Next Article in Journal
The Intestinal Microbiome after Traumatic Injury
Next Article in Special Issue
Critical Review of Tuberculosis Diagnosis in Children from Papua New Guinea Presenting to Health Facilities in the Torres Strait Islands, Australia
Previous Article in Journal
Biochemical, Antioxidant Properties and Antimicrobial Activity of Steno-Endemic Origanum onites
Previous Article in Special Issue
Machine Learning of the Whole Genome Sequence of Mycobacterium tuberculosis: A Scoping PRISMA-Based Review
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Use of Particulate Systems for Tuberculosis Prophylaxis and Treatment: Opportunities and Challenges

by
Alejandra Barrera-Rosales
1,
Romina Rodríguez-Sanoja
2,
Rogelio Hernández-Pando
3 and
Silvia Moreno-Mendieta
2,4,*
1
Posgrado en Ciencias Bioquímicas, Universidad Nacional Autónoma de México (UNAM), A.P. 70228, Ciudad Universitaria, Ciudad de México 04510, México
2
Instituto de Investigaciones Biomédicas, Universidad Nacional Autónoma de México (UNAM), A.P. 70228, Ciudad Universitaria, Ciudad de México 04510, México
3
Sección de Patología Experimental, Instituto Nacional de Ciencias Médicas y Nutrición Salvador Zubirán, Vasco de Quiroga 15, Delegación Tlalpan, Ciudad de México 14080, México
4
CONAHCyT, Instituto de Investigaciones Biomédicas, Universidad Nacional Autónoma de México (UNAM), A.P. 70228, Ciudad Universitaria, Ciudad de México 04510, México
*
Author to whom correspondence should be addressed.
Microorganisms 2023, 11(8), 1988; https://doi.org/10.3390/microorganisms11081988
Submission received: 16 June 2023 / Revised: 28 July 2023 / Accepted: 29 July 2023 / Published: 2 August 2023
(This article belongs to the Special Issue Prevention, Treatment and Diagnosis of Tuberculosis)

Abstract

:
The use of particles to develop vaccines and treatments for a wide variety of diseases has increased, and their success has been demonstrated in preclinical investigations. Accurately targeting cells and minimizing doses and adverse side effects, while inducing an adequate biological response, are important advantages that particulate systems offer. The most used particulate systems are liposomes and their derivatives, immunostimulatory complexes, virus-like particles, and organic or inorganic nano- and microparticles. Most of these systems have been proven using therapeutic or prophylactic approaches to control tuberculosis, one of the most important infectious diseases worldwide. This article reviews the progress and current state of the use of particles for the administration of TB vaccines and treatments in vitro and in vivo, with a special emphasis on polymeric particles. In addition, we discuss the challenges and benefits of using these particulate systems to provide researchers with an overview of the most promising strategies in current preclinical trials, offering a perspective on their progress to clinical trials.

1. Introduction

Tuberculosis (TB), caused by the intracellular bacillus Mycobacterium tuberculosis (Mtb), remains one of the most prevalent infectious diseases, representing the leading cause of death from a single infectious agent until the COVID-19 pandemic. Despite the large global drop in the diagnosis and reporting of cases in the pandemic period, it is estimated that 10.6 million people contracted the disease in 2021, of whom 450,000 were rifampicin-resistant cases, and 1.4 million of people died. Compared with the 10 million cases of TB and the 1.2 million deaths reported in 2019, the worrying increase in cases and the setbacks in the quest to end tuberculosis after COVID-19 are evident [1].
Although this pathogen preferentially generates a pulmonary disease, the infection can be disseminated by generating extrapulmonary TB, such as lymphatic, miliary, and central nervous system TB, which represent approximately 15% of all TB infections and are prevalent in immunocompromised patients [2].
The probability of developing TB disease is much higher among individuals with medical conditions that weaken the immune system, such as HIV/AIDS, diabetes, cancer, renal disease, and severe fungal infections; in individuals who have received organ transplantation or tumor necrosis factor alpha (TNF-α) antagonist therapy; or in individuals that have been exposed to alcohol and tobacco abuse, malnutrition, or air pollution [3,4,5]. Recently, new hypotheses derived from meta-analyses have stated that coinfection with SARS-CoV-2, or the use of drugs to treat it, could accelerate the progression of a preexisting TB infection to pulmonary disease, suggesting that coinfection is a predictor of poor prognosis [6,7].
Adding to the complexity, latent TB infection is recognized as the main source of new TB cases, favoring the prevalence of the disease and its high morbidity and mortality in all countries with a high TB burden. This is not a minor problem if we consider that, according to the latest reports, it is estimated that 23% of the world’s population has a latent infection and the diagnostic methods currently available do not allow us to distinguish between a latent infection and active disease [8].
No less important are the factors that have contributed to the increasing emergence of strains that are resistant to the available antibiotics. Some of these factors include the incomplete and variable protection provided by the existing Bacillus Calmette–Guérin (BCG) vaccines against pulmonary TB [9], late diagnosis, the lack of the timely and proper administration of effective drugs, and extensive treatment regimens that have led to poor patient adherence [5]. Consequently, it is extremely important to develop alternatives that increase or reinforce the protective efficacy of the BCG vaccine [10] and therapeutic alternatives to shorten the treatment timespan and ideally decrease the side effects generated by antibiotics [11].
For TB prevention, the strategies have included the design, development, and evaluation of recombinant BCG strains [12,13,14]; the live-attenuated Mtb strain (MTBVAC) [15,16]; other Mycobacterium strains, such as M. vaccae [17]; subunit recombinant vaccines [18,19,20,21]; and vectorized recombinant vaccines [22,23]. The most recent and promising candidates that have shown evidence of efficacy in animal studies and human trials are summarized in the current TB vaccine pipeline updated as of October 2022 by the TuBerculosis Vaccine Initiative (TBVI) [24]. Treatment strategies have included the use of therapeutic vaccines produced from different Mycobacterium strains, such as M. indicus pranii and M. vaccae [25,26], BCG recombinant strains [12], multiantigenic and multiphasic vectors [27], adjuvanted antigens alone or in combination with nonsteroidal anti-inflammatory drugs (NSAIDs) [28,29], and detoxified cellular fragments of Mtb, such as RUTI [30,31]. Other promising candidates include the use of peptides from different natural sources, such as antimicrobial peptides [32,33] and scorpion venom peptides [34]. Moreover, the use of cytokine gene therapy has also been proven its efficacy and prevented reactivation in experimental TB models [35,36]. No less important are the opportunities that emerge with natural compounds such as flavonoids and lignan aglycones, diferuloylmethane, polyphenols, and aldehydes, among others, which are isolated from plants, fungi, marine species, and bacteria and have shown interesting results alone or in combination with already approved medications [37,38,39].
Among other interesting strategies, the use of particles is becoming more frequent for both TB prophylaxis and therapy [40]. Particles not only protect the molecules that are administered from degradation but also facilitate their controlled and directed release to the target cells, allowing dose optimization [41,42]. They also contribute to overcoming the lack of immunogenicity of subunit vaccines, because many of these particles are inherently immunogenic or can be manipulated to promote enhanced antigenic uptake and processing, mediating adaptive immune responses [43]. As a result of all their properties and benefits, they have been used to formulate vaccines and treatments for TB. Our main objective is to provide a comprehensive perspective on the current state of the preclinical investigation of TB vaccines and treatments formulated with polymeric particles, the challenges and opportunities in the field, and the impact that they could have in future clinical trials.

2. Particulate Systems for the Administration of Vaccines and Therapeutics

Particulate systems are important biotechnological tools that have had an enormous impact on biomedical applications, including basic research, imaging, theranostics, and especially therapeutic or vaccine design and delivery [44,45]. They have sizes ranging from nanometers to micrometers and can be manufactured from inorganic materials (i.e., gold, metal oxides, or silica), synthetic or natural polymers (i.e., aliphatic polyesters or chitosan respectively), or synthetic or natural lipids, among other materials [45,46]. Similar to the materials with which they are manufactured, the methods for their loading and functionalization are varied. Some of these methods include adsorption or immobilization onto the surface, dispersion inside the matrix, linking between the matrix and the bioactive molecule, and encapsulation [47].
The materials and preparation methods of particles define their physicochemical characteristics, such as their size, shape, and charge, which in turn define their biodistribution, targeting, release profiles, toxicity, accumulation time, and clearance [48]. Other properties, such as bioavailability, biodegradability, biocompatibility, and bioadhesiveness, are influenced by the intrinsic properties of the particles and their route of administration [49,50]. Consequently, the main challenge is to reach the best combination of materials to obtain the best particles whose properties guarantee their function and safety in vivo.
Currently, almost all routes of administration can be used to deliver particulate systems, including oral, transdermal, intravenous, subcutaneous, topical, intranasal, and pulmonary routes, the last of which is particularly important for TB treatment and prophylaxis because inhalable formulations are the most effective to induce a memory immune response in the lungs [23,51,52].
In addition, particulate systems can be useful to expand the type of immune response generated, considering that the few currently approved adjuvants are effective in inducing antibody responses but are less successful in inducing cell-mediated immunity, which is very important to eliminate intracellular pathogens such as Mtb [53]. In Figure 1, we consider the advantages and the most common characteristics of particles intended for the nasal and pulmonary administration of vaccines and treatments. In subsequent sections, we address the main factors that justify the use of particles to develop TB vaccines or treatments and summarize the most recent preclinical studies with polymeric formulations.

2.1. Particulate Systems for TB Vaccine Development

Greater comprehension of the roles that immune cells play in response to Mtb infection is of vital importance for the development of vaccines against this pathogen [54]. For many years, exhaustive efforts have been made to modify, improve, or find an alternative to the BCG vaccine [13,55,56]. This vaccine, the only anti-TB vaccine approved in humans, confers effective protection against disseminated and meningeal TB only in children, with variable protection in adults. The factors that mainly affect its protective efficacy include coinfections with viruses or parasites, comorbidities, environmental factors, intrinsic genetic factors of both mycobacteria and humans, and, importantly, the route of vaccination [57,58,59]. After intradermal vaccination, the BCG vaccine interacts with resident epidermal macrophages, whereas Mtb interacts, in most cases, with resident alveolar macrophages (AMs) and does not suffer opsonization. Consequently, antigenic recognition, uptake, processing, and presentation are different, with implications for the induction of the T-cell memory response required for protection against lung disease [57]. This complex situation has justified the administration of the BCG vaccine directly into the respiratory system as a strategy to induce resident memory T cells in the lung [60,61,62] and the exploration of new vaccines against TB that can be administered by nasal or pulmonary routes, which favor the retention of the antigen at mucosal sites, the induction of systemic and mucosal immunity, and, importantly, the development of lung-resident memory T cells. These are important advantages of mucosal vaccination and, of course, an opportunity for the use of particulate systems [63,64].

2.1.1. Immune Activation Induced by Mtb and Particulate Systems

After inhalation, mycobacteria in the deep lung (alveoli) can interact through pattern recognition receptors (PRRs) with AMs and dendritic cells (DCs). Mycobacterial endocytosis leads to the activation and maturation of these cells and the migration of DCs toward the lung-draining lymph nodes for antigenic presentation and the differentiation of T lymphocytes toward a Th1 type. Th1 cells contribute to the elimination of bacilli and create a positive feedback loop by secreting IFN-γ, which in turn activates more macrophages, enhancing the microbicidal response against Mtb by executing functions including the secretion of microbicidal factors and cytokines such as TNF-α [65,66]. In the same way, particles formulated in prophylactic or therapeutic vaccines can also interact with and activate innate immune cells, increasing their mycobactericidal performance to prevent or combat the infection (Figure 2). Particles can also promote endocytosis by professional phagocytes, induce the production of cytokines and microbicidal factors such as nitric oxide and reactive oxygen species (ROS) [67,68], or induce apoptosis and autophagy [69,70,71], which together are very important mechanisms to eliminate bacilli.
Importantly, these particulate formulations can be administered by several routes, such as parenteral, nasal, and pulmonary, protecting the antigen and supplying it to immune cells, and they can also be engineered to have intrinsic immunostimulant activity that increases the microbicidal performance of cells. In such scenarios, they can act as delivery systems, adjuvants, and immunostimulants, simultaneously or separately, which is highly desirable for the formulation of subunit vaccines against TB. For this purpose, the most used nano- and microparticles include natural and synthetic polymeric capsules and spheres (mainly of chitosan and poly(lactide-co-glycolide) (PLGA)) [72], followed by liposomes and derivatives, solid lipid nanoparticles (SLNs), and immune-stimulating complexes (ISCOMs) [73,74,75]. In Figure 3, we summarize the main functions of particles in vaccines against TB depending on their use as delivery systems, adjuvants, or immunostimulants.

2.1.2. In Vitro and In Vivo Evaluation of Particulate TB Vaccines

When particles are added into a vaccine formulation, in addition to antigens and adjuvants, in vitro preclinical studies are necessary to characterize the particles’ properties, their capacity to transport and release antigens, and their stability, safety, and efficacy in the formulation in terms of the immune response induced in cell lines or primary isolates [76,77]. For instance, one of the most complete in vitro characterization studies was carried out on the subunit vaccine candidate ID93 [76,78]. The authors used the recombinant TB antigen ID93 (composed of three immune-dominant antigens and one latency-associated antigen) conjugated to a modified liposome (mGLA-LSQ). This liposome has intrinsic adjuvant properties because it contains the TLR4 agonist glucopyranosyl lipid adjuvant (GLA) and the saponin QS21. The authors demonstrated that the vaccine was stable and bioactive for 3 months, being able to induce the secretion of IL-2, INF-γ, and TNF-α in a cytokine stimulation assay using fresh whole blood from 10 healthy donors [78]. Most of the time, and if the formulation is successful in vitro, the next step is to test it in vivo. These studies are more robust in exploring the immune response generated after administration by different routes and are a requirement to proceed to clinical phase studies. In the last decade, most of the particulate TB vaccine candidates tested have contained polymeric particles that encapsulate, accompany, or present the antigen on their surfaces and have been administered by the parenteral or mucosal routes. In Table 1, we summarize some recent in vivo studies carried out with particulate TB vaccine candidates based on natural and synthetic polymers, showing the scheme of immunization and the immune response induced.
In contrast to the growing number of preclinical phase studies conducted with particulate TB vaccine formulations, progression to clinical phase trials is scarce. Ongoing clinical trials of new TB vaccines were recently reviewed by Saramago et al. [86]. Based on their review, and in our search, only two particulate vaccine candidates have progressed to clinical studies: ID93+GLA-SE and GamTBvac. Coler et al., conducted a randomized, double-blind phase I study in 60 healthy non-TB-exposed non-vaccinated adults. The purpose was to evaluate two dose levels of the ID93 antigen, administered intramuscularly alone or in combination with two different doses of the GLA-SE adjuvant. The vaccine was safe and well tolerated under all regimes and induced antigen-specific IgG responses in subjects that also received the adjuvant. The use of the adjuvant also enhanced the magnitude and cytokine profile of polyfunctional CD4+ T cells [87]. Tkachuk et al., in 2020, conducted a phase II study with 180 healthy volunteers previously vaccinated with BCG and immunized subcutaneously twice at 8-week intervals with their vaccine, GamTBvac. This was a particulate system composed of a multi-antigen fusion protein (the TB antigens Ag85A-ESAT6-CFP10 and a dextran-binding domain) immobilized on dextran NPs and a CpG adjuvant. The vaccine was also safe and well tolerated and induced antigen-specific IFN-γ release, augmented Th1 cytokine-expressing CD4+ T cells, and a higher IgG response in vaccinated subjects [88].

2.2. Disadvantages of Conventional Treatments for TB and Opportunities for Particulate Formulations

After infection, the main objective is to target the mycobacteria that are present inside macrophages, which the immune system is unable to eliminate. It would also be relevant to target the bacteria that are present inside neutrophils or DCs, with therapeutic agents. However, most of the WHO-recommended drugs for TB treatment, which show variable permeability, are administered by oral or intravenous routes, implying that they are present at high concentrations in serum but not in the lungs. This partially explains their lower effectiveness in pulmonary disease treatment and their higher toxicity [89]. Additionally, Mtb not only survives inside the cells but also in the granuloma, the complex multicellular structure formed as a result of the host immune response, and drugs must also permeate these structures and reach the mycobacteria that are contained within them [90]. Importantly, prolonged treatments for drug-susceptible TB (6 months of isoniazid and rifampicin, with the addition of pyrazinamide and ethambutol in the initial 2 months) and drug-resistant TB (between 9 and 24 months depending on the strain) become very toxic, increase secondary adverse effects, and therefore decrease patient adherence to the treatment scheme [91,92].
One way to overcome or at least partially resolve these problems is to use engineered carriers for the directed administration of TB drugs, such as liposomes, solid lipid nanoparticles (SLNs), and polymeric micro- and nanoparticles [89]. The physicochemical characteristics of these particles, mainly but not exclusively the size, surface charge, and functionalization, are crucial in the design and must be considered together with the route of administration. Particularly for the treatment of TB, the inhalation route is of interest to target resident AMs loaded with bacteria. In this regard, several investigations have been developed to find the ideal characteristics that a particle must have to reach this cell population and deliver its load (Figure 1) [93]. Another important advantage of this route of administration is that it mimics the course of bacterial spread: because the AMs are the first cells to phagocytize Mtb and drug-containing particles upon inhalation, they traffic them to the lung interstitium and travel to the site at which the bacteria tend to migrate, which can guarantee the directed and controlled release of anti-TB drugs and, consequently, a more precise dosage with fewer side effects [94].

2.2.1. In Vitro Evaluation of Particulate TB Drug Delivery Systems

In the same way as for evaluating particulate vaccines, in vitro assays are also required for the preclinical investigation of anti-TB treatments formulated with particles. Each study has its limitations and advantages, but they are essential in determining the safety and efficacy of these systems. In vitro studies are also very important to standardize and achieve particles with an optimal aerodynamic diameter for pulmonary delivery, ensuring deposition in the apical and deep regions of the lung [95]. These studies are also critical in characterizing the physicochemical properties, stability, loading efficiency, and release of anti-TB drugs, as exemplified in the works of Garg et al. and Desai et al. [96,97]. Additionally, they allow us to evaluate the phagocytosis of loaded particles, their intracellular accumulation, and their cytotoxicity, because the main objective is to induce lower cytotoxicity than that induced by free drug administration [98,99,100,101,102].
Novel tools such as the in silico stochastic lung model have also been developed to correlate with in vitro studies and to predict the amount of drug deposited quantitatively in the lungs. Mukhtar et al., who reported the fabrication and characterization of a chitosan/hyaluronic acid nanoparticle and isoniazid suspension, predicted, with this model, that a very low fraction of particles was exhaled, while the particle deposition was high in the lung–bronchial and acinar regions, correlating with their in vitro observations [103].
Interestingly, several authors have also evaluated in vitro drug-free microparticles as a strategy to reduce the bacillary load in infected cell lines. For instance, Lawlor et al. reported the use of PLGA particles to reduce the bacillary load in THP-1-derived macrophages infected with the H37Rv strain. Without altering cell viability and without modifying proinflammatory cytokine secretion, they demonstrated that the particles induced NF-kB activation and autophagy in a dose-dependent manner, which in turn increased the killing performance of macrophages [70]. Bai et al., used curcumin particles to treat human alveolar and THP-1-derived macrophages before infection with the H37Rv strain, and curcumin also reduced the bacillary load through the induction of autophagy and caspase-3-dependent apoptosis [69]. Machelart et al., using beta cyclodextrin NPs, demonstrated that they were efficiently captured by bone-marrow-derived macrophages and bone-marrow-derived dendritic cells and were able to impair Mtb replication and induce apoptosis in infected macrophages [71].
In Table 2, we summarize some studies carried out in the last decade with particulate TB drug delivery systems based on natural and synthetic polymers, tested on different cell lines before or after infection with mycobacteria.
Although less frequent, the study of inorganic particles that have a direct microbicidal effect is also of interest. Gold and silver NPs have been functionalized with variable ligands, such as citrate or polyallylamine hydrochloride A, to effectively reduce the cell viability of mycobacteria [113]. The antimycobacterial properties of gold have been supported by auranofin, a gold-based antirheumatic drug that inhibits bacterial thioredoxin reductase, making replicating and nonreplicating mycobacteria susceptible to oxidative species; consequently, gold has become a suitable material to develop particles for the treatment of TB [114,115].

2.2.2. In Vivo Evaluation of Particulate TB Drug Delivery Systems

In vivo studies conducted to evaluate particulate TB drugs are performed with antibiotic-loaded particles (against drug-sensitive and drug-resistant Mtb strains) and are useful in showing prolonged drug release, long-term antibacterial effects, reduced toxicity, and the prevention of infection relapse. There is agreement that, for inhalable formulations, the most appropriate materials are natural or synthetic polymers, and those made from polysaccharides are especially promising. Wu et al. evaluated the in vivo toxicity and release properties of an inhalable preparation of chitosan nanogel particles loaded with genipin, isoniazid, and rifampicin. They demonstrated enhanced antimycobacterial activity in mice infected with the resistant H37Rv strain [116]. Machelart et al., with their beta-cyclodextrin NPs administered by direct aerosolization, were also able to decrease the Mtb burden in the lung after infection, and the authors proposed that this observation was a result of AMs’ reprogramming by these particles, which had intrinsic immunostimulant properties [71].
Grehna et al., showed that after the pulmonary administration of spray-dried locust bean gum MPs loaded with isoniazid and rifabutin, lung infection and mycobacterial growth rate values were decreased in the spleens and livers of infected mice. The short-term treatment regimen (five times per week) that the authors used was more effective than the oral coadministration of both antibiotics, even at lower doses. Additionally, they highlighted that polysaccharide-based particles are promising for pulmonary administration because they contain sugar units that are recognized by surface receptors expressed by AMs [117]. Singh et al. in 2021 also developed a dry powder for inhalation, composed of 25% isoniazid, 25% rifabutin, and 50% biodegradable polymer poly(L-lactide). The authors demonstrated the efficacy, safety, and tolerability of the inhalable particles in three TB models (high-dose intravenous and low-dose aerosol infection in mice and low-dose aerosol infection in guinea pigs). They were also able to prevent the relapse of infection four weeks after stopping the treatment, using the combination strategy of half the oral dose of antibiotics with inhalable particles [118]. Antonov et al. showed that encapsulated levofloxacin in PLGA MPs achieved greater bacterial clearance than the free drug orally administered after infecting mice with the H37Rv strain. The particles demonstrated suitable biocompatibility and release kinetics [119].
In contrast to the growing number of preclinical phase studies conducted with particulate formulations for TB treatment, progression to clinical phase trials is also scarce and, based on our search, there are no polymeric formulations at this stage of investigation. Srichana et al., demonstrated the safety of a dry powder formulation with liposomes containing four anti-tuberculosis drugs (isoniazid, rifampicin, pyrazinamide, and levofloxacin) administered via inhalation to 40 healthy adults. After successfully passing this clinical phase I trial [120], the formulation was evaluated for approximately eight weeks in 44 adult patients with active pulmonary TB. Although the treatment did not increase Mtb sputum culture conversion after two months, the percentage of patients having adverse side effects was significantly lower. The main results were decreased cough at 4 weeks of treatment, substantially reduced hemoptysis at 2 weeks of treatment, and lower incidences of nausea and vomiting [121].

2.2.3. Opportunities for Particulate Systems for TB Theranostics

Recent studies have focused their attention on theranostics as means to combine early diagnosis and the administration of targeted treatments in a single system. Particulate systems applied to TB theranostics must be developed with a favorable aerodynamic diameter for pulmonary delivery, to maximize drug delivery while avoiding toxic systemic side effects and potentially shortening the treatment duration. These systems are composed of a biocompatible metal organic framework (MOF) as a drug carrier, which usually has synergistic therapeutic activity and one or several anti-TB drugs [122]. The MOF delivers its cargo upon activation by endogenous stimuli such as pH, redox, or ATP or by exogenous stimuli such as temperature, ions, pressure, light, humidity, or a magnetic field [123]. Recently, Jiménez-Rodríguez et al. successfully encapsulated RIF in liposomes and silver nanoparticles to develop a luminescent biomarker for its evaluation as a TB theranostic. The particles permitted early diagnosis and treatment, and, due to their optical properties, the authors highlighted their utility in pharmacokinetic studies [124].
An emerging opportunity for TB theranostics is the tracking of complex structures such as granulomas and encapsulating various anti-TB drugs for directed administration. In latent TB that can become active TB, this strategy is a priority because granulomas contribute to the persistence and/or spread of the bacilli present inside them. For this reason, in recent studies, the use of sophisticated systems to localize and treat early granulomas has been explored. Liao et al. designed a TB granuloma imaging-guided photodynamic therapy (PDT) using an aggregation-induced emission carrier. After exposure to white light, the carrier generated ROS and simultaneously released rifampicin. With this system, the authors were able to perform an early diagnosis ex vivo using a granuloma tail model in mice and control the drug-sensitive and drug-resistant bacteria in vitro [125,126]. However, these strategies are in the preliminary stage of investigation and their efficacy and safety levels need to be further studied and characterized.
In Figure 4, we provide an overview of the main advantages, disadvantages, challenges, and opportunities regarding particulate systems for the formulation of TB vaccines or treatments.

3. Concluding Remarks and Prospects

In addition to the challenges involved in combating the immune evasion and resistance mechanisms generated by mycobacteria, challenges related to the development of novel and safe protective and therapeutic alternatives against TB persist. A sophisticated approach that matches the sophisticated evasion mechanisms of Mtb is needed to target infected cells or cells that potentially will be infected. One important approach is the use of particles. They are not only ideal for pulmonary administration, imitating the portal of entry and path of bacilli, but can also be engineered by selecting the most desirable materials, ligands, and physicochemical characteristics depending on whether it is a vaccine or a drug delivery system.
However, the landscape is complicated because few studies have moved on to clinical research phases. This may be because, despite the versatility and advantages of particles, they also have intrinsic limitations, and their performance is affected by external factors, which must be weighed against the benefits, especially regarding to formulating a vaccine or a treatment for a disease of such relevance as TB. Complicating the academic scenario, other associated problems that likely affect translation to clinical practice are related to the lack of clinical phase III trials that evaluate the efficacy and safety of these systems in large populations, the availability of resources and infrastructure for research and development, and the consideration of the health emergency that TB represents globally, especially in countries with a higher prevalence and the presence of hypervirulent multi-drug-resistant strains.
As researchers involved in the field of particulate vaccine development for mucosal administration, we aimed to provide a grounded perspective on the importance of the rational design of these systems, since they are recognized, phagocytosed, and can influence the same cells that the tubercle bacillus interacts with. Moreover, we sought to highlight the value of multidisciplinary work to advance the field and the work that is required to move beyond a proof of concept without losing sight of the challenges imposed by particles and Mtb in its progression.

Author Contributions

All authors made significant contributions to the work reported. Conceptualization and editing, S.M.-M.; research and writing, A.B.-R. and S.M.-M.; critical review of the article, R.R.-S. and R.H.-P. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by grant A1-S-14446 Ciencia Basica (SEP-CONAHCyT).

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. World Health Organization Global Tuberculosis Report 2022; World Health Organization: Geneva, Switzerland, 2022; ISBN 9789240061729.
  2. Rodriguez-Takeuchi, S.Y.; Renjifo, M.E.; Medina, F.J. Extrapulmonary Tuberculosis: Pathophysiology and Imaging Findings. Radiographics 2019, 39, 2023–2037. [Google Scholar] [CrossRef]
  3. Bates, M.; Marais, B.J.; Zumla, A. Tuberculosis Comorbidity with Communicable and Noncommunicable Diseases. Cold Spring Harb. Perspect. Med. 2015, 5, a017889. [Google Scholar] [CrossRef] [Green Version]
  4. Hameed, H.M.A.; Islam, M.M.; Chhotaray, C.; Wang, C.; Liu, Y.; Tan, Y.; Li, X.; Tan, S.; Delorme, V.; Yew, W.W.; et al. Molecular Targets Related Drug Resistance Mechanisms in MDR-, XDR-, and TDR-Mycobacterium tuberculosis Strains. Front. Cell. Infect. Microbiol. 2018, 8, 114. [Google Scholar] [CrossRef]
  5. Singh, R.; Dwivedi, S.P.; Gaharwar, U.S.; Meena, R.; Rajamani, P.; Prasad, T. Recent Updates on Drug Resistance in Mycobacterium tuberculosis. J. Appl. Microbiol. 2019, 128, 1547–1567. [Google Scholar] [CrossRef] [Green Version]
  6. Khurana, A.K.; Aggarwal, D. The (in)Significance of TB and COVID-19 Co-Infection. Eur. Respir. J. 2020, 56, 2002105. [Google Scholar] [CrossRef]
  7. Song, W.M.; Zhao, J.Y.; Zhang, Q.Y.; Liu, S.Q.; Zhu, X.H.; An, Q.Q.; Xu, T.T.; Li, S.J.; Liu, J.Y.; Tao, N.N.; et al. COVID-19 and Tuberculosis Coinfection: An Overview of Case Reports/Case Series and Meta-Analysis. Front. Med. 2021, 8, 657006. [Google Scholar] [CrossRef]
  8. Gong, W.; Wu, X. Differential Diagnosis of Latent Tuberculosis Infection and Active Tuberculosis: A Key to a Successful Tuberculosis Control Strategy. Front. Microbiol. 2021, 12, 745592. [Google Scholar] [CrossRef]
  9. Dockrell, H.M.; Smith, S.G. What Have We Learnt about BCG Vaccination in the Last 20 Years? Front. Immunol. 2017, 8, 1134. [Google Scholar] [CrossRef]
  10. Kaufmann, S.H.E. Vaccine Development Against Tuberculosis Over the Last 140 Years: Failure as Part of Success. Front. Microbiol. 2021, 12, 750124. [Google Scholar] [CrossRef]
  11. Silva, D.R.; de Queiroz Mello, F.C.; Migliori, G.B. Shortened Tuberculosis Treatment Regimens: What Is New? J Bras Pneumol 2020, 46, e20200009. [Google Scholar] [CrossRef]
  12. Hoft, D.F.; Blazevic, A.; Selimovic, A.; Turan, A.; Tennant, J.; Abate, G.; Fulkerson, J.; Zak, D.E.; Walker, R.; McClain, B.; et al. Safety and Immunogenicity of the Recombinant BCG Vaccine AERAS-422 in Healthy BCG-Naïve Adults: A Randomized, Active-Controlled, First-in-Human Phase 1 Trial. EBioMedicine 2016, 7, 278–286. [Google Scholar] [CrossRef] [Green Version]
  13. Nieuwenhuizen, N.E.; Kulkarni, P.S.; Shaligram, U.; Cotton, M.F.; Rentsch, C.A.; Eisele, B.; Grode, L.; Kaufmann, S.H.E. The Recombinant Bacille Calmette-Guérin Vaccine VPM1002: Ready for Clinical Efficacy Testing. Front. Immunol. 2017, 8, 1147. [Google Scholar] [CrossRef]
  14. Carvalho Dos Santos, C.; Rodriguez, D.; Kanno Issamu, A.; Cezar De Cerqueira Leite, L.; Pereira Nascimento, I. Recombinant BCG Expressing the LTAK63 Adjuvant Induces Increased Early and Long-Term Immune Responses against Mycobacteria. Hum. Vaccines Immunother. 2020, 16, 673–683. [Google Scholar] [CrossRef]
  15. Broset, E.; Saubi, N.; Guitart, N.; Aguilo, N.; Uranga, S.; Kilpeläinen, A.; Eto, Y.; Hanke, T.; Gonzalo-Asensio, J.; Martín, C.; et al. MTBVAC-Based TB-HIV Vaccine Is Safe, Elicits HIV-T Cell Responses, and Protects against Mycobacterium tuberculosis in Mice. Mol. Ther. Methods Clin. Dev. 2019, 13, 253–264. [Google Scholar] [CrossRef] [Green Version]
  16. Martín, C.; Marinova, D.; Aguiló, N.; Gonzalo-Asensio, J. MTBVAC, a Live TB Vaccine Poised to Initiate Efficacy Trials 100 Years after BCG. Vaccine 2021, 39, 7277–7285. [Google Scholar] [CrossRef]
  17. Gong, W.P.; Liang, Y.; Ling, Y.B.; Zhang, J.X.; Yang, Y.R.; Wang, L.; Wang, J.; Shi, Y.C.; Wu, X.Q. Effects of Mycobacterium vaccae Vaccine in a Mouse Model of Tuberculosis: Protective Action and Differentially Expressed Genes. Mil. Med. Res. 2020, 7, 25. [Google Scholar] [CrossRef]
  18. Ma, J.; Teng, X.; Wang, X.; Fan, X.; Wu, Y.; Tian, M.; Zhou, Z.; Li, L. A Multistage Subunit Vaccine Effectively Protects Mice Against Primary Progressive Tuberculosis, Latency and Reactivation. EBioMedicine 2017, 22, 143–154. [Google Scholar] [CrossRef] [Green Version]
  19. Tait, D.R.; Hatherill, M.; Van Der Meeren, O.; Ginsberg, A.M.; Van Brakel, E.; Salaun, B.; Scriba, T.J.; Akite, E.J.; Ayles, H.M.; Bollaerts, A.; et al. Final Analysis of a Trial of M72/AS01 E Vaccine to Prevent Tuberculosis. N. Engl. J. Med. 2019, 381, 2429–2439. [Google Scholar] [CrossRef]
  20. Fan, X.; Li, X.; Wan, K.; Zhao, X.; Deng, Y.; Chen, Z.; Luan, X.; Lu, S.; Liu, H. Construction and Immunogenicity of a T Cell Epitope-Based Subunit Vaccine Candidate against Mycobacterium tuberculosis. Vaccine 2021, 39, 6860–6865. [Google Scholar] [CrossRef]
  21. Woodworth, J.S.; Clemmensen, H.S.; Battey, H.; Dijkman, K.; Lindenstrøm, T.; Laureano, R.S.; Taplitz, R.; Morgan, J.; Aagaard, C.; Rosenkrands, I.; et al. A Mycobacterium tuberculosis-Specific Subunit Vaccine That Provides Synergistic Immunity upon Co-Administration with Bacillus Calmette-Guérin. Nat. Commun. 2021, 12, 6658. [Google Scholar] [CrossRef]
  22. Hu, Z.; Jiang, W.; Gu, L.; Qiao, D.; Shu, T.; Lowrie, D.B.; Lu, S.H.; Fan, X.Y. Heterologous Prime-Boost Vaccination against Tuberculosis with Recombinant Sendai Virus and DNA Vaccines. J. Mol. Med. 2019, 97, 1685–1694. [Google Scholar] [CrossRef]
  23. Khan, A.; Sayedahmed, E.E.; Singh, V.K.; Mishra, A.; Dorta-Estremera, S.; Nookala, S.; Canaday, D.H.; Chen, M.; Wang, J.; Sastry, K.J.; et al. A Recombinant Bovine Adenoviral Mucosal Vaccine Expressing Mycobacterial Antigen-85B Generates Robust Protection against Tuberculosis in Mice. Cell Reports Med. 2021, 2, 100372. [Google Scholar] [CrossRef]
  24. TBVI TuBerculosis Vaccine Innitiative. Available online: https://www.tbvi.eu/what-we-do/pipeline-of-vaccines/ (accessed on 16 January 2023).
  25. Sharma, S.K.; Katoch, K.; Sarin, R.; Balambal, R.; Kumar Jain, N.; Patel, N.; Murthy, K.J.R.; Singla, N.; Saha, P.K.; Khanna, A.; et al. Efficacy and Safety of Mycobacterium indicus pranii as an Adjunct Therapy in Category II Pulmonary Tuberculosis in a Randomized Trial. Sci. Rep. 2017, 7, 3354. [Google Scholar] [CrossRef] [Green Version]
  26. Huang, C.Y.; Hsieh, W.Y. Efficacy of Mycobacterium vaccae Immunotherapy for Patients with Tuberculosis: A Systematic Review and Meta-Analysis. Hum. Vaccines Immunother. 2017, 13, 1960–1971. [Google Scholar] [CrossRef] [Green Version]
  27. Leung-Theung-Long, S.; Coupet, C.A.; Gouanvic, M.; Schmitt, D.; Ray, A.; Hoffmann, C.; Schultz, H.; Tyagi, S.; Soni, H.; Converse, P.J.; et al. A Multi-Antigenic MVA Vaccine Increases Efficacy of Combination Chemotherapy against Mycobacterium tuberculosis. PLoS ONE 2018, 13, e0196815. [Google Scholar] [CrossRef] [Green Version]
  28. Day, T.A.; Penn-Nicholson, A.; Luabeya, A.K.K.; Fiore-Gartland, A.; Du Plessis, N.; Loxton, A.G.; Vergara, J.; Rolf, T.A.; Reid, T.D.; Toefy, A.; et al. Safety and Immunogenicity of the Adjunct Therapeutic Vaccine ID93 + GLA-SE in Adults Who Have Completed Treatment for Tuberculosis: A Randomised, Double-Blind, Placebo-Controlled, Phase 2a Trial. Lancet Respir. Med. 2021, 9, 373–386. [Google Scholar] [CrossRef]
  29. Jenum, S.; Tonby, K.; Rueegg, C.S.; Rühwald, M.; Kristiansen, M.P.; Bang, P.; Olsen, I.C.; Sellæg, K.; Røstad, K.; Mustafa, T.; et al. A Phase I/II Randomized Trial of H56:IC31 Vaccination and Adjunctive Cyclooxygenase-2-Inhibitor Treatment in Tuberculosis Patients. Nat. Commun. 2021, 12, 6774. [Google Scholar] [CrossRef]
  30. Prabowo, S.A.; Painter, H.; Zelmer, A.; Smith, S.G.; Seifert, K.; Amat, M.; Cardona, P.J.; Fletcher, H.A. RUTI Vaccination Enhances Inhibition of Mycobacterial Growth ex vivo and Induces a Shift of Monocyte Phenotype in Mice. Front. Immunol. 2019, 10, 894. [Google Scholar] [CrossRef] [Green Version]
  31. Russo, G.; Juarez, M.A.; Cardona, P.J.; Fichera, E.; Pappalardo, F.; Pennisi, M.; Coler, R.; Viceconti, M. Evaluation of the Efficacy of RUTI and ID93/GLA-SE Vaccines in Tuberculosis Treatment: In silico Trial Trhough UISS-TB Simulator. In Proceedings of the 2019 IEEE International Conference on Bioinformatics and Biomedicine (BIBM), San Diego, CA, USA, 18–21 November 2019; pp. 1–23. [Google Scholar]
  32. Portell-Buj, E.; Vergara, A.; Alejo, I.; López-Gavín, A.; Rosa Montè, M.; San Nicolàs, L.; Gonzàlez-Martín, J.; Tudó, G. In vitro Activity of 12 Antimicrobial Peptides against Mycobacterium tuberculosis and Mycobacterium avium Clinical Isolates. J. Med. Microbiol. 2019, 68, 211–215. [Google Scholar] [CrossRef]
  33. Peláez Coyotl, E.A.; Palacios, J.B.; Muciño, G.; Moreno-Blas, D.; Costas, M.; Montes, T.M.; Diener, C.; Uribe-Carvajal, S.; Massieu, L.; Castro-Obregón, S.; et al. Antimicrobial Peptide against Mycobacterium tuberculosis That Activates Autophagy Is an Effective Treatment for Tuberculosis. Pharmaceutics 2020, 12, 1071. [Google Scholar] [CrossRef]
  34. Carcamo-Noriega, E.N.; Sathyamoorthi, S.; Banerjee, S.; Gnanamani, E.; Mendoza-Trujillo, M.; Mata-Espinosa, D.; Hernández-Pando, R.; Veytia-Bucheli, J.I.; Possani, L.D.; Zare, R.N. 1,4-Benzoquinone Antimicrobial Agents against Staphylococcus aureus and Mycobacterium tuberculosis Derived from Scorpion Venom. Proc. Natl. Acad. Sci. USA 2019, 116, 12642–12647. [Google Scholar] [CrossRef] [Green Version]
  35. Francisco-Cruz, A.; Mata-Espinosa, D.; Ramos-Espinosa, O.; Marquina-Castillo, B.; Estrada-Parra, S.; Xing, Z.; Hernández-Pando, R. Efficacy of Gene-Therapy Based on Adenovirus Encoding Granulocyte-Macrophage Colony-Stimulating Factor in Drug-Sensitive and Drug-Resistant Experimental Pulmonary Tuberculosis. Tuberculosis 2016, 100, 5–14. [Google Scholar] [CrossRef] [PubMed]
  36. Ramos-Espinosa, O.; Hernández-Bazán, S.; Francisco-Cruz, A.; Mata-Espinosa, D.; Barrios-Payán, J.; Marquina-Castillo, B.; López-Casillas, F.; Carretero, M.; del Río, M.; Hernández-Pando, R. Gene Therapy Based in Antimicrobial Peptides and Proinflammatory Cytokine Prevents Reactivation of Experimental Latent Tuberculosis. Pathog. Dis. 2016, 74, ftw075. [Google Scholar] [CrossRef] [PubMed]
  37. Rodríguez-Flores, E.M.; Mata-Espinosa, D.; Barrios-Payan, J.; Marquina-Castillo, B.; Castañón-Arreola, M.; Hernández-Pando, R. A Significant Therapeutic Effect of Silymarin Administered Alone, or in Combination with Chemotherapy, in Experimental Pulmonary Tuberculosis Caused by Drug-Sensitive or Drugresistant Strains: In vitro and in vivo Studies. PLoS ONE 2019, 14, e0217457. [Google Scholar] [CrossRef] [PubMed]
  38. Maiolini, M.; Gause, S.; Taylor, J.; Steakin, T.; Shipp, G.; Lamichhane, P.; Deshmukh, B.; Shinde, V.; Bishayee, A.; Deshmukh, R.R. The War against Tuberculosis: A Review of Natural Compounds and Their Derivatives. Molecules 2020, 25, 3011. [Google Scholar] [CrossRef]
  39. Cazzaniga, G.; Mori, M.; Chiarelli, L.R.; Gelain, A.; Meneghetti, F.; Villa, S. Natural Products against Key Mycobacterium tuberculosis Enzymatic Targets: Emerging Opportunities for Drug Discovery. Eur. J. Med. Chem. 2021, 224, 113732. [Google Scholar] [CrossRef] [PubMed]
  40. Patil, K.; Bagade, S.; Bonde, S.; Sharma, S.; Saraogi, G. Recent Therapeutic Approaches for the Management of Tuberculosis: Challenges and Opportunities. Biomed. Pharmacother. 2018, 99, 735–745. [Google Scholar] [CrossRef] [PubMed]
  41. Nasiruddin, M.; Neyaz, M.K.; Das, S. Nanotechnology-Based Approach in Tuberculosis Treatment. Tuberc. Res. Treat. 2017, 2017, 4920209. [Google Scholar] [CrossRef] [Green Version]
  42. Gordillo-Galeano, A.; Ospina-Giraldo, L.F.; Mora-Huertas, C.E. Lipid Nanoparticles with Improved Biopharmaceutical Attributes for Tuberculosis Treatment. Int. J. Pharm. 2021, 596, 120321. [Google Scholar] [CrossRef]
  43. Chen, S.; Quan, D.H.; Wang, X.T.; Sandford, S.; Kirman, J.R.; Britton, W.J.; Rehm, B.H.A. Particulate Mycobacterial Vaccines Induce Protective Immunity against Tuberculosis in Mice. Nanomaterials 2021, 11, 2060. [Google Scholar] [CrossRef]
  44. Wallis, J.; Shenton, D.P.; Carlisle, R.C. Novel Approaches for the Design, Delivery and Administration of Vaccine Technologies. Clin. Exp. Immunol. 2019, 196, 189–204. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Batty, C.J.; Bachelder, E.M.; Ainslie, K.M. Historical Perspective of Clinical Nano and Microparticle Formulations for Delivery of Therapeutics. Trends Mol. Med. 2021, 27, 516–519. [Google Scholar] [CrossRef]
  46. Zaheer, T.; Pal, K.; Zaheer, I. Topical Review on Nano-Vaccinology: Biochemical Promises and Key Challenges. Process Biochem. 2020, 100, 237–244. [Google Scholar] [CrossRef]
  47. Kheirollahpour, M.; Mehrabi, M.; Dounighi, N.M.; Mohammadi, M.; Masoudi, A. Nanoparticles and Vaccine Development. Pharm. Nanotechnol. 2020, 8, 6–21. [Google Scholar] [CrossRef]
  48. Zhao, Z.; Ukidve, A.; Krishnan, V.; Mitragotri, S. Effect of Physicochemical and Surface Properties on in vivo Fate of Drug Nanocarriers. Adv. Drug Deliv. Rev. 2019, 143, 3–21. [Google Scholar] [CrossRef] [PubMed]
  49. Chenthamara, D.; Subramaniam, S.; Ramakrishnan, S.G.; Krishnaswamy, S.; Essa, M.M.; Lin, F.; Qoronfleh, M.W. Therapeutic Efficacy of Nanoparticles and Routes of Administration. Biomater. Res. 2019, 21, 20. [Google Scholar] [CrossRef]
  50. Vasquez-Martínez, N.; Guillen, D.; Moreno-Mendieta, S.A.; Sanchez, S.; Rodríguez-Sanoja, R. The Role of Mucoadhesion and Mucopenetration in the Immune Response Induced by Polymer-Based Mucosal Adjuvants. Polymers 2023, 15, 1615. [Google Scholar] [CrossRef] [PubMed]
  51. Nagpal, P.S.; Kesarwani, A.; Sahu, P.; Upadhyay, P. Aerosol Immunization by Alginate Coated Mycobacterium (BCG/MIP) Particles Provide Enhanced Immune Response and Protective Efficacy than Aerosol of Plain Mycobacterium against M.tb. H37Rv Infection in Mice. BMC Infect. Dis. 2019, 19, 568. [Google Scholar] [CrossRef]
  52. Gomez, M.; Archer, M.; Barona, D.; Wang, H.; Ordoubadi, M.; Bin Karim, S.; Carrigy, N.B.; Wang, Z.; McCollum, J.; Press, C.; et al. Microparticle Encapsulation of a Tuberculosis Subunit Vaccine Candidate Containing a Nanoemulsion Adjuvant via Spray Drying. Eur. J. Pharm. Biopharm. 2021, 163, 23–37. [Google Scholar] [CrossRef]
  53. Amini, Y.; Moradi, B.; Fasihi-Ramandi, M. Aluminum Hydroxide Nanoparticles Show Strong Activity to Stimulate Th-1 Immune Response against Tuberculosis. Artif. Cells Nanomed. Biotechnol. 2017, 45, 1331–1335. [Google Scholar] [CrossRef] [Green Version]
  54. Sia, J.K.; Rengarajan, J. Immunology of Mycobacterium tuberculosis Infections. Microbiol. Spectr. 2019, 7, 3–22. [Google Scholar] [CrossRef] [PubMed]
  55. Zhang, Y.; Yang, J.; Bai, G. Cyclic Di-AMP-Mediated Interaction between Mycobacterium tuberculosis ΔcnpB and Macrophages Implicates a Novel Strategy for Improving BCG Vaccination. Pathog. Dis. 2018, 76, fty008. [Google Scholar] [CrossRef] [PubMed]
  56. Cotton, M.F.; Madhi, S.A.; Luabeya, A.K.; Tameris, M.; Hesseling, A.C.; Shenje, J.; Schoeman, E.; Hatherill, M.; Desai, S.; Kapse, D.; et al. Safety and Immunogenicity of VPM1002 versus BCG in South African Newborn Babies: A Randomised, Phase 2 Non-Inferiority Double-Blind Controlled Trial. Lancet Infect. Dis. 2022, 22, 1472–1483. [Google Scholar] [CrossRef] [PubMed]
  57. Moliva, J.I.; Turner, J.; Torrelles, J.B. Immune Responses to Bacillus Calmette-Guérin Vaccination: Why Do They Fail to Protect against Mycobacterium tuberculosis? Front. Immunol. 2017, 8, 407. [Google Scholar] [CrossRef] [Green Version]
  58. Orgeur, M.; Brosch, R. Evolution of Virulence in the Mycobacterium tuberculosis Complex. Curr. Opin. Microbiol. 2018, 41, 68–75. [Google Scholar] [CrossRef]
  59. Boahen, C.K.; Moorlag, S.J.C.F.M.; Jensen, K.J.; Matzaraki, V.; Fanucchi, S.; Monteiro, I.; de Bree, C.; Fok, E.T.; Mhlanga, M.; Joosten, L.A.B.; et al. Genetic Regulators of Cytokine Responses upon BCG Vaccination in Children from West Africa. J. Genet. Genom. 2023, 50, 434–446. [Google Scholar] [CrossRef]
  60. Perdomo, C.; Zedler, U.; Kühl, A.A.; Lozza, L.; Saikali, P.; Sander, L.E.; Vogelzang, A.; Kaufmann, S.H.E.; Kupz, A. Mucosal BCG Vaccination Induces Protective Lung-Resident Memory T Cell Populations against Tuberculosis. MBio 2016, 7, e01686-16. [Google Scholar] [CrossRef] [Green Version]
  61. Wang, Y.; Yang, C.; He, Y.; Zhan, X.; Xu, L. Ipr1 Modified BCG as a Novel Vaccine Induces Stronger Immunity than BCG against Tuberculosis Infection in Mice. Mol. Med. Rep. 2016, 14, 1756–1764. [Google Scholar] [CrossRef]
  62. Bull, N.C.; Stylianou, E.; Kaveh, D.A.; Pinpathomrat, N.; Pasricha, J.; Harrington-Kandt, R.; Garcia-Pelayo, M.C.; Hogarth, P.J.; McShane, H. Enhanced Protection Conferred by Mucosal BCG Vaccination Associates with Presence of Antigen-Specific Lung Tissue-Resident PD-1 + KLRG1 − CD4+ T Cells. Mucosal. Immunol. 2019, 12, 555–564. [Google Scholar] [CrossRef] [Green Version]
  63. Counoupas, C.; Ferrell, K.C.; Ashhurst, A.; Bhattacharyya, N.D.; Nagalingam, G.; Stewart, E.L.; Feng, C.G.; Petrovsky, N.; Britton, W.J.; Triccas, J.A. Mucosal Delivery of a Multistage Subunit Vaccine Promotes Development of Lung-Resident Memory T Cells and Affords Interleukin-17-Dependent Protection against Pulmonary Tuberculosis. NPJ Vaccines 2020, 5, 105. [Google Scholar] [CrossRef]
  64. Gomez, M.; McCollum, J.; Wang, H.; Ordoubadi, M.; Jar, C.; Carrigy, N.B.; Barona, D.; Tetreau, I.; Archer, M.; Gerhardt, A.; et al. Development of a Formulation Platform for a Spray-Dried, Inhalable Tuberculosis Vaccine Candidate. Int. J. Pharm. 2021, 593, 120121. [Google Scholar] [CrossRef]
  65. O’Garra, A.; Redford, P.S.; McNab, F.W.; Bloom, C.I.; Wilkinson, R.J.; Berry, M.P.R. The Immune Response in Tuberculosis. Annu. Rev. Immunol. 2013, 31, 475–527. [Google Scholar] [CrossRef]
  66. de Martino, M.; Lodi, L.; Galli, L.; Chiappini, E. Immune Response to Mycobacterium tuberculosis: A Narrative Review. Front. Pediatr. 2019, 7, 350. [Google Scholar] [CrossRef] [Green Version]
  67. Counoupas, C.; Pinto, R.; Nagalingam, G.; Britton, W.J.; Petrovsky, N.; Triccas, J.A. Delta Inulin-Based Adjuvants Promote the Generation of Polyfunctional CD4+ T Cell Responses and Protection against Mycobacterium tuberculosis. Infect. Sci. Rep. 2017, 7, 8582. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  68. Tkachuk, A.P.; Gushchin, V.A.; Potapov, V.D.; Demidenko, A.V.; Lunin, V.G.; Gintsburg, A.L. Multi-Subunit BCG Booster Vaccine GamTBvac: Assessment of Immunogenicity and Protective Efficacy in Murine and Guinea Pig TB Models. PLoS ONE 2017, 12, e0176784. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  69. Bai, X.; Oberley-Deegan, R.E.; Bai, A.; Ovrutsky, A.R.; Kinney, W.H.; Weaver, M.; Zhang, G.; Honda, J.R.; Chan, E.D. Curcumin Enhances Human Macrophage Control of Mycobacterium tuberculosis Infection. Respirology 2016, 21, 951–957. [Google Scholar] [CrossRef] [PubMed]
  70. Lawlor, C.; O’Connor, G.; O’Leary, S.; Gallagher, P.J.; Cryan, S.A.; Keane, J.; O’Sullivan, M.P. Treatment of Mycobacterium tuberculosis-Infected Macrophages with Poly(Lactic-Co-Glycolic Acid) Microparticles Drives NFKB and Autophagy Dependent Bacillary Killing. PLoS ONE 2016, 11, e0149167. [Google Scholar] [CrossRef] [Green Version]
  71. Machelart, A.; Salzano, G.; Li, X.; Demars, A.; Debrie, A.S.; Menendez-Miranda, M.; Pancani, E.; Jouny, S.; Hoffmann, E.; Deboosere, N.; et al. Intrinsic Antibacterial Activity of Nanoparticles Made of β-Cyclodextrins Potentiates Their Effect as Drug Nanocarriers against Tuberculosis. ACS Nano 2019, 13, 3992–4007. [Google Scholar] [CrossRef]
  72. Khademi, F.; Derakhshan, M.; Yousefi-Avarvand, A.; Tafaghodi, M. Potential of polymeric particles as future vaccine delivery systems/adjuvants for parenteral and non-parenteral immunization against tuberculosis: A systematic review. Iran. J. Basic Med. Sci. 2018, 21, 116–123. [Google Scholar] [CrossRef]
  73. Rajput, A.; Mandlik, S.; Pokharkar, V. Nanocarrier-Based Approaches for the Efficient Delivery of Anti-Tubercular Drugs and Vaccines for Management of Tuberculosis. Front. Pharmacol. 2021, 12, 749945. [Google Scholar] [CrossRef]
  74. Duong, V.T.; Skwarczynski, M.; Toth, I. Towards the development of subunit vaccines against tuberculosis: The key role of adjuvant. Tuberculosis 2023, 139, 102307. [Google Scholar] [CrossRef] [PubMed]
  75. Pabreja, S.; Garg, T.; Rath, G.; Goyal, A.K. Mucosal Vaccination against Tuberculosis Using Ag85A-Loaded Immunostimulating Complexes. Artif. Cells Nanomed. Biotechnol. 2016, 44, 532–539. [Google Scholar] [CrossRef] [PubMed]
  76. Kramer, R.M.; Archer, M.C.; Orr, M.T.; Dubois Cauwelaert, N.; Beebe, E.A.; Huang, P.W.D.; Dowling, Q.M.; Schwartz, A.M.; Fedor, D.M.; Vedvick, T.S.; et al. Development of a Thermostable Nanoemulsion Adjuvanted Vaccine against Tuberculosis Using a Design-of-Experiments Approach. Int. J. Nanomed. 2018, 13, 3689–3711. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  77. Najafi, A.; Ghazvini, K.; Sankian, M.; Gholami, L.; Amini, Y.; Zare, S.; Khademi, F.; Tafaghodi, M. T Helper Type 1 Biased Immune Responses by PPE17 Loaded Core-Shell Alginate-Chitosan Nanoparticles after Subcutaneous and Intranasal Administration. Life Sci. 2021, 282, 119806. [Google Scholar] [CrossRef] [PubMed]
  78. Adeagbo, B.A.; Akinlalu, A.O.; Phan, T.; Guderian, J.; Boukes, G.; Willenburg, E.; Fenner, C.; Bolaji, O.O.; Fox, C.B. Controlled Covalent Conjugation of a Tuberculosis Subunit Antigen (ID93) to Liposome Improved in vitro Th1-Type Cytokine Recall Responses in Human Whole Blood. ACS Omega 2020, 5, 31306–31313. [Google Scholar] [CrossRef]
  79. Meerak, J.; Wanichwecharungruang, S.P.; Palaga, T. Enhancement of Immune Response to a DNA Vaccine against Mycobacterium tuberculosis Ag85B by Incorporation of an Autophagy Inducing System. Vaccine 2013, 31, 784–790. [Google Scholar] [CrossRef]
  80. Ahmed, M.; Jiao, H.; Domingo-Gonzalez, R.; Das, S.; Griffiths, K.; Rangel-Moreno, J.; Nagarajan, U.; Khader, S. Rationalized Design of a Mucosal Vaccine Protects against Mycobacterium tuberculosis Challenge in Mice. J. Leukoc. Biol. 2017, 101, 1373–1381. [Google Scholar] [CrossRef] [Green Version]
  81. Das, I.; Padhi, A.; Mukherjee, S.; Dash, D.P.; Kar, S.; Sonawane, A. Biocompatible Chitosan Nanoparticles as an Efficient Delivery Vehicle for Mycobacterium tuberculosis Lipids to Induce Potent Cytokines and Antibody Response through Activation of Γδ T Cells in Mice. Nanotechnology 2017, 28, 165101. [Google Scholar] [CrossRef]
  82. Yu, W.; Hu, T. Conjugation with an Inulin-Chitosan Adjuvant Markedly Improves the Immunogenicity of Mycobacterium tuberculosis CFP10-TB10.4 Fusion Protein. Mol. Pharm. 2016, 13, 3626–3635. [Google Scholar] [CrossRef]
  83. Diego-González, L.; Crecente-Campo, J.; Paul, M.J.; Singh, M.; Reljic, R.; Alonso, M.J.; González, Á. Design of Polymeric Nanocapsules for Intranasal Vaccination against Mycobacterium tuberculosis: Influence of the Polymeric Shell and Antigen Positioning. Pharmaceutics 2020, 12, 489. [Google Scholar] [CrossRef]
  84. Khademi, F.; Sahebkar, A.; Fasihi-Ramandi, M.; Taheri, R.A. Induction of Strong Immune Response against a Multicomponent Antigen of Mycobacterium tuberculosis in BALB/c Mice Using PLGA and DOTAP Adjuvant. APMIS 2018, 126, 509–514. [Google Scholar] [CrossRef]
  85. Dalirfardouei, R.; Tafaghodi, M.; Meshkat, Z.; Najafi, A.; Gholoobi, A.; Nabavinia, S.; Sajedifar, S.; Meshkat, M.; Badiee, A.; Ramezani, M. A Novel Formulation of Mtb72F DNA Vaccine for Immunization against Tuberculosis. Iran. J. Basic Med. Sci. 2020, 23, 826–832. [Google Scholar] [CrossRef] [PubMed]
  86. Saramago, S.; Magalhães, J.; Pinheiro, M. Tuberculosis Vaccines: An Update of Recent and Ongoing Clinical Trials. Appl. Sci. 2021, 11, 9250. [Google Scholar] [CrossRef]
  87. Coler, R.; Day, T.; Ellis, R.; Piazza, F.M.; Beckmann, A.; Vergara, J.A.; Rolf, T.; Lu, L.L.; Alter, G.; Hokey, D.; et al. The TLR-4 agonist adjuvant, GLA-SE, improves magnitude and quality of immune responses elicited by the ID93 tuberculosis vaccine: First-in-human trial. NPJ Vaccines 2018, 3, 34. [Google Scholar] [CrossRef] [Green Version]
  88. Tkachuk, A.P.; Bykonia, E.N.; Popova, L.I.; Kleymenov, D.A.; Semashko, M.A.; Chulanov, V.P.; Fitilev, S.B.; Maksimov, S.L.; Smolyarchuk, E.A.; Manuylov, V.A.; et al. Safety and Immunogenicity of the GamTBvac, the Recombinant Subunit Tuberculosis Vaccine Candidate: A Phase II, Multi-Center, Double-Blind, Randomized, Placebo-Controlled Study. Vaccines 2020, 8, 652. [Google Scholar] [CrossRef]
  89. Tan, Z.M.; Lai, G.P.; Pandey, M.; Srichana, T.; Pichika, M.R.; Gorain, B.; Bhattamishra, S.K.; Choudhury, H. Novel Approaches for the Treatment of Pulmonary Tuberculosis. Pharmaceutics 2020, 12, 1196. [Google Scholar] [CrossRef]
  90. Hortle, E.; Oehlers, S.H. Host-Directed Therapies Targeting the Tuberculosis Granuloma Stroma. Pathog. Dis. 2020, 78, ftaa015. [Google Scholar] [CrossRef] [PubMed]
  91. Yang, M.; Pan, H.; Lu, L.; He, X.; Chen, H.; Tao, B.; Liu, W.; Yi, H.; Tang, S. Home-Based Anti-Tuberculosis Treatment Adverse Reactions (HATTAR) Study: A Protocol for a Prospective Observational Study. BMJ Open 2019, 9, e027321. [Google Scholar] [CrossRef] [PubMed]
  92. Sant´Anna, F.M.; Araújo-Pereira, M.; Schmaltz, C.A.S.; Arriaga, M.B.; de Oliveira, R.V.C.; Andrade, B.B.; Rolla, V.C. Adverse Drug Reactions Related to Treatment of Drug-Susceptible Tuberculosis in Brazil: A Prospective Cohort Study. Front. Trop. Dis. 2022, 2, 748310. [Google Scholar] [CrossRef]
  93. Chae, J.; Choi, Y.; Tanaka, M.; Choi, J. Inhalable Nanoparticles Delivery Targeting Alveolar Macrophages for the Treatment of Pulmonary Tuberculosis. J. Biosci. Bioeng. 2021, 132, 543–551. [Google Scholar] [CrossRef]
  94. Cohen, S.B.; Gern, B.H.; Delahaye, J.L.; Adams, K.N.; Courtney, R.; Winkler, J.; Sherman, D.R.; Gerner, M.Y.; Kevin, B. Alveolar Macrophages Provide an Early Mycobacterium tuberculosis Niche and Initiate Dissemination. Cell Host Microbe 2018, 24, 439–446. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Rodrigues, S.; da Costa, A.M.R.; Flórez-Fernández, N.; Torres, M.D.; Faleiro, M.L.; Buttini, F.; Grenha, A. Inhalable Spray-Dried Chondroitin Sulphate Microparticles: Effect of Different Solvents on Particle Properties and Drug Activity. Polymers 2020, 12, 425. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Garg, T.; Goyal, A.K.; Rath, G.; Murthy, R.S.R. Spray-Dried Particles as Pulmonary Delivery System of Anti-Tubercular Drugs: Design, Optimization, in vitro and in vivo Evaluation. Pharm. Dev. Technol. 2016, 21, 951–960. [Google Scholar] [CrossRef]
  97. Desai, S.K.; Mondal, D.; Bera, S. Polyurethane-Functionalized Starch Nanocrystals as Anti-Tuberculosis Drug Carrier. Sci. Rep. 2021, 11, 8331. [Google Scholar] [CrossRef] [PubMed]
  98. Moretton, M.A.; Cagel, M.; Bernabeu, E.; Gonzalez, L.; Chiappetta, D.A. Nanopolymersomes as Potential Carriers for Rifampicin Pulmonary Delivery. Colloids Surf. B Biointerfaces 2015, 136, 1017–1025. [Google Scholar] [CrossRef] [PubMed]
  99. Alves, A.D.; Cavaco, J.S.; Guerreiro, F.; Lourenço, J.P.; Rosa Da Costa, A.M.; Grenha, A. Inhalable Antitubercular Therapy Mediated by Locust Bean Gum Microparticles. Molecules 2016, 21, 702. [Google Scholar] [CrossRef] [Green Version]
  100. Rodrigues, S.; Alves, A.D.; Cavaco, J.S.; Pontes, J.F.; Guerreiro, F.; Rosa da Costa, A.M.; Buttini, F.; Grenha, A. Dual Antibiotherapy of Tuberculosis Mediated by Inhalable Locust Bean Gum Microparticles. Int. J. Pharm. 2017, 529, 433–441. [Google Scholar] [CrossRef]
  101. Adeleke, O.A.; Hayeshi, R.K.; Davids, H. Development and Evaluation of a Reconstitutable Dry Suspension Containing Isoniazid for Flexible Pediatric Dosing. Pharmaceutics 2020, 12, 286. [Google Scholar] [CrossRef] [Green Version]
  102. Rodrigues, S.; Cunha, L.; Kollan, J.; Neumann, P.R.; Rosa da Costa, A.M.; Dailey, L.A.; Grenha, A. Cytocompatibility and Cellular Interactions of Chondroitin Sulfate Microparticles Designed for Inhaled Tuberculosis Treatment. Eur. J. Pharm. Biopharm. 2021, 163, 171–178. [Google Scholar] [CrossRef]
  103. Mukhtar, M.; Pallagi, E.; Csóka, I.; Benke, E.; Farkas, Á.; Zeeshan, M.; Burián, K.; Kókai, D.; Ambrus, R. Aerodynamic Properties and in Silico Deposition of Isoniazid Loaded Chitosan/Thiolated Chitosan and Hyaluronic Acid Hybrid Nanoplex DPIs as a Potential TB Treatment. Int. J. Biol. Macromol. 2020, 165, 3007–3019. [Google Scholar] [CrossRef]
  104. Cunha, L.; Rodrigues, S.; Rosa da Costa, A.M.; Faleiro, L.; Buttini, F.; Grenha, A. Inhalable Chitosan Microparticles for Simultaneous Delivery of Isoniazid and Rifabutin in Lung Tuberculosis Treatment. Drug. Dev. Ind. Pharm. 2019, 45, 1313–1320. [Google Scholar] [CrossRef]
  105. Pires, D.; Mandal, M.; Matos, A.I.; Peres, C.; Catalão, M.J.; Azevedo-Pereira, J.M.; Satchi-Fainaro, R.; Florindo, H.F.; Anes, E. Development of Chitosan Particles Loaded with SiRNA for Cystatin C to Control Intracellular Drug-Resistant Mycobacterium tuberculosis. Antibiotics 2023, 12, 729. [Google Scholar] [CrossRef] [PubMed]
  106. Cunha, L.; Rodrigues, S.; da Costa, A.M.R.; Faleiro, M.L.; Buttini, F.; Grenha, A. Inhalable Fucoidan Microparticles Combining Two Antitubercular Drugs with Potential Application in Pulmonary Tuberculosis Therapy. Polymers 2018, 10, 636. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  107. Upadhyay, T.K.; Fatima, N.; Sharma, A.; Sharma, D.; Sharma, R. Nano-Rifabutin Entrapment within Glucan Microparticles Enhances Protection against Intracellular Mycobacterium tuberculosis. Artif. Cells Nanomed. Biotechnol. 2019, 47, 427–435. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  108. Kalluru, R.; Fenaroli, F.; Westmoreland, D.; Ulanova, L.; Maleki, A.; Roos, N.; Madsen, M.P.; Koster, G.; Egge-Jacobsen, W.; Wilson, S.; et al. Poly(Lactide-Co-Glycolide)-Rifampicin Nanoparticles Efficiently Clear Mycobacterium bovis BCG Infection in Macrophages and Remain Membrane-Bound in Phago-Lysosomes. J. Cell Sci. 2013, 126, 3043–3054. [Google Scholar] [CrossRef] [Green Version]
  109. Edagwa, B.J.; Guo, D.; Puligujja, P.; Chen, H.; McMillan, J.E.; Liu, X.; Gendelman, H.E.; Narayanasamy, P. Long-Acting Antituberculous Therapeutic Nanoparticles Target Macrophage Endosomes. FASEB J. 2014, 28, 5071–5082. [Google Scholar] [CrossRef] [PubMed]
  110. Andreu, V.; Larrea, A.; Rodriguez-fernandez, P.; Alfaro, S. Matryoshka-Type Gastro-Resistant Microparticles for the Oral Treatment of Mycobacterium tuberculosis. Nanomedicine (Lond) 2019, 14, 707–726. [Google Scholar] [CrossRef] [Green Version]
  111. Daddio, S.M.; Reddy, V.M.; Liu, Y.; Sinko, P.J.; Einck, L.; Prudhomme, R.K. Antitubercular Nanocarrier Combination Therapy: Formulation Strategies and in vitro Efficacy for Rifampicin and SQ641. Mol. Pharm. 2015, 12, 1554–1563. [Google Scholar] [CrossRef] [Green Version]
  112. Jahagirdar, P.S.; Gupta, P.K.; Kulkarni, S.P.; Devarajan, P.V. Intramacrophage Delivery of Dual Drug Loaded Nanoparticles for Effective Clearance of Mycobacterium tuberculosis. J. Pharm. Sci. 2020, 109, 2262–2270. [Google Scholar] [CrossRef]
  113. Zhou, Y.; Kong, Y.; Kundu, S.; Cirillo, J.D.; Liang, H. Antibacterial Activities of Gold and Silver Nanoparticles against Escherichia coli and Bacillus Calmette-Guérin. J. Nanobiotechnol. 2012, 10, 19. [Google Scholar] [CrossRef] [Green Version]
  114. Harbut, M.B.; Vilchèze, C.; Luo, X.; Hensler, M.E.; Guo, H.; Yang, B.; Chatterjee, A.K.; Nizet, V.; Jacobs, W.R.; Schultz, P.G.; et al. Auranofin Exerts Broad-Spectrum Bactericidal Activities by Targeting Thiol-Redox Homeostasis. Proc. Nat. Acad. Sci. USA 2015, 112, 4453–4458. [Google Scholar] [CrossRef] [PubMed]
  115. Frei, A.; Verderosa, A.D.; Elliott, A.G.; Zuegg, J.; Blaskovich, M.A.T. Metals to Combat Antimicrobial Resistance. Nat. Rev. Chem. 2023, 7, 202–224. [Google Scholar] [CrossRef] [PubMed]
  116. Wu, T.; Liao, W.; Wang, W.; Zhou, J.; Tan, W.; Xiang, W.; Zhang, J.; Guo, L.; Chen, T.; Ma, D.; et al. Genipin-Crosslinked Carboxymethyl Chitosan Nanogel for Lung-Targeted Delivery of Isoniazid and Rifampin. Carbohydr. Polym. 2018, 197, 403–413. [Google Scholar] [CrossRef] [PubMed]
  117. Grenha, A.; Alves, A.D.; Guerreiro, F.; Pinho, J.; Simões, S.; Almeida, A.J.; Gaspar, M.M. Inhalable Locust Bean Gum Microparticles Co-Associating Isoniazid and Rifabutin: Therapeutic Assessment in a Murine Model of Tuberculosis Infection. Eur. J. Pharm. Biopharm. 2020, 147, 38–44. [Google Scholar] [CrossRef]
  118. Singh, A.K.; Verma, R.K.; Mukker, J.K.; Yadav, A.B.; Muttil, P.; Sharma, R.; Mohan, M.; Agrawal, A.K.; Gupta, A.; Dwivedi, A.K.; et al. Inhalable Particles Containing Isoniazid and Rifabutin as Adjunct Therapy for Safe, Efficacious and Relapse-Free Cure of Experimental Animal Tuberculosis in One Month. Tuberculosis 2021, 128, 102081. [Google Scholar] [CrossRef]
  119. Antonov, E.N.; Andreevskaya, S.N.; Bocharova, I.V.; Bogorodsky, S.E.; Krotova, L.I.; Larionova, E.E.; Mariyanats, A.O.; Mishakov, G.V.; Smirnova, T.G.; Chernousova, L.N.; et al. PLGA Carriers for Controlled Release of Levofloxacin in Anti-Tuberculosis Therapy. Pharmaceutics 2022, 14, 1275. [Google Scholar] [CrossRef]
  120. Srichana, T.; Ratanajamit, C.; Juthong, S.; Suwandecha, T.; Laohapojanart, N.; Pungrassami, P.; Padmavathi, A.R. Evaluation of Proinflammatory Cytokines and Adverse Events in Healthy Volunteers upon Inhalation of Antituberculosis Drugs. Biol. Pharm. Bull. 2016, 39, 1815–1822. [Google Scholar] [CrossRef] [Green Version]
  121. Laohapojanart, N.; Ratanajamit, C.; Kawkitinarong, K.; Srichana, T. Efficacy and Safety of Combined Isoniazid-Rifampicin-Pyrazinamide-Levofloxacin Dry Powder Inhaler in Treatment of Pulmonary Tuberculosis: A Randomized Controlled Trial. Pulm. Pharmacol. Ther. 2021, 70, 102056. [Google Scholar] [CrossRef]
  122. Luz, I.; Stewart, I.E.; Mortensen, N.P.; Hickey, A.J. Designing Inhalable Metal Organic Frameworks for Pulmonary Tuberculosis Treatment and Theragnostics: Via Spray Drying. Chem. Commun. 2020, 56, 13339–13342. [Google Scholar] [CrossRef]
  123. Cai, W.; Wang, J.; Chu, C.; Chen, W.; Wu, C.; Liu, G. Metal–Organic Framework-Based Stimuli-Responsive Systems for Drug Delivery. Adv. Sci. 2019, 6, 1801526. [Google Scholar] [CrossRef] [Green Version]
  124. Jiménez-Rodríguez, R.; Douda, J.; Mota-Díaz, I.I.; Luna-Herrera, J.; Romera-Ibarra, I.C.; Casas-Espínola, J.L. Theragnostic Liposomes for the Diagnosis and Treatment of Tuberculosis. MRS Adv. 2023, 8, 67–70. [Google Scholar] [CrossRef]
  125. Li, B.; Tan, Q.; Fan, Z.; Xiao, K.; Liao, Y. Next-Generation Theranostics: Functionalized Nanomaterials Enable Efficient Diagnosis and Therapy of Tuberculosis. Adv. Ther. 2020, 3, 1900189. [Google Scholar] [CrossRef]
  126. Liao, Y.; Li, B.; Zhao, Z.; Fu, Y.; Tan, Q.; Li, X.; Wang, W.; Yin, J.; Shan, H.; Tang, B.Z.; et al. Targeted Theranostics for Tuberculosis: A Rifampicin-Loaded Aggregation-Induced Emission Carrier for Granulomas Tracking and Anti-Infection. ACS Nano 2020, 14, 8046–8058. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Particles for nasal and pulmonary administration. The main physicochemical properties to consider for the preparation of particles intended for TB prophylaxis and treatment are size, shape, and surface charge. Sizes from 100 to 500 nm are preferable but not exclusive for nasal administration, while 1–5 μm particles are optimal for intrapulmonary administration. Shapes such as those present in nature are also favorable for the internalization of particles in the lung, mainly those that are spherical or rod-shaped, as is the case for Mtb. Additionally, a preferential but not exclusive interaction between cationic microparticles (+) and macrophages and anionic nanoparticles (−) with dendritic and epithelial cells has been documented. All these properties in turn will depend on the fabrication materials. For the formulation of TB vaccines, the most reported materials are natural and synthetic polymers, followed by lipids (for the fabrication of solid lipid nanoparticles and liposomes), while, for TB treatment formulations, the use of lipids is most frequent, followed by synthetic and natural polymers.
Figure 1. Particles for nasal and pulmonary administration. The main physicochemical properties to consider for the preparation of particles intended for TB prophylaxis and treatment are size, shape, and surface charge. Sizes from 100 to 500 nm are preferable but not exclusive for nasal administration, while 1–5 μm particles are optimal for intrapulmonary administration. Shapes such as those present in nature are also favorable for the internalization of particles in the lung, mainly those that are spherical or rod-shaped, as is the case for Mtb. Additionally, a preferential but not exclusive interaction between cationic microparticles (+) and macrophages and anionic nanoparticles (−) with dendritic and epithelial cells has been documented. All these properties in turn will depend on the fabrication materials. For the formulation of TB vaccines, the most reported materials are natural and synthetic polymers, followed by lipids (for the fabrication of solid lipid nanoparticles and liposomes), while, for TB treatment formulations, the use of lipids is most frequent, followed by synthetic and natural polymers.
Microorganisms 11 01988 g001
Figure 2. Immune activation induced by Mtb and particulate systems after inhalation.
Figure 2. Immune activation induced by Mtb and particulate systems after inhalation.
Microorganisms 11 01988 g002
Figure 3. Roles of particles in the formulation of vaccines against TB.
Figure 3. Roles of particles in the formulation of vaccines against TB.
Microorganisms 11 01988 g003
Figure 4. SWOT analysis of particulate systems for the formulation of TB vaccines or treatments. HMDR: hypervirulent multi-drug resistance.
Figure 4. SWOT analysis of particulate systems for the formulation of TB vaccines or treatments. HMDR: hypervirulent multi-drug resistance.
Microorganisms 11 01988 g004
Table 1. Polymeric particulate TB vaccines and the immune response induced in vivo.
Table 1. Polymeric particulate TB vaccines and the immune response induced in vivo.
Particulate SystemVaccine
Formulation
(Antigen/Adjuvant)
Scheme of Immunization (Model/Route/Dose)Immune Response InducedRef
NATURAL POLYMERSChitosan NPs
Encapsulation
pDNA encoding Ag 85B
No extra adjuvant
BALB/c mice
50 μg
SC, 1× at day 0
IN, 2× at 2w interval
-
Higher levels of Ag-specific IgG and IgG2a
-
Enhanced proliferation of CD4+ T cells
-
Higher release of IFN-γ and IL-2 in splenocytes restimulated with Ag 85B
-
Induction of autophagy
[79]
Chitosan NPs
Mix
ESAT-6 (1–20 peptide)
A: MPL
C57BL/6J, IFN-γ−/− and IL-17−/− mice
133 μg/50 μg
IN, 3× at 2w interval
-
Potent induction of Th17 in lung and Th1/Th17 in spleen
-
Higher protection against Mtb infection
-
Induction of NLRP3-independent inflammasome and IL-1β
[80]
Chitosan NPs
Coating
Mtb cell wall lipids
No extra adjuvant
BALB/c mice
0.5 mg/kg
SC, 1×
IP, 4× 0, 21, 45, 66 days
-
Higher levels of Th1 (IFN-γ, IL-2, TNF-α) and Th2 cytokines (IL-4, IL-5, IL-13) in LNs and spleen
-
γδ T-cell activation in LNs
-
Higher levels of IgG, IgG1, IgG2, and IgM
[81]
Inulin chitosan NPs
Conjugation
Fusion
CT (CFP10-TB10.4)
No extra adjuvant
C57BL/6 mice
100 μg/mL
SC, 3× 0, 14, 28 days
-
Higher release of IFN-γ, TNF-α, IL-2, and IL-4 in splenocytes restimulated with CT
-
Higher levels of CT-specific IgG1 and IgG2b
[82]
Advax™
(δ-Inulin-NPs)
Mix
Fusion
CysVac2 (Ag85B-CysD)
A: Advax CpG
C57BL/6 mice
3 μg of fusion/mg inulin
IM, 3× at 2w interval
-
Induction of specific multifunctional CD4+ T cells (IFN-γ+, TNF+, IL-2+)
-
Reduction in CFU in lung after Mtb infection
-
Strong immunogenicity and protection
[67]
Advax™
(δ-Inulin-NPs)
Mix
Fusion
CysVac2 (Ag85B-CysD)
A: Advax
C57BL/6 mice
3 μg of fusion/mg inulin
IT, 3× at 2w interval
-
Induction of lung-resident antigen-specific IL17-secreting CD4+ cells
-
Higher protection against Mtb infection compared with BCG-vaccinated mice
[63]
INU/pArg NCs
Adsorption
Fusion
ECH (ESAT6/CFP-10)
A: Imiquimod
C57BL/6 mice
10 μg of fusion
IN, 3×
-
Higher levels of IgA in bronchoalveolar fluid
-
Higher titers of IgG in sera
-
Higher release of IFN-γ and IL-17 in splenocytes restimulated with ECH
[83]
Dextran NPs
Immobilization
Fusion
GamTBvac (Ag85A-ESAT6-CFP10-DBD)
A: DEAE-dextran-CpG
C57BL/6 mice and guinea pigs
5, 10, 20 μg of fusion
SC, 2× at 3w interval
As booster of BCG vaccine in mice:
-
Higher levels of IFN-γ and Ag-specific IgG
-
Reduction in CFU in lung
-
Higher survival
[68]
SYNTHETIC POLYMERSPLGA NPs EncapsulationFusion
HspX/EsxS
A: DOTAP
BALB/c mice
25 μg of fusion/5 mg NPs
SC, 3× at 2w interval
-
Higher levels of IFN-γ
-
Higher titers of specific IgG1 and IgG2a compared with BCG
[84]
PLGA NPs
Encapsulation
Plasmid pcDNA3.1/Mtb72F
A: TB10.4 and/or CpG
BALB/c mice
SC, 1× BCG or plasmid at day 0
SC, 3× 7, 14, 21 days
As booster of BCG vaccine:
-
Higher levels of IFN-γ in splenocytes restimulated with BCG
[85]
Polyester NPs
Coating
Fusion
H28 (Ag85B-TB10.4-Rv2660c)
H4 (Ag85B-TB10.4)
A: DDA
C57BL/6 mice
2–10 μg of fusion
SC, 3× at 9-day intervals
PNPs-H4 induced:
-
Long-lasting antigen-specific T-cell responses
-
Protective immunity in infected mice
-
Reduction in CFU in lung
-
Similar protective immunity to BCG
[43]
NPs: nanoparticles; INU/pArgNCs: inulin/polyarginine nanocapsules; PLGA: poly(lactide-co-glycolide); A: adjuvant; MPL: monophosphoryl lipid A; DEAE: diethylaminoethyl; DOTAP: 1,2-dioleoyl-3-trimethylammonium propane; DDA: dimethyldioctadecyl ammonium bromide; SC: subcutaneous; IN: intranasal; IM: intramuscular; IP: intraperitoneal; IT: intratracheal; LNs: lymph nodes.
Table 2. Particulate systems evaluated in vitro as carriers of anti-tuberculosis drugs.
Table 2. Particulate systems evaluated in vitro as carriers of anti-tuberculosis drugs.
Particulate SystemDrugAdministration Scheme
(Cell Line/Strategy)
ObservationsRef
NATURAL POLYMERSChitosan MPsINH
RFB
A549, THP-1 Mφ
AI with Mb BCG
-
Cell viability above 70% for A549 cells
-
Dose-dependent effect on THP-1 Mφ
-
Microencapsulation preserved antibacterial activity of drugs
-
Free and drug-loaded MPs induced increased secretion of TNF-α and IL-18 in THP-1 Mφ
[104]
Chitosan NPsAnti-Cystatin C siRNAHMDM, THP-1 Mφ
BI with Mtb H37Rv and susceptible and resistant isolates
-
Loaded NPs were non-cytotoxic and were efficiently internalized by cells
-
Significant reduction in intracellular bacteria
[105]
Fucoidan MPsRFB
INH
A549, THP-1 Mφ
AI with Mb BCG
-
Cell viability above 65% at 24 h
-
Encapsulation reduced RFB cytotoxicity
-
Free and loaded MPs induced TNF-α and IL-8
-
Dose-dependent uptake of MPs
[106]
Glucan NPsRFBJ774
AI with Mtb H37Ra
-
Induction of ROS and NO within infected Mφ
-
Induction of lysosome accumulation and phagolysosomal maturation in infected cells
-
The efficacy of RFB was enhanced 2.5-fold
[107]
SYNTHETIC POLYMERSPLGA NPsRIFRAW 264.7, BMDM
BI with Mb BCG
-
Loaded NPs promoted the efficient clearing of BCG infection over a 12-day period
[108]
PLGA NPsRIF
INHP
HMDM
BI with Ms
-
Sustained release of drugs over 15 days
-
Six-fold increase in therapeutic efficacy
-
Higher cell uptake and better antimicrobial activity than free drugs
[109]
PLGA NPs
encapsulated inside MAAEA MPs
RIFCaco2, MH-S
AI with Mtb H37Rv
-
Loaded NPs translocated to the basolateral side of Caco2 cells and were not cytotoxic
-
Loaded and empty NPs decreased growth of intracellular bacteria
[110]
Poly(ε-caprolactone)INH
SQ641 + CsA + VE
J774A.1
AI with Mtb H37Rv
-
Better inhibition of intracellular replication of Mtb with SQ641-CsA-VE than SQ641 alone or INH
[111]
Poly(ethylene sebacate) NPsRIF-CURRAW 264.7
AI with Mtb H37Rv
-
NPs were non-cytotoxic
-
Showed 1.5-fold higher drug internalization compared to free drugs
-
Significant killing of intracellular bacteria
[112]
NPs: nanoparticles; MPs: microparticles; PLGA: poly(lactic-co-glycolic) acid; MAAEA: methacrylic acid–ethyl acrylate copolymer; RIF: rifampicin; INHP: pentenyl–isoniazid; INH: isoniazid; SQ641 + CsA + VE: natural analogue of capuramycin + cyclosporine A + vitamin E; RFB: rifabutin; CUR: curcumin; BI: before infection; AI: after infection; BMDM: bone-marrow-derived monocytes; HMDM: human-monocyte-derived macrophages; Ms: Mycobacterium smegmatis; Mb: Mycobacterium bovis.
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

Barrera-Rosales, A.; Rodríguez-Sanoja, R.; Hernández-Pando, R.; Moreno-Mendieta, S. The Use of Particulate Systems for Tuberculosis Prophylaxis and Treatment: Opportunities and Challenges. Microorganisms 2023, 11, 1988. https://doi.org/10.3390/microorganisms11081988

AMA Style

Barrera-Rosales A, Rodríguez-Sanoja R, Hernández-Pando R, Moreno-Mendieta S. The Use of Particulate Systems for Tuberculosis Prophylaxis and Treatment: Opportunities and Challenges. Microorganisms. 2023; 11(8):1988. https://doi.org/10.3390/microorganisms11081988

Chicago/Turabian Style

Barrera-Rosales, Alejandra, Romina Rodríguez-Sanoja, Rogelio Hernández-Pando, and Silvia Moreno-Mendieta. 2023. "The Use of Particulate Systems for Tuberculosis Prophylaxis and Treatment: Opportunities and Challenges" Microorganisms 11, no. 8: 1988. https://doi.org/10.3390/microorganisms11081988

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop