibition in computational and preliminary in vitro studiesExisting FDA approval, known safety profileHigh cytotoxicity limits therapeutic window[92,93]
EluxadolineIBS treatmentDNA-dependent RNA polymeraseBlocks viral transcriptionMolecular docking, MD simulationsIdentified via computational screeningFDA-approved, potential oral formulationLacks clinical validation[13,94]
ElvitegravirHIV integrase inhibitorVP39 / DNA replication complex (putative)Binds nucleic-acid–processing pocketsDocking (AutoDock/Vina), ADMET filtersNone for MPXV wet labKnown human PK, safetyOff-target risk; needs efficacy data[61]
EtanerceptAutoimmune disordersTNF-α (host)Prevents cytokine overproductionNetwork-based approachIdentified as a potential adjunctive therapyCould reduce MPXV disease severityMay impair immune response to infection[38]
FostamatinibSYK inhibitor (ITP)VP39 (mRNA cap 2-O-MTase); host SYK pathwayInterference with RNA capping; host immune modulationDocking to VP39 (PDB 8CEQ), MD refinementNone specific to MPXV in cited papersOral, known safety profileOff-target immunomodulation; needs MPXV validation[34,37]
InfliximabAutoimmune diseases (TNF-α inhibitor)TNF-α (Host pathway)Suppresses immune hyperactivationProtein-ligand interaction networksReduces cytokine storm in MPXV casesPrevents immune overactivationPotential immunosuppressive side effects[38]
InfliximabAutoimmune disordersTNF-α (host)Modulates immune response, reducing viral pathogenesisNetwork-based approachPotential benefit in controlling cytokine storm in MPXV casesImmunomodulatory effectsRisk of immune suppression[38]
MefloquineAntimalarialD13L capsid proteinDisrupts viral assembly and replicationMolecular docking, MD simulationsEffective against poxviruses in preclinical studiesLong half-life, immune-modulatory propertiesCNS side effects limit broad use[92,95]
Methisazone (Marboran)Historical anti-poxHost translation (eIF-2–dependent early protein synthesis)Inhibits early viral protein synthesis(historic pharmacology)Historical prophylaxis/limited efficacy vs. variola/vaccinia in older reports; not used currentlyMechanistic precedent for pox antiviralsPoor efficacy/tolerability; obsolete[96]
MinocyclineAntibioticDNA-dependent RNA polymeraseInhibits viral transcriptionMolecular docking, MD simulationsComputational predictions suggest strong inhibitionFDA-approved, broad antiviral potentialRequires further in vivo validation[56]
NiclosamideAnthelminticVP37/F13 (envelopment/egress)Inhibits virion egress by binding F13 pocket (putative)Structure-based docking to F13/VP37No MPXV wet-lab signal reported in cited workOral, generic, broad antiviral reportsPoor solubility; primarily computational evidence here[34]
NitroxolineUrinary tract infections (UTI)Thymidylate kinase (TMPK)Blocks DNA synthesis and viral replicationAI-driven screening, QSARIn vitro antiviral activity confirmedFDA-approved; broad-spectrum activityRequires systemic efficacy validation[97]
RemdesivirRdRp inhibitor (Ebola/COVID-19)MPXV DNA polymerase complex (off-target)Nucleoside analogue mismatch (unlikely optimal for DNA viruses)Docking screensWeak/variable activity vs. orthopox in vitro; no MPXV clinical signalKnown safety; IVLimited mechanism fit for DNA poxviruses[41,98]
RibavirinBroad-spectrum antiviralIMPDH/guanosine pools; viral polymerasesDepletes GTP; error catastrophe (RNA-virus-centric)(mechanistic repurposing)Limited/variable anti-orthopox effects in vitro; toxicity at doses requiredOral option; cheapHemolysis; limited pox efficacy[98]
Tecovirimat (ST-246)Smallpox (Orthopoxviruses)F13L (Envelope protein)Inhibits viral egress and prevents virion maturationMolecular docking, MD simulationsFDA-approved; in vitro MPXV efficacy confirmedSpecific to orthopoxviruses; FDA-approvedResistance mutations (F13L escape variants)[99]
Tigecycline/Omadacycline Tetracycline antibioticMulti-target: thymidylate kinase, DNA topoisomerase I, F13 (p37)Binds nucleotide/ATP pockets; blocks DNA synthesis/egress (putative)Multi-target docking, MM-GBSA rescoring, MDNo MPXV wet lab in cited studyMulti-site binding; favorable docking energiesIV use; antibacterial AEs; translational gap[56]
TiloroneAntiviral (Influenza)Viral helicaseDisrupts viral genome processingMolecular docking, MD simulationsComputationally predicted to inhibit MPXV replicationBroad-spectrum antiviral, immune modulatingRequires further in vivo and clinical testing[100]
ZINC22060520Not approved (In silico lead)JAK1Targets host immune pathwaysPharmacophore modeling, MD simulationsStrong binding affinity in MD simulationsNovel candidate with high specificityNo clinical validation yet[38]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Mushebenge, A.G.-A.; Mphuthi, D.D. Deciphering Drug Repurposing Strategies: Antiviral Properties of Candidate Agents Against the Mpox Virus. Sci. Pharm. 2025, 93, 51. https://doi.org/10.3390/scipharm93040051

AMA Style

Mushebenge AG-A, Mphuthi DD. Deciphering Drug Repurposing Strategies: Antiviral Properties of Candidate Agents Against the Mpox Virus. Scientia Pharmaceutica. 2025; 93(4):51. https://doi.org/10.3390/scipharm93040051

Chicago/Turabian Style

Mushebenge, Aganze Gloire-Aimé, and David Ditaba Mphuthi. 2025. "Deciphering Drug Repurposing Strategies: Antiviral Properties of Candidate Agents Against the Mpox Virus" Scientia Pharmaceutica 93, no. 4: 51. https://doi.org/10.3390/scipharm93040051

APA Style

Mushebenge, A. G.-A., & Mphuthi, D. D. (2025). Deciphering Drug Repurposing Strategies: Antiviral Properties of Candidate Agents Against the Mpox Virus. Scientia Pharmaceutica, 93(4), 51. https://doi.org/10.3390/scipharm93040051

Article Metrics

Back to TopTop