Next Article in Journal
A New Perspective for Vineyard Terroir Identity: Looking for Microbial Indicator Species by Long Read Nanopore Sequencing
Previous Article in Journal
Comparison of Atmospheric and Lithospheric Culturable Bacterial Communities from Two Dissimilar Active Volcanic Sites, Surtsey Island and Fimmvörðuháls Mountain in Iceland
Previous Article in Special Issue
Modulation of Virulence-Associated Traits in Aspergillus fumigatus by BET Inhibitor JQ1
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Antibodies to Combat Fungal Infections: Development Strategies and Progress

by
Ali A. Rabaan
1,2,3,*,
Amal H. Alfaraj
4,
Amer Alshengeti
5,6,
Abdulsalam Alawfi
5,
Sara Alwarthan
7,
Mashael Alhajri
7,
Amal H. Al-Najjar
8,
Mona A. Al Fares
9,
Mustafa A. Najim
10,
Souad A. Almuthree
11,
Sultan T. AlShurbaji
12,
Fadwa S. Alofi
13,
Bashayer M. AlShehail
14,
Buthina AlYuosof
15,
Ahlam Alynbiawi
16,
Suha A. Alzayer
17,
Nawal Al Kaabi
18,
Wesam A. Abduljabbar
19,
Zakiyah A. Bukhary
20 and
Ahmed S. Bueid
21
1
Molecular Diagnostic Laboratory, Johns Hopkins Aramco Healthcare, Dhahran 31311, Saudi Arabia
2
College of Medicine, Alfaisal University, Riyadh 11533, Saudi Arabia
3
Department of Public Health and Nutrition, The University of Haripur, Haripur 22610, Pakistan
4
Pediatric Department, Abqaiq General Hospital, First Eastern Health Cluster, Abqaiq 33261, Saudi Arabia
5
Department of Pediatrics, College of Medicine, Taibah University, Al-Madinah 41491, Saudi Arabia
6
Department of Infection Prevention and Control, Prince Mohammad Bin Abdulaziz Hospital, National Guard Health Affairs, Al-Madinah 41491, Saudi Arabia
7
Department of Internal Medicine, College of Medicine, Imam Abdulrahman Bin Faisal University, Dammam 34212, Saudi Arabia
8
Drug & Poison Information Center, Pharmacy Department, Security Forces Hospital Program, Riyadh 11481, Saudi Arabia
9
Department of Internal Medicine, King Abdulaziz University Hospital, Jeddah 21589, Saudi Arabia
10
Department of Medical Laboratories Technology, College of Applied Medical Sciences, Taibah University, Madinah 41411, Saudi Arabia
11
Department of Infectious Disease, King Abdullah Medical City, Makkah 43442, Saudi Arabia
12
Outpatient Pharmacy, Dr. Sulaiman Alhabib Medical Group, Diplomatic Quarter, Riyadh 91877, Saudi Arabia
13
Department of Infectious Diseases, King Fahad Hospital, Madinah 42351, Saudi Arabia
14
Pharmacy Practice Department, College of Clinical Pharmacy, Imam Abdulrahman Bin Faisal University, Dammam 31441, Saudi Arabia
15
Directorate of Public Health, Dammam Network, Eastern Health Cluster, Dammam 31444, Saudi Arabia
16
Infectious Diseases Section, Medical Specialties Department, King Fahad Medical City, Riyadh 12231, Saudi Arabia
17
Parasitology Laboratory Department, Qatif Comprehensive Inspection Center, Qatif 31911, Saudi Arabia
18
Department of Pediatric Infectious Disease, Sheikh Khalifa Medical City, Abu Dhabi 51900, United Arab Emirates
19
Department of Medical Laboratory Sciences, Fakeeh College for Medical Science, Jeddah 21134, Saudi Arabia
20
Department of Internal Medicine, King Fahad General Hospital, Jeddah 23325, Saudi Arabia
21
Microbiology Laboratory, King Faisal General Hospital, Al-Ahsa 31982, Saudi Arabia
*
Author to whom correspondence should be addressed.
Microorganisms 2023, 11(3), 671; https://doi.org/10.3390/microorganisms11030671
Submission received: 2 February 2023 / Revised: 23 February 2023 / Accepted: 24 February 2023 / Published: 6 March 2023

Abstract

:
The finding that some mAbs are antifungal suggests that antibody immunity may play a key role in the defense of the host against mycotic infections. The discovery of antibodies that guard against fungi is a significant advancement because it gives rise to the possibility of developing vaccinations that trigger protective antibody immunity. These vaccines might work by inducing antibody opsonins that improve the function of non-specific (such as neutrophils, macrophages, and NK cells) and specific (such as lymphocyte) cell-mediated immunity and stop or aid in eradicating fungus infections. The ability of antibodies to defend against fungi has been demonstrated by using monoclonal antibody technology to reconsider the function of antibody immunity. The next step is to develop vaccines that induce protective antibody immunity and to comprehend the mechanisms through which antibodies mediate protective effects against fungus.

1. Introduction

Every environmental niche, including the air, soil, freshwater, and oceans, might provide fungi that could be separated. They play a significant role in the world’s biogeochemical cycles. Many of them are also employed for human welfare. Since a very long time ago, fungi have been employed to ferment nutrients. For instance, yeasts are used to produce food both traditionally and today, as well as to break down the rash and produce items that are helpful to the industry [1,2]. Fungi have been utilized in traditional Chinese medicine for thousands of years, and recent research has concentrated on polysaccharides, an essential part of fungi cell walls [3]. Among the various polysaccharides, fungal β-glucans play a significant role in developing cosmetics, food additives, and medicine [4]. Its positive effects on the course of numerous diseases have also been demonstrated [5]. In this way, fungi contribute significantly to the biomedical and manufacturing industries and impact human health, agriculture, and biodiversity.
Although the great majority of fungus lacks pathogenic characteristics, some species inflict life-threatening superficial illnesses on humans. Invasive infections are far more difficult to cure and tremendously influence human health, leading to a significant mortality rate, whereas superficial fungal infections are typically benign [5]. There are more than a hundred species of fungi that are regarded as primary pathogens of humans and animals [6].
Primary pathogens and opportunistic pathogens are the two broad categories into which fungal pathogens can be subdivided. The fungus can establish themselves and spread infection because of their capacity to develop at 37 °C and adapt to the environment inside the host tissues [7]. Primary (true) fungal pathogens (Blastomyces dermatitidis, Coccidioides immitis, Paracoccidioides brasiliensis, Histoplasma capsulatum, and Penicillium marneffei) attack healthy, immunocompetent hosts and are distributed in identifiably geographic regions [8]. Opportunistic pathogens employ weak or immune-compromised hosts to spread infection and cause considerable morbidity and mortality. Over the past 20 years, fungal infections have paradoxically gained in significance as more immunosuppressed individuals have survived due to the triumph of modern medicine.
Such patients are highly susceptible to disease by opportunistic pathogens such as Candida species, C. neoformans, A. fumigatus, other Aspergillus species, and zygomycetes [7]. Table 1 summarizes the list of common human pathogens.
The processes of bacterial pathogens are far better understood than those of fungal infections. Fungal infections currently make for 10% of all nosocomial infections [9].
The fungus-specific factors that are responsible for the pathogenic nature of fungi are unique to that species. The enzymes and toxins significantly influence the virulence and pathogenicity of many fungi and the byproducts they produce, for example, proteases and aspartic acid proteinase, are made by Aspergillus species. The toxins that are generated by Aspergillus species include gliotoxin and restriction. An opportunistic fungus called Candida spp., a typical component of human flora, can lead to dangerous infections ranging from candidemia and urinary tract infections to skin and nail infections [10].
Human fungi diseases provide a severe threat but are often ignored. Fungi’s effects on human health are a growing concern. Every year, 4.9 million people are affected by invasive or persistent fungal diseases [11]. By comparison, the financial impact is significant. For example, fungal infections are estimated to have cost the United States more than USD 7.2 billion in damages in 2017 [12].
Despite the availability of antifungal medications, invasive fungal infections are thought to be responsible for 1.5 million annual deaths, a similar number to tuberculosis [13]. The global burden of common life-threatening fungal infections data have been summarized by Gago et al. (2017) in Table 2 [11].
A patient with an invasive fungal infection may incur additional medicine and hospitalization costs of between EUR 10,000 and EUR 51,000, according to limited studies on the economic impact conducted in Europe. A lack of access to diagnostic and antifungal medications is a severe issue in lower-income nations. For instance, in sub-Saharan Africa, cryptococcal meningitis is the most common cause of mortality for HIV/AIDS patients [14]. The most effective antifungal medications to treat this infection have been in circulation since the 1970s, yet they are still unavailable in many nations.
Despite the fact that some fungi can infect healthy people and cause disease, the majority of fungal infections occur in sick persons and commonly compromise the success of cutting-edge treatments for cancer, complex surgery, autoimmune disease treatments, organ and stem cell transplantation and trauma [15]. As a result, AIDS-related illnesses such as pneumocystis and cryptococcus are common in developing nations where antiretroviral medication is not widely available [16,17]. Diagnoses of invasive candidiasis, allergic bronchopulmonary disease, and invasive aspergillosis associated with cancer and surgery are becoming more common in wealthier countries [18]. The presence of fungi in the environment is a major determinant of the epidemiology of primary fungal diseases. For example, one of Latin America’s most common systemic mycoses is paracoccidioidomycosis. The geographic distribution of several paracoccidioidomycosis species in these regions significantly limits the epidemiology of this disease [19].
Antibody-based treatments have emerged as a promising approach for treating fungal infections, particularly for those caused by difficult-to-treat species or for patients who are immunocompromised. Some of the specific logic put forth by the researchers to emphasize the superiority of antibodies over other interventions includes their specificity, durability, and adaptability. Antibody-based treatments can be designed to specifically target the fungus of interest while leaving other healthy cells and microbes unharmed. This specificity can lead to fewer side effects and less damage to the patient’s immune system.
Antibodies are naturally long-lasting molecules in the body and can provide prolonged protection against infections. In contrast, some antifungal drugs need to be administered frequently and for extended periods of time, which can lead to drug resistance. Combining antibody-based treatments with other antifungal strategies, such as antifungal drugs, can lead to a synergistic effect and enhance the efficacy of the treatment, which appears to an added advantage. Moreover, antibody-based treatments can be used for prophylaxis to prevent fungal infections in high-risk patients, such as those undergoing chemotherapy or organ transplantation. Finally, antibodies can be engineered to target multiple fungal species and strains, making them adaptable to changing pathogen populations. Scientific evidence has shown the effectiveness of antibody-based treatments for fungal infections. For example, a monoclonal antibody called Mycograb was developed to target the fungus Candida albicans. Clinical trials showed that Mycograb improved the survival rates of patients with invasive candidiasis compared to standard antifungal therapy alone [20]. Another example is the use of passive immunization with polyclonal antibodies against Aspergillus fumigatus in a mouse model. The antibodies were able to reduce the fungal burden and increase survival rates compared to control mice [21].

2. General Mechanism of Fungal Infection

When it comes to the pathogenesis of fungal infections, the nature and life cycle of fungi are crucial. As fungal infections in humans are caused by various types of fungi that can invade tissue, these infections can range from superficial skin infections to life-threatening systemic infections. The general mechanism of fungal infections in humans involves several steps, including colonization, adhesion, invasion, and proliferation of the fungi within the host tissues [22]. The first step in fungal infections is colonization, which involves the adherence of fungal cells to the host cells or extracellular matrix (Figure 1). This process is facilitated by the fungal cell walls’ components, such as adhesins and glycoproteins, which interact with host receptors. For example, Candida albicans, a common fungal pathogen, can adhere to host cells by binding to host proteins such as fibronectin and laminin [23]. Studies have shown that the adhesion of C. albicans to host cells is a critical step in the development of invasive candidiasis, a serious bloodstream infection. The second step in fungal infections is adhesion, which is followed by the invasion of host tissues. Fungi can penetrate host tissues through various mechanisms, such as hyphal growth, phagocytosis, and endocytosis. Once inside the host tissues, fungi can evade the host immune system by several mechanisms, including the modulation of host immune responses and the secretion of virulence factors. For example, Aspergillus fumigatus, a common cause of invasive aspergillosis, can produce gliotoxin, a virulence factor that suppresses host immune responses and promotes fungal growth [24]. The third step in fungal infections is proliferation, which involves the growth and replication of fungal cells within host tissues. Fungi can utilize host nutrients and evade host immune responses to promote their growth and proliferation. For example, Histoplasma capsulatum, a fungal pathogen that causes histoplasmosis, can survive and replicate within host macrophages by avoiding host immune recognition and response [25]. In addition to these steps, fungal infections can also trigger host inflammatory responses, which can contribute to tissue damage and disease progression. Such inflammatory responses triggered by fungal infections can be known to contribute to tissue damage and disease progression. Fungal cell wall components, such as β-glucans and chitin, can activate host immune cells and promote the production of proinflammatory cytokines. These responses can contribute to tissue damage and organ dysfunction. For example, studies have shown that β-glucans from Candida albicans can activate host immune cells and contribute to the development of systemic candidiasis [26].

3. Host Defense Mechanism against Fungi

So, what attributes do human fungal pathogens have? Four requirements must be met. (1) It must be capable of growing at temperatures of at least 37 °C. (2) It must be able to pass through host tissue barriers or find a way around them by penetrating air-filled spaces such as the sinuses and lungs. (3) It must be capable of absorbing and digesting parts of human tissues. (4) It must also be resilient to the human immune system [15].
The way a person’s body fights disease caused by fungi depends on their immune system. The immune system is made up of two parts: basic defenses called “innate immunity” and more complex defenses called “adaptive immunity.” These defenses are created by cells such as macrophages, monocytes, dendritic cells, PMNs, MDSCs, NK cells, alarmins, and antimicrobial peptides. The type of defense a person’s body uses against fungi depends on where the infection is located. However, the relative importance of each is primarily determined by the infection site. The innate response occurs either directly by the phagocytic mechanism, through the secretion of antifungal compounds, or through the production of cytokines and chemokines [27].
Macrophages, antigen-presenting cells (APC), take over the fungi and start different events. The macrophages of tissues work as effector cells and create cytokines and chemokines that stimulate more immune cells. Moreover, they are essential in developing granuloma [28,29]. Additionally, dendritic cells are APCs that process antigens following fungal recognition and deliver them to T cells with a further linage toward T-helper (Th) subsets, such as Th1, Th2, and Th17 cells [30]. Recent studies recognize the significance of plasmacytoid dendritic cells in aspergillosis [31]. Monocytes were classified into classical (CD14++CD16), intermediate (CD14+CD16+), and nonclassical (CD14+CD16++) subsets based on CD14 and CD16 expression [32]. Th17 differentiation is initiated by classical monocytes stimulated with IL-1 and prostaglandin E2 against Candida albicans [33]. Natural killer (NK) cells produce cytokines, primarily IFN-, as well as other soluble factors that regulate the activity of other immune cells [34]. NK cells harm the hyphae of both C. albicans [35] and A. fumigatus [36]. Recent research has expanded on the role of invariant NKT (iNKT) cells in cryptococcosis [37] as well as in aspergillosis [38].
The lack of Polymorphonuclear neutrophils (PMNs) shows an overall risk of invasive Candidiasis [39] and Aspergillosis [40]. They are the storehouse for several antifungal peptides, myeloperoxidase, and nucleolytic enzymes that eradicate several pathogens. Myeloid-derived suppressor cells (MDSCs) showed higher levels in people infected with either Candida albicans and Aspergillus fumigatus [41]. Β defensins, histantin, LL-37, and serprocidins are the antimicrobial peptides (AMP) which showed anti-candidacidal activity [42].
There are two types of adaptive immune responses: cell-mediated responses and humoral immune responses, and both can be elaborated in terms of the aforementioned context.

4. Cell-Mediated Response

Helper CD4 and cytotoxic CD8 T cells are examples of cell-mediated responses. When fungal proteins are detected, they are cleaved into small antigenic peptides, which are then assembled with MHC-I or MHC II and transported to the surface of active dendritic cells. Dendritic cells then migrate to secondary lymphoid organs, where T cells are activated. Dendritic cells then migrate to secondary lymphoid organs and activate T cells. The antigenic peptide sequence influences T cell activation. CD8 T cells contain MHC-I and MHC-II peptides, while CD4 T cells contain MHC-II related peptides.
CD4+ T lymphocytes play an important protective role in patients with human immunodeficiency virus or acquired immunodeficiency syndrome who are susceptible to fungal infections such as Candida albicans, P. jirovecii, C. neoformans, and Aspergillus fumigatus [43].
The depletion of CD8 T cells in mice reduces survival in a cryptococcal infection model and inhibits C. neoformans growth in macrophages by secreting IFN g [44]. CD8 T cells become crucial and immunologically vital for suppressing fungal infections, especially when CD4 T cells are absent or perhaps have relatively low concentrations [45]. Figure 2 explains the cellular innate response and humoral response against fungal infections.

5. Humoral Immune Response

Glycoproteins, also known as antibodies or immunoglobulins (Igs), are critical components of the immune system. The fungal burden is reduced by these effector molecules. After binding to surface antigens, antibodies can act as opsonins, increasing the phagocytic activity of immune cells. The fungal cell wall, which is primarily composed of carbohydrate polymers and specific proteins, is the primary target structure for opsonizing antibodies. The control mechanism for fungal infections is determined by the type of fungus and the specificity of the epitope. Natural antibodies are polyreactive, have a low to medium affinity, and are frequently germ-line encoded. Natural antibodies of the IgM, IgA, and IgG classes are primarily produced by B1 cells. [46,47,48,49,50,51,52].
As we all know, a variety of antifungal antibodies protect the host against specific fungal derivatives such as glycoproteins, glycolipids or peptides, and polysaccharides. Because the majority of these components are found in the fungi’s cell wall, antifungal antibodies concentrate on its development, dynamics, and remodeling [53]. Humoral immunity has long been thought to support cellular immunity, and that the cellular immune response contributes significantly to protection.

6. Antibodies against Fungal Infection

Numerous studies have shown the crucial function of natural antibodies in the body’s fight against fungus infection. It has been demonstrated in multiple models that involve the transfer of normal mice serum to MT mice that they limit the proliferation of microbial fungi [54,55,56,57]. Natural antibodies have a low-to-medium affinity, are often germ-line encoded, and are polyreactive. Natural antibodies such as IgM, IgA, and IgG classes are primarily produced by B1 cells. [46,47,48,49,50,51,52]. This natural antibody’s antifungal properties include the direct reduction in germ tube development and the enhancement of candida phagocytosis through macrophage-mediated opsonization [58,59].
Studies using Candida albicans and Candida neoformans, which are two of the most well-studied and widespread human pathogenic fungi, have provided much of the current knowledge on the mechanism of AMI against fungi. In the case of C. neoformans, opsonization is essential for host defense because polysaccharide capsules block phagocytosis and complement is comparatively weaker than the opsonin-mediated response for a few fungal strains [60]. As C. neoformans can multiply intracellularly and escape phagocytosis through non-proliferative exocytosis, the efficacy of ingestion as a method of modulating this fungus depends on cellular activation and host immunity. In animal models of cryptococcosis, complement activation appears to be required for the function of specific IgMs, but it may not be required for IgGs [61,62]. According to Van Spiiel et al. (2001), Fc receptor-dependent ADCC may protect against C. albicans [63], while C. neoformans have been shown to be sensitive to NK cell antibodies [64,65]. The ability of killer toxin-mimicking antibodies to directly damage fungal cells by interacting with the killer toxin receptor serves as an example of a mediated antifungal mechanism and other antibody intervals [66]. Table 3 [67,68,69] summarizes the host PRRs (pattern recognition receptor) and their related fungal PAMPs (pathogen-associated molecular patterns).
However, in addition to T cell deficiencies, B cell deficits and hypogammaglobulinemia [70,71,72] have also been linked to cryptococcal illness [73] in patients. In addition, IgG2 deficiency has been found in healthy subjects with C. gattii meningitis [74], and in adults and children with HIV/AIDS with specific levels of IgG2, the polysaccharide glucuronoxylomannan (GXM) was lower than those without the disease [75,76]. In humans, IgG2 is the most common IgG subtype of the polysaccharide antigen [77].
Natural IgM are crucial for mice’s resistance to Pneumocystis murina and C. neoformans through various mechanisms. Native IgM promotes the development of Th2 cells, facilitates DC detection and the migration of fungal antigens to the draining lymph nodes, and promotes B cell class switching by recruiting macrophages to the sites of infection and improving fungal phagocytosis [78,79]. According to these results, infection with C. neoformans was more severe in X-linked immunodeficiency mice with much lower IgM levels [80].
Additionally, new mechanisms of antifungal antibodies have been discovered. For instance, a mAb to the mannoprotein of Candida albicans has been demonstrated to mediate three distinct activities: the inhibition of adhesion, suppression of germination, and direct candidacidal activity [81]. The direct effects include iron deprivation caused by the antibody.
The ability of antibodies to cell wall β-glucans to directly limit the growth of Cryptococcus neoformans and C. albicans has been shown to interfere with adhesion and cell wall remodeling [82,83]. Data demonstrating that C. neoformans opsonized by an antibody leaves macrophages as microcolonies that resemble biofilms, in contrast to planktonic cells that leave after complement opsonization, serve as an example of how a specific antibody can change the nature of the interaction between C. neoformans and macrophages [84].
Mice with B cell depletion are more susceptible to developing systemic candidiasis [85]. As demonstrated in erythemic mice challenged with Candida albicans at the systemic level [86], a mice model of severe combined immunodeficiency or SCID [87], and antibody-deficient animal models which lack Lyb-5 B cells [88], humoral immune responses are crucial in providing protection against systemic candidiasis.
The multiple stages of human C. albicans infection, from homosexuality to disease, pose a particular challenge in determining the role of AMI in disease resistance and maximizing the potential for antibody therapy in patients. Mice studies are simpler, but the results vary depending on the fungal strain, infection technique, and animal model. B-lymphocyte-depleted mice are more susceptible to filamentous C. albicans infection, according to previous studies [89].
Mice with B cell depletion are more susceptible to developing systemic candidiasis [85]. As demonstrated by systemic challenges with C. albicans in the athymic mice [86], severely combined immunodeficiency (SCID) mice [87], and antibody-deficient CBA/N mice lacking Lyb-5 B cells [88], the humoral immune response is crucial in protection against systemic candidiasis.
According to a study by Romani and colleagues, antibodies are essential for the development of memory of antifungal immunity [57]. They demonstrated that B-deficient (μMT) mice could effectively control the growth of fungal pathogens during different stages (primary and secondary) of infections by comparing their susceptibility to Aspergillus fumigatus Candida albicans infections via either intravenous or intratracheal routes. Thus, their findings demonstrate the significance of Th1 cells in mediating the protective response against these two fungi infections. However, μMT mice were unable to survive a subsequent C. albicans infection. These findings showed that even though B cells and antibodies are not necessary for Aspergillus resistance, the shield against Candida infections are facilitated by both pathways, antibody-dependent and independent [57].
Circulating healthy antibodies were found to limit the size and organization of the granulomatous lesions in the experimental paracoccidioidomycosis, a chronic granulomatous illness brought by the thermally dimorphic fungus called Paracoccidioides brasiliensis [90].
Critical antifungal agents should adhere to several standards [91]:
  • (a) Should be a broad spectrum for a variety of fungi;
  • (b) Should be fungicidal as opposed to fungistatic;
  • (c) Should be against specific fungal target region;
  • (d) Should have spare interference with host targets;
  • (e) Should have minimal side effects or toxicities.
It was evident that finding or creating a medication that meets the aforementioned requirements is extremely difficult and necessitates discovering biological targets specific to fungus. Table 4 shows the list of some antifungal vaccines (Nami et al. (2019) with some modifications) [92].
When polyene antifungal agents interact with ergosterol in the plasma membrane, they disrupt the membrane [146], increase permeability, cause cytoplasmic leakage [147], and ultimately cause cell death [148]. According to recent research, they may also produce oxidative damage, which could help explain their fungicidal effect [149]. Amphotericin B, nystatin, natamycin (pimaricin), rimocidin, filipin, and candicin have a higher affinity for ergosterol than mammalian cholesterol, resulting in less toxicity to mammalian cells and clinically useful polyenes [150]. The advancement of polyenes is based on changes in the drug structure that convert molecular umbrella complexes into nanoparticles and polysaccharide complexes. The oral drug amphotericin B cochleate lipid crystal nanoparticles were developed and demonstrated in vivo activity [151]. Various PRRs and their counterpart PAMPs in fungi are enumerated in Table 4.
Viamet Pharmaceuticals’ azole enhancements are based on chemical changes that can modulate the metal-binding groups of azole compounds, reducing interactions with cytochrome P450s and the potential for drug–drug interactions [152]. Triazoles were shown to be more potent than itraconazole in vitro against Candida albicans, Trichosporon beigelli, Candida parapsilosis, Aspergillus fumigatus, Candida glabrata, and Candida neoformans [153].
There are three new compounds which are currently in phase II clinical trials with reference IDs: NCT02267356; NCT02267382; and VT-1129. These have been designed to combat vaginal candidiasis and onychomycosis. VT-1129 was created to treat cryptococcosis, and a related compound, VT-1598, is potent and effective against endemic mycoses and cryptococcus [153].
IPC synthase is targeted by drugs such as aureobasidin A, caspofungin, fosfomycin, 21-hydroxylstrin, galbonolide, 21-hydroxygalbonolide, and profungin A. Sphingolipids are important membrane components of fungi and mammals, and are mostly found in the fungal cytoplasmic membrane’s outer layer [154]. Serine palmitoyltransferase is an enzyme that catalyzes the condensation of serine with fatty acyl coenzyme A (acyl-CoA, usually palmitoyl-CoA) to produce the long-chain base keto-Dihydrosphingosine. This is the first and most important step in sphingolipid biosynthesis. The first fungal-specific sphingolipid biosynthetic enzyme is inositol phosphoramide (IPC) synthase [155]. It catalyzes the transfer of phosphatidylinositol from glycophosphatidylinositol to the C1 hydroxyl of ceramide to produce inositol phosphoramide.
According to Monk and Perlin [156], omeprazole, which blocks the stomach K+ H+-ATPase, can prevent the growth of Candida albicans. They also showed a correlation between the inhibition of the organism’s H+-ATPase and the inhibition of the growth of Candida albicans.
The first echinocandin antifungal agent licensed for use in clinical settings was caspofungin. Micafungin and anidulafungin were shortly introduced after it. These substances prevent the synthesis of -1,3-Dglucan, a crucial and essential component of the cell walls of numerous fungus, including Candidaspp. [157,158]. It has been demonstrated that mulundocandin and deoxymulundocandin are effective against A. niger and C. albicans [159].
Rezafungin is an anidulafungin structural analogue with in vivo activity against Candida and Aspergillus species. Piperazine pyridazinones have antifungal activity against Aspergillus and Candida species in vitro [160].
Cell viability depends on chitin, which is a linear homopolymer of N-acetylglucosamine (GlcNAc, 1–4 linked). A series of membrane-localized chitin synthases (Chs1, 2, and 3) were characterized and found in Candida albicans and Saccharomyces cerevisiae, and Chs8 was further observed in Candida albicans, which synthesizes chitin and remains crucial for survival [161].
By inhibiting fungal chitin and glucan synthases, arthrichitin has a broad spectrum of activity against Candida, Trichophyton, and many plant pathogens but shows more significant activity against chitin synthase [162].
The alkaloid inhibits C. albicans growth by stabilizing the cleavage complex formed by C. albicans topoisomerase [163]. Further, it was found that the activity of topoisomerase I from C. albicans to eupolauridine is better than that homo sapiens enzyme [164].
No Aspergillus-licensed vaccinations are available currently to protect people from contracting aspergillosis [165]. Since protein carriers considerably enhance specific antibody responses and carbohydrate antigens are not very immunogenic, they protect A. fumigatus and C. albicans [82].
Several studies also show that mice can be immunized with the suitable fungal antigens to produce protective antifungal antibody responses. As a result, immunization with a liposomal–mannan admixture provided C. albicans resistance dependent on antibodies. In addition, high-titer antibodies to test antigens, including β-mannan and fructose bisphosphate aldolase (EC 4.1. 2.13), methyltetrahydropteroyltriglutamate (EC 2.1.1.14), and mycelial wall protein-1, were also demonstrated in glycoprotein-based synthetic vaccines which were primarily peptide epitopes with β-mannan. As a result of the DC immunization strategy, these antibodies also provided protection against experimental disseminated candidiasis [113].
Mucosal protective IgG and IgA antibodies against Candida albicans Sap2t are produced after an intravaginal vaccination with the secreted aspartic protease (Sap2t) family. Mice were protected from vaginal candidiasis by passive infusion of these or anti-Sap2t IgM and IgG monoclonal antibodies [95].

7. Future Prospects

Antibody-based therapies for fungal infections are an active area of research with several future perspectives and potential benefits. Researchers are developing new antibodies against various fungal pathogens, which can be used for the treatment and prevention of infections. Antibody-based therapies can be combined with other antifungal drugs to increase efficacy and reduce the likelihood of developing resistance. Advances in technologies such as genetic engineering and synthetic biology may enable the development of more effective and specific antibodies. Antibody-based therapies can be tailored to the individual patient’s needs, based on the specific fungal strain causing the infection and the patient’s immune status. However, there are also limitations and challenges associated with antibody-based therapies. The high cost of these therapies may limit their accessibility to patients who cannot afford them. Although antibodies are generally well tolerated, they can still cause adverse effects in some patients. As with all antimicrobial agents, there is a risk of developing a resistance to antibody-based therapies. Currently, there are only a few approved antibody-based therapies for fungal infections, and they may not be effective against all fungal pathogens. In terms of benefits to future humans, antibody-based therapies can improve the outcomes of fungal infections, especially for patients who are immunocompromised. They are highly specific, which can reduce the incidence of side effects compared to broad-spectrum antifungal drugs. Antibody-based therapies can also be used for prophylaxis to prevent fungal infections in high-risk patients, such as those undergoing chemotherapy or organ transplantation. With continued research to address the challenges and limitations and to develop new antibodies and technologies, antibody-based therapies offer a promising approach for treating fungal infections.

8. Conclusions

There is no doubt that the threat posed by FIs will continue to grow globally, but a number of challenges (such as the emergence of resistance) must be faced. At various levels, quick and inventive action is required.
By developing and improving novel therapeutic approaches, antifungal susceptibility testing for the identification of antifungal defiance, rapid diagnostics, enhancing public health capabilities by raising awareness of FIs at social and governmental levels, and conducting research on understanding fungal infection, the burden of fungus infections could be significantly reduced. The findings indicated that the primary determinant of the function of antibodies in the defense against fungi is the number of protective antibodies.

Author Contributions

Conceptualization, A.A.R., A.H.A., A.A. (Amer Alshengeti) and A.S.B.; methodology, A.A.R., A.H.A., A.A. (Amer Alshengeti) and A.S.B.; resources, A.A.R., A.H.A., A.A. (Amer Alshengeti) and A.S.B.; data curation, A.A.R., A.H.A., A.A. (Amer Alshengeti), A.A. (Abdulsalam Alawfi), S.A., M.A., A.H.A.-N., M.A.A.F., M.A.N., S.A.A. (Suha A. Alzayer), S.T.A., F.S.A., B.M.A., B.A., A.A. (Ahlam Alynbiawi), S.A.A. (Suha A. Alzayer), N.A.K., W.A.A., Z.A.B. and A.S.B.; writing—original draft preparation, A.A.R. and A.S.B.; writing—review and editing, A.A.R., A.H.A., A.A. (Amer Alshengeti), A.A. (Abdulsalam Alawfi), S.A., M.A., A.H.A.-N., M.A.A.F., M.A.N., S.A.A. (Souad A. Almuthree), S.T.A., F.S.A., B.M.A., B.A., A.A. (Ahlam Alynbiawi), S.A.A. (Souad A. Almuthree), N.A.K., W.A.A., Z.A.B. and A.S.B.; supervision, A.A.R. project administration, A.A.R. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Cairns, T.C.; Nai, C.; Meyer, V. How a fungus shapes biotechnology: 100 years of Aspergillus niger research. Fungal Biol. Biotechnol. 2018, 5, 13. [Google Scholar] [CrossRef] [PubMed]
  2. Spagnuolo, M.; Yaguchi, A.; Blenner, M. Oleaginous yeast for biofuel and oleochemical production. Curr. Opin. Biotechnol. 2019, 57, 73–81. [Google Scholar] [CrossRef] [PubMed]
  3. Chang, R. Bioactive polysaccharides from traditional Chinese medicine herbs as anticancer adjuvants. J. Altern. Complement. Med. 2002, 8, 559–565. [Google Scholar] [CrossRef] [PubMed]
  4. Rop, O.; Mlcek, J.; Jurikova, T. Beta-glucans in higher fungi and their health effects. Nutr. Rev. 2009, 67, 624–631. [Google Scholar] [CrossRef] [PubMed]
  5. Pelley, R.P.; Strickland, F.M. Plants, polysaccharides, and the treatment and prevention of neoplasia. Crit. Rev. ™ Oncog. 2000, 11, 189–225. [Google Scholar] [CrossRef]
  6. Leal, F. Fungal Pathogenesis in Humans: The Growing Threat. Genes 2019, 10, 136. [Google Scholar] [CrossRef]
  7. Van Burik, J.-A.H.; Magee, P.T. Aspects of fungal pathogenesis in humans. Annu. Rev. Microbiol. 2001, 55, 743–772. [Google Scholar] [CrossRef]
  8. Bitew, A. Dermatophytosis: Prevalence of dermatophytes and non-dermatophyte fungi from patients attending Arsho advanced medical laboratory, Addis Ababa, Ethiopia. Dermatol. Res. Pract. 2018, 2018, 8164757. [Google Scholar] [CrossRef]
  9. Khan, H.A.; Baig, F.K.; Mehboob, R. Nosocomial infections: Epidemiology, prevention, control and surveillance. Asian Pac. J. Trop. Biomed. 2017, 7, 478–482. [Google Scholar] [CrossRef]
  10. Gnat, S.; Łagowski, D.; Nowakiewicz, A.; Dyląg, M. A global view on fungal infections in humans and animals: Opportunistic infections and microsporidioses. J. Appl. Microbiol. 2021, 131, 2095–2113. [Google Scholar] [CrossRef]
  11. Bongomin, F.; Gago, S.; Oladele, R.O.; Denning, D.W. Global and multi-national prevalence of fungal diseases—Estimate precision. J. Fungi 2017, 3, 57. [Google Scholar] [CrossRef]
  12. Benedict, K.; Jackson, B.R.; Chiller, T.; Beer, K.D. Estimation of direct healthcare costs of fungal diseases in the United States. Clin. Infect. Dis. 2019, 68, 1791–1797. [Google Scholar] [CrossRef]
  13. Almeida, F.; Rodrigues, M.L.; Coelho, C. The still underestimated problem of fungal diseases worldwide. Front. Microbiol. 2019, 10, 214. [Google Scholar] [CrossRef]
  14. Darwin, J. Improving Outcomes for Patients with Fungal Infections across the World a Road Map for the Next Decade; Global Action Fund for Fungal Infections: Geneva, Switzerland, 2015. [Google Scholar]
  15. Köhler, J.R.; Casadevall, A.; Perfect, J. The spectrum of fungi that infects humans. Cold Spring Harb. Perspect. Med. 2015, 5, a019273. [Google Scholar] [CrossRef]
  16. Kwon-Chung, K.J.; Fraser, J.A.; Doering, T.L.; Wang, Z.A.; Janbon, G.; Idnurm, A.; Bahn, Y.-S. Cryptococcus neoformans and Cryptococcus gattii, the etiologic agents of cryptococcosis. Cold Spring Harb. Perspect. Med. 2014, 4, a019760. [Google Scholar] [CrossRef]
  17. Limper, A.H.; Adenis, A.; Le, T.; Harrison, T.S. Fungal infections in HIV/AIDS. Lancet Infect. Dis. 2017, 17, e334–e343. [Google Scholar] [CrossRef]
  18. Sipsas, N.V.; Kontoyiannis, D.P. Invasive fungal infections in patients with cancer in the Intensive Care Unit. Int. J. Antimicrob. Agents 2012, 39, 464–471. [Google Scholar] [CrossRef]
  19. Martinez, R. New trends in paracoccidioidomycosis epidemiology. J. Fungi 2017, 3, 1. [Google Scholar] [CrossRef]
  20. Pachl, J.; Svoboda, P.; Jacobs, F.; Vandewoude, K.; Van Der Hoven, B.; Spronk, P.; Masterson, G.; Malbrain, M.; Aoun, M.; Garbino, J. A randomized, blinded, multicenter trial of lipid-associated amphotericin B alone versus in combination with an antibody-based inhibitor of heat shock protein 90 in patients with invasive candidiasis. Clin. Infect. Dis. 2006, 42, 1404–1413. [Google Scholar] [CrossRef]
  21. Wharton, R.E.; Stefanov, E.K.; King, R.G.; Kearney, J.F. Antibodies generated against Streptococci protect in a mouse model of disseminated aspergillosis. J. Immunol. 2015, 194, 4387–4396. [Google Scholar] [CrossRef]
  22. Strickland, A.B.; Shi, M. Mechanisms of fungal dissemination. Cell. Mol. Life Sci. 2021, 78, 3219–3238. [Google Scholar] [CrossRef] [PubMed]
  23. Moyes, D.L.; Richardson, J.P.; Naglik, J.R. Candida albicans-epithelial interactions and pathogenicity mechanisms: Scratching the surface. Virulence 2015, 6, 338–346. [Google Scholar] [CrossRef] [PubMed]
  24. Dagenais, T.R.; Keller, N.P. Pathogenesis of Aspergillus fumigatus in invasive aspergillosis. Clin. Microbiol. Rev. 2009, 22, 447–465. [Google Scholar] [CrossRef] [PubMed]
  25. Mittal, J.; Ponce, M.G.; Gendlina, I.; Nosanchuk, J.D. Histoplasma capsulatum: Mechanisms for pathogenesis. Fungal Physiol. Immunopathogenesis 2019, 422, 157–191. [Google Scholar]
  26. Richardson, J.P.; Moyes, D.L. Adaptive immune responses to Candida albicans infection. Virulence 2015, 6, 327–337. [Google Scholar] [CrossRef]
  27. Romani, L. Immunity to fungal infections. Nat. Rev. Immunol. 2004, 4, 11–24. [Google Scholar] [CrossRef]
  28. Heninger, E.; Hogan, L.H.; Karman, J.; Macvilay, S.; Hill, B.; Woods, J.P.; Sandor, M. Characterization of the Histoplasma capsulatum-induced granuloma. J. Immunol. 2006, 177, 3303–3313. [Google Scholar] [CrossRef]
  29. Shibuya, K.; Hirata, A.; Omuta, J.; Sugamata, M.; Katori, S.; Saito, N.; Murata, N.; Morita, A.; Takahashi, K.; Hasegawa, C. Granuloma and cryptococcosis. J. Infect. Chemother. 2005, 11, 115–122. [Google Scholar] [CrossRef]
  30. Netea, M.G.; Gijzen, K.; Coolen, N.; Verschueren, I.; Figdor, C.; Van der Meer, J.W.; Torensma, R.; Kullberg, B.J. Human dendritic cells are less potent at killing Candida albicans than both monocytes and macrophages. Microbes Infect. 2004, 6, 985–989. [Google Scholar] [CrossRef]
  31. Ramirez-Ortiz, Z.G.; Lee, C.K.; Wang, J.P.; Boon, L.; Specht, C.A.; Levitz, S.M. A nonredundant role for plasmacytoid dendritic cells in host defense against the human fungal pathogen Aspergillus fumigatus. Cell Host Microbe 2011, 9, 415–424. [Google Scholar] [CrossRef]
  32. Ziegler-Heitbrock, L.; Ancuta, P.; Crowe, S.; Dalod, M.; Grau, V.; Hart, D.N.; Leenen, P.J.; Liu, Y.-J.; MacPherson, G.; Randolph, G.J. Nomenclature of monocytes and dendritic cells in blood. Blood J. Am. Soc. Hematol. 2010, 116, e74–e80. [Google Scholar] [CrossRef]
  33. Smeekens, S.P.; van de Veerdonk, F.L.; Joosten, L.A.; Jacobs, L.; Jansen, T.; Williams, D.L.; van der Meer, J.W.; Kullberg, B.J.; Netea, M.G. The classical CD14++ CD16− monocytes, but not the patrolling CD14+ CD16+ monocytes, promote Th17 responses to Candida albicans. Eur. J. Immunol. 2011, 41, 2915–2924. [Google Scholar] [CrossRef]
  34. Campbell, K.S.; Hasegawa, J. Natural killer cell biology: An update and future directions. J. Allergy Clin. Immunol. 2013, 132, 536–544. [Google Scholar] [CrossRef]
  35. Voigt, J.; Hünniger, K.; Bouzani, M.; Jacobsen, I.D.; Barz, D.; Hube, B.; Löffler, J.; Kurzai, O. Human natural killer cells acting as phagocytes against Candida albicans and mounting an inflammatory response that modulates neutrophil antifungal activity. J. Infect. Dis. 2014, 209, 616–626. [Google Scholar] [CrossRef]
  36. Schmidt, S.; Tramsen, L.; Hanisch, M.; Latgé, J.-P.; Huenecke, S.; Koehl, U.; Lehrnbecher, T. Human natural killer cells exhibit direct activity against Aspergillus fumigatus hyphae, but not against resting conidia. J. Infect. Dis. 2011, 203, 430–435. [Google Scholar] [CrossRef]
  37. Nakamatsu, M.; Yamamoto, N.; Hatta, M.; Nakasone, C.; Kinjo, T.; Miyagi, K.; Uezu, K.; Nakamura, K.; Nakayama, T.; Taniguchi, M. Role of interferon-γ in Vα14+ natural killer T cell-mediated host defense against Streptococcus pneumoniae infection in murine lungs. Microbes Infect. 2007, 9, 364–374. [Google Scholar] [CrossRef]
  38. Cohen, N.R.; Tatituri, R.V.; Rivera, A.; Watts, G.F.; Kim, E.Y.; Chiba, A.; Fuchs, B.B.; Mylonakis, E.; Besra, G.S.; Levitz, S.M. Innate recognition of cell wall β-glucans drives invariant natural killer T cell responses against fungi. Cell Host Microbe 2011, 10, 437–450. [Google Scholar] [CrossRef]
  39. Horn, D.L.; Neofytos, D.; Anaissie, E.J.; Fishman, J.A.; Steinbach, W.J.; Olyaei, A.J.; Marr, K.A.; Pfaller, M.A.; Chang, C.-H.; Webster, K.M. Epidemiology and outcomes of candidemia in 2019 patients: Data from the prospective antifungal therapy alliance registry. Clin. Infect. Dis. 2009, 48, 1695–1703. [Google Scholar] [CrossRef]
  40. Feldmesser, M. Prospects of vaccines for medically important fungi. Sabouraudia 2005, 43, 571–587. [Google Scholar] [CrossRef]
  41. Gabrilovich, D.I.; Nagaraj, S. Myeloid-derived suppressor cells as regulators of the immune system. Nat. Rev. Immunol. 2009, 9, 162–174. [Google Scholar] [CrossRef]
  42. Hayes, B.M.; Bleackley, M.R.; Wiltshire, J.L.; Anderson, M.A.; Traven, A.; van der Weerden, N.L. Identification and mechanism of action of the plant defensin NaD1 as a new member of the antifungal drug arsenal against Candida albicans. Antimicrob. Agents Chemother. 2013, 57, 3667–3675. [Google Scholar] [CrossRef] [PubMed]
  43. LeibundGut-Landmann, S.; Wüthrich, M.; Hohl, T.M. Immunity to fungi. Curr. Opin. Immunol. 2012, 24, 449–458. [Google Scholar] [CrossRef] [PubMed]
  44. Lindell, D.M.; Moore, T.A.; McDonald, R.A.; Toews, G.B.; Huffnagle, G.B. Generation of antifungal effector CD8+ T cells in the absence of CD4+ T cells during Cryptococcus neoformans infection. J. Immunol. 2005, 174, 7920–7928. [Google Scholar] [CrossRef] [PubMed]
  45. Nanjappa, S.G.; Heninger, E.; Wüthrich, M.; Gasper, D.J.; Klein, B.S. Tc17 cells mediate vaccine immunity against lethal fungal pneumonia in immune deficient hosts lacking CD4+ T cells. PLoS Pathog. 2012, 8, e1002771. [Google Scholar] [CrossRef] [PubMed]
  46. Berland, R.; Wortis, H.H. Origins and Functions of B-1 Cells. Annu. Rev. Immunol 2002, 20, 253–300. [Google Scholar] [CrossRef]
  47. Hardy, R.R. B-1 B cells: Development, selection, natural autoantibody and leukemia. Curr. Opin. Immunol. 2006, 18, 547–555. [Google Scholar] [CrossRef]
  48. Kohler, H.; Bayry, J.; Nicoletti, A.; Kaveri, S. Natural autoantibodies as tools to predict the outcome of immune response? Scand. J. Immunol. 2003, 58, 285–289. [Google Scholar] [CrossRef]
  49. Zhou, Z.-h.; Wild, T.; Xiong, Y.; Sylvers, L.H.; Zhang, Y.; Zhang, L.; Wahl, L.; Wahl, S.M.; Kozlowski, S.; Notkins, A.L. Polyreactive antibodies plus complement enhance the phagocytosis of cells made apoptotic by UV-light or HIV. Sci. Rep. 2013, 3, 2271. [Google Scholar] [CrossRef]
  50. Notkins, A.L. Polyreactivity of antibody molecules. Trends Immunol. 2004, 25, 174–179. [Google Scholar] [CrossRef]
  51. Ehrenstein, M.R.; Notley, C.A. The importance of natural IgM: Scavenger, protector and regulator. Nat. Rev. Immunol. 2010, 10, 778–786. [Google Scholar] [CrossRef]
  52. Coutinho, A.; Kazatchkine, M.D.; Avrameas, S. Natural autoantibodies. Curr. Opin. Immunol. 1995, 7, 812–818. [Google Scholar] [CrossRef]
  53. López-Ribot, J.L.; Casanova, M.; Murgui, A.; Martínez, J.P. Antibody response to Candida albicans cell wall antigens. FEMS Immunol. Med. Microbiol. 2004, 41, 187–196. [Google Scholar] [CrossRef]
  54. Kuruganti, U.; Henderson, L.A.; Garner, R.E.; Asofsky, R.; Baker, P.J.; Domer, J.E. Nonspecific and Candida-Specific Immune Responses in Mice Suppressed by Chronic Administration of Anti-μ. J. Leukoc. Biol. 1988, 44, 422–433. [Google Scholar] [CrossRef]
  55. Han, Y.; Morrison, R.P.; Cutler, J.E. A vaccine and monoclonal antibodies that enhance mouse resistance to Candida albicans vaginal infection. Infect. Immun. 1998, 66, 5771–5776. [Google Scholar] [CrossRef]
  56. Mukherjee, J.; Scharff, M.D.; Casadevall, A. Protective murine monoclonal antibodies to Cryptococcus neoformans. Infect. Immun. 1992, 60, 4534–4541. [Google Scholar] [CrossRef]
  57. Montagnoli, C.; Bozza, S.; Bacci, A.; Gaziano, R.; Mosci, P.; Morschhäuser, J.; Pitzurra, L.; Kopf, M.; Cutler, J.; Romani, L. A role for antibodies in the generation of memory antifungal immunity. Eur. J. Immunol. 2003, 33, 1193–1204. [Google Scholar] [CrossRef]
  58. Li, W.; Fu, M.; An, J.G.; Xing, Y.; Zhang, P.; Zhang, X.; Wang, Y.C.; Li, C.X.; Tian, R.; Su, W.J. Host defence against C. albicans infections in IgH transgenic mice with VH derived from a natural anti-keratin antibody. Cell. Microbiol. 2007, 9, 306–315. [Google Scholar] [CrossRef]
  59. Tian, R.; Fu, M.; Zhang, Z.; Ren, J.; An, J.; Liu, Y.; Li, W. In situ IgM production and clonal expansion of B-1 cells in peritoneal cavity promote elimination of C. albicans infection in IgH transgenic mice with VH derived from a natural antibody. PLoS ONE 2013, 8, e60779. [Google Scholar] [CrossRef]
  60. Zaragoza, O.; Taborda, C.P.; Casadevall, A. The efficacy of complement-mediated phagocytosis of Cryptococcus neoformans is dependent on the location of C3 in the polysaccharide capsule and involves both direct and indirectC3-mediated interactions. Eur. J. Immunol. 2003, 33, 1957–1967. [Google Scholar] [CrossRef]
  61. Shapiro, S.; Beenhouwer, D.O.; Feldmesser, M.; Taborda, C.; Carroll, M.C.; Casadevall, A.; Scharff, M.D. Immunoglobulin G monoclonal antibodies to Cryptococcus neoformans protect mice deficient in complement component C3. Infect. Immun. 2002, 70, 2598–2604. [Google Scholar] [CrossRef]
  62. Casadevall, A.; Pirofski, L.A. Immunoglobulins in defense, pathogenesis, and therapy of fungal diseases. Cell Host Microbe 2012, 11, 447–456. [Google Scholar] [CrossRef] [PubMed]
  63. Van Spriel, A.B.; Sofi, M.; Gartlan, K.H.; van der Schaaf, A.; Verschueren, I.; Torensma, R.; Raymakers, R.A.; Loveland, B.E.; Netea, M.G.; Adema, G.J. The tetraspanin protein CD37 regulates IgA responses and anti-fungal immunity. PLoS Pathog. 2009, 5, e1000338. [Google Scholar] [CrossRef] [PubMed]
  64. Miller, M.F.; Mitchell, T.G.; Storkus, W.; Dawson, J.R. Human natural killer cells do not inhibit growth of Cryptococcus neoformans in the absence of antibody. Infect. Immun. 1990, 58, 639–645. [Google Scholar] [CrossRef] [PubMed]
  65. Nabavi, N.; Murphy, J.W. Antibody-dependent natural killer cell-mediated growth inhibition of Cryptococcus neoformans. Infect. Immun. 1986, 51, 556–562. [Google Scholar] [CrossRef] [PubMed]
  66. Magliani, W.; Conti, S.; Frazzi, R.; Ravanetti, L.; Maffei, D.; Polonelli, L. Protective antifungal yeast killer toxin-like antibodies. Curr. Mol. Med. 2005, 5, 443–452. [Google Scholar] [CrossRef]
  67. Kainz, K.; Bauer, M.A.; Madeo, F.; Carmona-Gutierrez, D. Fungal infections in humans: The silent crisis. Microb. Cell 2020, 7, 143. [Google Scholar] [CrossRef]
  68. Marcos, C.M.; De Oliveira, H.C.; de Melo, W.d.C.M.A.; Da Silva, J.D.F.; Assato, P.A.; Scorzoni, L.; Rossi, S.A.; de Paula e Silva, A.C.; Mendes-Giannini, M.J.; Fusco-Almeida, A.M. Anti-immune strategies of pathogenic fungi. Front. Cell. Infect. Microbiol. 2016, 6, 142. [Google Scholar] [CrossRef]
  69. Patin, E.C.; Thompson, A.; Orr, S.J. Pattern recognition receptors in fungal immunity. Semin. Cell Dev. Biol. 2019, 89, 24–33. [Google Scholar] [CrossRef]
  70. Gupta, S.; Ellis, M.; Cesario, T.; Ruhling, M.; Vayuvegula, B. Disseminated cryptococcal infection in a patient with hypogammaglobulinemia and normal T cell functions. Am. J. Med. 1987, 82, 129–131. [Google Scholar] [CrossRef]
  71. Wahab, J.; Hanifah, M.; Choo, K. Bruton’s agamma globulinaemia in a child presenting with cryptococcal, empyemia thoracis and periauricular pyogenic abscess. Singap. Med. J. 1995, 36, 686–689. [Google Scholar]
  72. Pires, N.; Guimarães, M.C.; Moya, M.J.; Oliveira, F.R.; Louzada, J.; Martinez, R. Hypogammaglobulinemia as predisposing factor for Cryptococcus neoformans infection: Report of two cases. Rev. Da Soc. Bras. De Med. Trop. 2000, 33, 603–608. [Google Scholar]
  73. Agematsu, K.; Nagumo, H.; Shinozaki, K.; Hokibara, S.; Yasui, K.; Terada, K.; Kawamura, N.; Toba, T.; Nonoyama, S.; Ochs, H.D. Absence of IgD-CD27 (+) memory B cell population in X-linked hyper-IgM syndrome. J. Clin. Investig. 1998, 102, 853–860. [Google Scholar] [CrossRef]
  74. Marr, K.A.; Datta, K.; Pirofski, L.-a.; Barnes, R. Cryptococcus gattii infection in healthy hosts: A sentinel for subclinical immunodeficiency? Clin. Infect. Dis. 2012, 54, 153–154. [Google Scholar] [CrossRef]
  75. Abadi, J.; Pirofski, L.-A. Antibodies reactive with the cryptococcal capsular polysaccharide glucuronoxylomannan are present in sera from children with and without human immunodeficiency virus infection. J. Infect. Dis. 1999, 180, 915–919. [Google Scholar] [CrossRef]
  76. Deshaw, M.; Pirofski, L.A. Antibodies to the Cryptococcus neoformans capsular glucuronoxylomannan are ubiquitous in serum from HIV+ and HIV− individuals. Clin. Exp. Immunol. 1995, 99, 425–432. [Google Scholar] [CrossRef]
  77. Pirofski, L.-a. Polysaccharides, mimotopes and vaccines for fungal and encapsulated pathogens. Trends Microbiol. 2001, 9, 445–451. [Google Scholar] [CrossRef]
  78. Subramaniam, K.S.; Datta, K.; Quintero, E.; Manix, C.; Marks, M.S.; Pirofski, L.-A. The absence of serum IgM enhances the susceptibility of mice to pulmonary challenge with Cryptococcus neoformans. J. Immunol. 2010, 184, 5755–5767. [Google Scholar] [CrossRef]
  79. Rapaka, R.R.; Ricks, D.M.; Alcorn, J.F.; Chen, K.; Khader, S.A.; Zheng, M.; Plevy, S.; Bengtén, E.; Kolls, J.K. Conserved natural IgM antibodies mediate innate and adaptive immunity against the opportunistic fungus Pneumocystis murina. J. Exp. Med. 2010, 207, 2907–2919. [Google Scholar] [CrossRef]
  80. Szymczak, W.A.; Davis, M.J.; Lundy, S.K.; Dufaud, C.; Olszewski, M.; Pirofski, L.-a. X-linked immunodeficient mice exhibit enhanced susceptibility to Cryptococcus neoformans infection. MBio 2013, 4, e00265-13. [Google Scholar] [CrossRef]
  81. Moragues, M.D.; Omaetxebarria, M.J.; Elguezabal, N.; Sevilla, M.J.; Conti, S.; Polonelli, L.; Pontón, J. A monoclonal antibody directed against a Candida albicans cell wall mannoprotein exerts three anti-C. albicans activities. Infect. Immun. 2003, 71, 5273–5279. [Google Scholar] [CrossRef]
  82. Torosantucci, A.; Bromuro, C.; Chiani, P.; De Bernardis, F.; Berti, F.; Galli, C.; Norelli, F.; Bellucci, C.; Polonelli, L.; Costantino, P. A novel glyco-conjugate vaccine against fungal pathogens. J. Exp. Med. 2005, 202, 597–606. [Google Scholar] [CrossRef] [PubMed]
  83. Torosantucci, A.; Chiani, P.; Bromuro, C.; De Bernardis, F.; Palma, A.S.; Liu, Y.; Mignogna, G.; Maras, B.; Colone, M.; Stringaro, A. Protection by anti-β-glucan antibodies is associated with restricted β-1,3 glucan binding specificity and inhibition of fungal growth and adherence. PLoS ONE 2009, 4, e5392. [Google Scholar] [CrossRef] [PubMed]
  84. Alvarez, M.; Saylor, C.; Casadevall, A. Antibody action after phagocytosis promotes Cryptococcus neoformans and Cryptococcus gattii macrophage exocytosis with biofilm-like microcolony formation. Cell. Microbiol. 2008, 10, 1622–1633. [Google Scholar] [CrossRef] [PubMed]
  85. Maiti, P.K.; Kumar, A.; Kumar, R.; Mohapatra, L. Role of antibodies and effect of BCG vaccination in experimental candidiasis in mice. Mycopathologia 1985, 91, 79–85. [Google Scholar] [CrossRef]
  86. Cantorna, M.T.; Balish, E. Acquired immunity to systemic candidiasis in immunodeficient mice. J. Infect. Dis. 1991, 164, 936–943. [Google Scholar] [CrossRef]
  87. Narayanan, R.; Joyce, W.; Greenfield, R. Gastrointestinal candidiasis in a murine model of severe combined immunodeficiency syndrome. Infect. Immun. 1991, 59, 2116–2119. [Google Scholar] [CrossRef]
  88. Carrow, E.W.; Hector, R.; Domer, J.E. Immunodeficient CBA/N mice respond effectively to Candida albicans. Clin. Immunol. Immunopathol. 1984, 33, 371–380. [Google Scholar] [CrossRef]
  89. Sinha, B.; Prasad, S.; Monga, D. Studies of the role of B-cells in the resistance of mice to experimental candidiasis. Zent. Für Bakteriol. Mikrobiol. Und Hygiene. Ser. A Med. Microbiol. Infect. Dis. Virol. Parasitol. 1987, 266, 316–322. [Google Scholar] [CrossRef]
  90. Tristão, F.; Panagio, L.; Rocha, F.; Cavassani, K.; Moreira, A.; Rossi, M.; Silva, J. B cell-deficient mice display enhanced susceptibility to Paracoccidioides brasiliensis infection. Mycopathologia 2013, 176, 1–10. [Google Scholar] [CrossRef]
  91. K Mazu, T.; A Bricker, B.; Flores-Rozas, H.; Y Ablordeppey, S. The mechanistic targets of antifungal agents: An overview. Mini Rev. Med. Chem. 2016, 16, 555–578. [Google Scholar] [CrossRef]
  92. Nami, S.; Mohammadi, R.; Vakili, M.; Khezripour, K.; Mirzaei, H.; Morovati, H. Fungal vaccines, mechanism of actions and immunology: A comprehensive review. Biomed. Pharmacother. 2019, 109, 333–344. [Google Scholar] [CrossRef]
  93. Schmidt, C.S.; White, C.J.; Ibrahim, A.S.; Filler, S.G.; Fu, Y.; Yeaman, M.R.; Edwards, J.E., Jr.; Hennessey, J.P., Jr. NDV-3, a recombinant alum-adjuvanted vaccine for Candida and Staphylococcus aureus, is safe and immunogenic in healthy adults. Vaccine 2012, 30, 7594–7600. [Google Scholar] [CrossRef]
  94. Spellberg, B.J.; Ibrahim, A.S.; Avanesian, V.; Fu, Y.; Myers, C.; Phan, Q.T.; Filler, S.G.; Yeaman, M.R.; Edwards, J.E., Jr. Efficacy of the anti-Candida rAls3p-N or rAls1p-N vaccines against disseminated and mucosal candidiasis. J. Infect. Dis. 2006, 194, 256–260. [Google Scholar] [CrossRef]
  95. De Bernardis, F.; Amacker, M.; Arancia, S.; Sandini, S.; Gremion, C.; Zurbriggen, R.; Moser, C.; Cassone, A. A virosomal vaccine against candidal vaginitis: Immunogenicity, efficacy and safety profile in animal models. Vaccine 2012, 30, 4490–4498. [Google Scholar] [CrossRef]
  96. Ibrahim, A.S.; Spellberg, B.J.; Avanesian, V.; Fu, Y.; Edwards, J.E., Jr. The anti-Candida vaccine based on the recombinant N-terminal domain of Als1p is broadly active against disseminated candidiasis. Infect. Immun. 2006, 74, 3039–3041. [Google Scholar] [CrossRef]
  97. Baquir, B.; Lin, L.; Ibrahim, A.S.; Fu, Y.; Avanesian, V.; Tu, A.; Edwards, J., Jr.; Spellberg, B. Immunological reactivity of blood from healthy humans to the rAls3p-N vaccine protein. J. Infect. Dis. 2010, 201, 473–477. [Google Scholar] [CrossRef]
  98. Bozza, S.; Clavaud, C.; Giovannini, G.; Fontaine, T.; Beauvais, A.; Sarfati, J.; D’Angelo, C.; Perruccio, K.; Bonifazi, P.; Zagarella, S. Immune sensing of Aspergillus fumigatus proteins, glycolipids, and polysaccharides and the impact on Th immunity and vaccination. J. Immunol. 2009, 183, 2407–2414. [Google Scholar] [CrossRef]
  99. Sandini, S.; La Valle, R.; Deaglio, S.; Malavasi, F.; Cassone, A.; De Bernardis, F. A highly immunogenic recombinant and truncated protein of the secreted aspartic proteases family (rSap2t) of Candida albicans as a mucosal anticandidal vaccine. FEMS Immunol. Med. Microbiol. 2011, 62, 215–224. [Google Scholar] [CrossRef]
  100. Saville, S.P.; Lazzell, A.L.; Chaturvedi, A.K.; Monteagudo, C.; Lopez-Ribot, J.L. Efficacy of a genetically engineered Candida albicans tet-NRG1 strain as an experimental live attenuated vaccine against hematogenously disseminated candidiasis. Clin. Vaccine Immunol. 2009, 16, 430–432. [Google Scholar] [CrossRef]
  101. Martínez-López, R.; Nombela, C.; Diez-Orejas, R.; Monteoliva, L.; Gil, C. Immunoproteomic analysis of the protective response obtained from vaccination with Candida albicans ecm33 cell wall mutant in mice. Proteomics 2008, 8, 2651–2664. [Google Scholar] [CrossRef]
  102. Bistoni, F.; Vecchiarelli, A.; Cenci, E.; Puccetti, P.; Marconi, P.; Cassone, A. Evidence for macrophage-mediated protection against lethal Candida albicans infection. Infect. Immun. 1986, 51, 668–674. [Google Scholar] [CrossRef] [PubMed]
  103. Carneiro, C.; Correia, A.; Lima, T.; Vilanova, M.; Pais, C.; Gomes, A.C.; Oliveira, M.E.C.R.; Sampaio, P. Protective effect of antigen delivery using monoolein-based liposomes in experimental hematogenously disseminated candidiasis. Acta Biomater. 2016, 39, 133–145. [Google Scholar] [CrossRef] [PubMed]
  104. Han, Y.; Ulrich, M.A.; Cutler, J.E. Candida albicans mannan extract—Protein conjugates induce a protective immune response against experimental candidiasis. J. Infect. Dis. 1999, 179, 1477–1484. [Google Scholar] [CrossRef] [PubMed]
  105. Pietrella, D.; Rachini, A.; Torosantucci, A.; Chiani, P.; Brown, A.J.; Bistoni, F.; Costantino, P.; Mosci, P.; d’Enfert, C.; Rappuoli, R. A β-glucan-conjugate vaccine and anti-β-glucan antibodies are effective against murine vaginal candidiasis as assessed by a novel in vivo imaging technique. Vaccine 2010, 28, 1717–1725. [Google Scholar] [CrossRef]
  106. Chiani, P.; Bromuro, C.; Cassone, A.; Torosantucci, A. Anti-β-glucan antibodies in healthy human subjects. Vaccine 2009, 27, 513–519. [Google Scholar] [CrossRef]
  107. Bromuro, C.; Romano, M.; Chiani, P.; Berti, F.; Tontini, M.; Proietti, D.; Mori, E.; Torosantucci, A.; Costantino, P.; Rappuoli, R. Beta-glucan-CRM197 conjugates as candidates antifungal vaccines. Vaccine 2010, 28, 2615–2623. [Google Scholar] [CrossRef]
  108. Paulovičová, E.; Machová, E.; Tulinská, J.; Bystrický, S. Cell and antibody mediated immunity induced by vaccination with novel Candida dubliniensis mannan immunogenic conjugate. Int. Immunopharmacol. 2007, 7, 1325–1333. [Google Scholar] [CrossRef]
  109. Cutler, J.E.; Corti, M.; Lambert, P.; Ferris, M.; Xin, H. Horizontal transmission of Candida albicans and evidence of a vaccine response in mice colonized with the fungus. PLoS ONE 2011, 6, e22030. [Google Scholar] [CrossRef]
  110. Xin, H.; Cutler, J.E. Vaccine and monoclonal antibody that enhance mouse resistance to candidiasis. Clin. Vaccine Immunol. 2011, 18, 1656–1667. [Google Scholar] [CrossRef]
  111. Cárdenas-Freytag, L.; Cheng, E.; Mayeux, P.; Domer, J.E.; Clements, J.D. Effectiveness of a vaccine composed of heat-killed Candida albicans and a novel mucosal adjuvant, LT (R192G), against systemic candidiasis. Infect. Immun. 1999, 67, 826–833. [Google Scholar] [CrossRef]
  112. Luo, G.; Ibrahim, A.S.; French, S.W.; Edwards, J.E., Jr.; Fu, Y. Active and passive immunization with rHyr1p-N protects mice against hematogenously disseminated candidiasis. PLoS ONE 2011, 6, e25909. [Google Scholar] [CrossRef]
  113. Xin, H.; Dziadek, S.; Bundle, D.R.; Cutler, J.E. Synthetic glycopeptide vaccines combining β-mannan and peptide epitopes induce protection against candidiasis. Proc. Natl. Acad. Sci. USA 2008, 105, 13526–13531. [Google Scholar] [CrossRef]
  114. Cenci, E.; Mencacci, A.; Bacci, A.; Bistoni, F.; Kurup, V.P.; Romani, L. T cell vaccination in mice with invasive pulmonary aspergillosis. J. Immunol. 2000, 165, 381–388. [Google Scholar] [CrossRef]
  115. Diaz-Arevalo, D.; Bagramyan, K.; Hong, T.B.; Ito, J.I.; Kalkum, M. CD4+ T cells mediate the protective effect of the recombinant Asp f3-based anti-aspergillosis vaccine. Infect. Immun. 2011, 79, 2257–2266. [Google Scholar] [CrossRef]
  116. Ito, J.I.; Lyons, J.M. Vaccination of corticosteroid immunosuppressed mice against invasive pulmonary aspergillosis. J. Infect. Dis. 2002, 186, 869–871. [Google Scholar] [CrossRef]
  117. Ito, J.I.; Lyons, J.M.; Hong, T.B.; Tamae, D.; Liu, Y.-K.; Wilczynski, S.P.; Kalkum, M. Vaccinations with recombinant variants of Aspergillus fumigatus allergen Asp f 3 protect mice against invasive aspergillosis. Infect. Immun. 2006, 74, 5075–5084. [Google Scholar] [CrossRef]
  118. Clemons, K.V.; Martinez, M.; Chen, V.; Liu, M.; Yoon, H.J.; Stevens, D.A. Protection against experimental aspergillosis by heat-killed yeast is not antibody dependent. Sabouraudia 2014, 52, 422–426. [Google Scholar] [CrossRef]
  119. Stuehler, C.; Khanna, N.; Bozza, S.; Zelante, T.; Moretti, S.; Kruhm, M.; Lurati, S.; Conrad, B.; Worschech, E.; Stevanović, S. Cross-protective TH1 immunity against Aspergillus fumigatus and Candida albicans. Blood J. Am. Soc. Hematol. 2011, 117, 5881–5891. [Google Scholar] [CrossRef]
  120. Bozza, S.; Gaziano, R.; Lipford, G.B.; Montagnoli, C.; Bacci, A.; Di Francesco, P.; Kurup, V.P.; Wagner, H.; Romani, L. Vaccination of mice against invasive aspergillosis with recombinant Aspergillus proteins and CpG oligodeoxynucleotides as adjuvants. Microbes Infect. 2002, 4, 1281–1290. [Google Scholar] [CrossRef]
  121. Wüthrich, M.; Filutowicz, H.I.; Warner, T.; Deepe, G.S., Jr.; Klein, B.S. Vaccine immunity to pathogenic fungi overcomes the requirement for CD4 help in exogenous antigen presentation to CD8+ T cells: Implications for vaccine development in immune-deficient hosts. J. Exp. Med. 2003, 197, 1405–1416. [Google Scholar] [CrossRef]
  122. Pinto, A.R.; Puccia, R.; Diniz, S.N.; Franco, M.F.; Travassos, L.R. DNA-based vaccination against murine paracoccidioidomycosis using the gp43 gene from Paracoccidioides brasiliensis. Vaccine 2000, 18, 3050–3058. [Google Scholar] [CrossRef] [PubMed]
  123. Assis-Marques, M.A.; Oliveira, A.F.; Ruas, L.P.; Reis, T.F.d.; Roque-Barreira, M.C.; Coelho, P.S.R. Saccharomyces cerevisiae expressing Gp43 protects mice against Paracoccidioides brasiliensis infection. PLoS ONE 2015, 10, e0120201. [Google Scholar] [CrossRef] [PubMed]
  124. Braga, C.J.; Rittner, G.M.; Munoz Henao, J.E.; Teixeira, A.F.; Massis, L.M.; Sbrogio-Almeida, M.E.; Taborda, C.P.; Travassos, L.R.; Ferreira, L.C. Paracoccidioides brasiliensis vaccine formulations based on the gp43-derived P10 sequence and the Salmonella enterica FliC flagellin. Infect. Immun. 2009, 77, 1700–1707. [Google Scholar] [CrossRef] [PubMed]
  125. Fernandes, V.C.; Martins, E.M.; Boeloni, J.N.; Coitinho, J.B.; Serakides, R.; Goes, A.M. Additive effect of rPb27 immunization and chemotherapy in experimental paracoccidioidomycosis. PLoS ONE 2011, 6, e17885. [Google Scholar] [CrossRef]
  126. De Bastos Ascenço Soares, R.; Gomez, F.J.; de Almeida Soares, C.M.; Deepe, G.S., Jr. Vaccination with heat shock protein 60 induces a protective immune response against experimental Paracoccidioides brasiliensis pulmonary infection. Infect. Immun. 2008, 76, 4214–4221. [Google Scholar] [CrossRef]
  127. Ribeiro, A.M.; Bocca, A.L.; Amaral, A.C.; Souza, A.C.C.; Faccioli, L.H.; Coelho-Castelo, A.A.; Figueiredo, F.; Silva, C.L.; Felipe, M.S.S. HSP65 DNA as therapeutic strategy to treat experimental paracoccidioidomycosis. Vaccine 2010, 28, 1528–1534. [Google Scholar] [CrossRef]
  128. Devi, S.J. Preclinical efficacy of a glucuronoxylomannan-tetanus toxoid conjugate vaccine of Cryptococcus neoformans in a murine model. Vaccine 1996, 14, 841–844. [Google Scholar] [CrossRef]
  129. Chow, S.-K.; Casadevall, A. Evaluation of Cryptococcus neoformans galactoxylomannan–protein conjugate as vaccine candidate against murine cryptococcosis. Vaccine 2011, 29, 1891–1898. [Google Scholar] [CrossRef]
  130. Wozniak, K.L.; Young, M.L.; Wormley, F.L., Jr. Protective immunity against experimental pulmonary cryptococcosis in T cell-depleted mice. Clin. Vaccine Immunol. 2011, 18, 717–723. [Google Scholar] [CrossRef]
  131. Rella, A.; Mor, V.; Farnoud, A.M.; Singh, A.; Shamseddine, A.A.; Ivanova, E.; Carpino, N.; Montagna, M.T.; Luberto, C.; Del Poeta, M. Role of Sterylglucosidase 1 (Sgl1) on the pathogenicity of Cryptococcus neoformans: Potential applications for vaccine development. Front. Microbiol. 2015, 6, 836. [Google Scholar] [CrossRef]
  132. Specht, C.A.; Nong, S.; Dan, J.M.; Lee, C.K.; Levitz, S.M. Contribution of glycosylation to T cell responses stimulated by recombinant Cryptococcus neoformans mannoprotein. J. Infect. Dis. 2007, 196, 796–800. [Google Scholar] [CrossRef]
  133. Oscarson, S.; Alpe, M.; Svahnberg, P.; Nakouzi, A.; Casadevall, A. Synthesis and immunological studies of glycoconjugates of Cryptococcus neoformans capsular glucuronoxylomannan oligosaccharide structures. Vaccine 2005, 23, 3961–3972. [Google Scholar] [CrossRef]
  134. Rachini, A.; Pietrella, D.; Lupo, P.; Torosantucci, A.; Chiani, P.; Bromuro, C.; Proietti, C.; Bistoni, F.; Cassone, A.; Vecchiarelli, A. An anti-β-glucan monoclonal antibody inhibits growth and capsule formation of Cryptococcus neoformans in vitro and exerts therapeutic, anticryptococcal activity in vivo. Infect. Immun. 2007, 75, 5085–5094. [Google Scholar] [CrossRef]
  135. Zheng, M.; Ramsay, A.J.; Robichaux, M.B.; Norris, K.A.; Kliment, C.; Crowe, C.; Rapaka, R.R.; Steele, C.; McAllister, F.; Shellito, J.E. CD4+ T cell–independent DNA vaccination against opportunistic infections. J. Clin. Investig. 2005, 115, 3536–3544. [Google Scholar] [CrossRef]
  136. Feng, Y.; Guo, S.; Jiang, T.; Han, X.; Liu, P.; Wu, T.; Luo, Y. Active immunization against Pneumocystis carinii with p55-v3 DNA vaccine in rats. Can. J. Microbiol. 2011, 57, 375–381. [Google Scholar] [CrossRef]
  137. Theus, S.A.; Smulian, A.G.; Steele, P.; Linke, M.J.; Walzer, P.D. Immunization with the major surface glycoprotein of Pneumocystis carinii elicits a protective response. Vaccine 1998, 16, 1149–1157. [Google Scholar] [CrossRef]
  138. Garcia, J.P.; Howard, D.H. Characterization of antigens from the yeast phase of Histoplasma capsulatum. Infect. Immun. 1971, 4, 116–125. [Google Scholar] [CrossRef]
  139. Tewari, R.; Sharma, D.; Solotorovsky, M.; Lafemina, R.; Balint, J. Adoptive transfer of immunity from mice immunized with ribosomes or live yeast cells of Histoplasma capsulatum. Infect. Immun. 1977, 15, 789–795. [Google Scholar] [CrossRef]
  140. Nosanchuk, J.D.; Steenbergen, J.N.; Shi, L.; Deepe, G.S.; Casadevall, A. Antibodies to a cell surface histone-like protein protect against Histoplasma capsulatum. J. Clin. Investig. 2003, 112, 1164–1175. [Google Scholar] [CrossRef]
  141. Deepe, J.; George, S.; Gibbons, R.S. Cellular and molecular regulation of vaccination with heat shock protein 60 from Histoplasma capsulatum. Infect. Immun. 2002, 70, 3759–3767. [Google Scholar] [CrossRef]
  142. Gomez, F.; Gomez, A.; Deepe, G., Jr. Protective efficacy of a 62-kilodalton antigen, HIS-62, from the cell wall and cell membrane of Histoplasma capsulatum yeast cells. Infect. Immun. 1991, 59, 4459–4464. [Google Scholar] [CrossRef] [PubMed]
  143. Gomez, F.J.; Gomez, A.; Deepe, G.S., Jr. An 80-kilodalton antigen from Histoplasma capsulatum that has homology to heat shock protein 70 induces cell-mediated immune responses and protection in mice. Infect. Immun. 1992, 60, 2565–2571. [Google Scholar] [CrossRef] [PubMed]
  144. Scheckelhoff, M.R.; Deepe, G.S. Pulmonary Vβ4+ T Cells from Histoplasma capsulatum–Infected Mice Respond to a Homologue of Sec31 That Confers a Protective Response. J. Infect. Dis. 2006, 193, 888–897. [Google Scholar] [CrossRef] [PubMed]
  145. Deepe, G.S., Jr.; Gibbons, R. Protective efficacy of H antigen from Histoplasma capsulatum in a murine model of pulmonary histoplasmosis. Infect. Immun. 2001, 69, 3128–3134. [Google Scholar] [CrossRef]
  146. Cheah, H.-L.; Lim, V.; Sandai, D. Inhibitors of the glyoxylate cycle enzyme ICL1 in Candida albicans for potential use as antifungal agents. PLoS ONE 2014, 9, e95951. [Google Scholar] [CrossRef]
  147. Gallis, H.A.; Drew, R.H.; Pickard, W.W. Amphotericin B: 30 years of clinical experience. Rev. Infect. Dis. 1990, 12, 308–329. [Google Scholar] [CrossRef]
  148. Bolard, J. How do the polyene macrolide antibiotics affect the cellular membrane properties? Biochim. Et Biophys. Acta (BBA)-Rev. Biomembr. 1986, 864, 257–304. [Google Scholar] [CrossRef]
  149. Villa, N.Y.; Moussatche, P.; Chamberlin, S.G.; Kumar, A.; Lyons, T.J. Phylogenetic and preliminary phenotypic analysis of yeast PAQR receptors: Potential antifungal targets. J. Mol. Evol. 2011, 73, 134–152. [Google Scholar] [CrossRef]
  150. Warnock, D.W. Amphotericin B: An introduction. J. Antimicrob. Chemother. 1991, 28, 27–38. [Google Scholar] [CrossRef]
  151. Delmas, G.; Park, S.; Chen, Z.; Tan, F.; Kashiwazaki, R.; Zarif, L.; Perlin, D. Efficacy of orally delivered cochleates containing amphotericin B in a murine model of aspergillosis. Antimicrob. Agents Chemother. 2002, 46, 2704–2707. [Google Scholar] [CrossRef]
  152. Di Mambro, T.; Guerriero, I.; Aurisicchio, L.; Magnani, M.; Marra, E. The yin and yang of current antifungal therapeutic strategies: How can we harness our natural defenses? Front. Pharmacol. 2019, 10, 80. [Google Scholar] [CrossRef]
  153. Hata, K.; Kimura, J.; Miki, H.; Toyosawa, T.; Nakamura, T.; Katsu, K. In vitro and in vivo antifungal activities of ER-30346, a novel oral triazole with a broad antifungal spectrum. Antimicrob. Agents Chemother. 1996, 40, 2237–2242. [Google Scholar] [CrossRef]
  154. Patton, J.L.; Lester, R.L. The phosphoinositol sphingolipids of Saccharomyces cerevisiae are highly localized in the plasma membrane. J. Bacteriol. 1991, 173, 3101–3108. [Google Scholar] [CrossRef]
  155. Nimrichter, L.; Rodrigues, M.L. Fungal glucosylceramides: From structural components to biologically active targets of new antimicrobials. Front. Microbiol. 2011, 2, 212. [Google Scholar] [CrossRef]
  156. Monk, B.C.; Mason, A.B.; Abramochkin, G.; Haber, J.E.; Seto-Young, D.; Perlin, D.S. The yeast plasma membrane proton pumping ATPase is a viable antifungal target. I. Effects of the cysteine-modifying reagent omeprazole. Biochim. Biophys. Acta (BBA)-Biomembr. 1995, 1239, 81–90. [Google Scholar] [CrossRef]
  157. Chen, S.C.-A.; Slavin, M.A.; Sorrell, T.C. Echinocandin antifungal drugs in fungal infections. Drugs 2011, 71, 11–41. [Google Scholar] [CrossRef]
  158. Perlin, D.S. Current perspectives on echinocandin class drugs. Future Microbiol. 2011, 6, 441–457. [Google Scholar] [CrossRef]
  159. Mukhopadhyay, T.; Roy, K.; Bhat, R.; Sawant, S.; Blumbach, J.; Ganguli, B.; Fehlhaber, H.; Kogler, H. Deoxymulundocandin-a new echinocandin type antifungal antibiotic. J. Antibiot. 1992, 45, 618–623. [Google Scholar] [CrossRef]
  160. Walker, S.S.; Xu, Y.; Triantafyllou, I.; Waldman, M.F.; Mendrick, C.; Brown, N.; Mann, P.; Chau, A.; Patel, R.; Bauman, N. Discovery of a novel class of orally active antifungal β-1, 3-d-glucan synthase inhibitors. Antimicrob. Agents Chemother. 2011, 55, 5099–5106. [Google Scholar] [CrossRef]
  161. Shubitz, L.F.; Trinh, H.T.; Perrill, R.H.; Thompson, C.M.; Hanan, N.J.; Galgiani, J.N.; Nix, D.E. Modeling nikkomycin Z dosing and pharmacology in murine pulmonary coccidioidomycosis preparatory to phase 2 clinical trials. J. Infect. Dis. 2014, 209, 1949–1954. [Google Scholar] [CrossRef]
  162. Vijayakumar, E.; Roy, K.; Chatterjee, S.; Deshmukh, S.; Ganguli, B.; Fehlhaber, H.-W.; Kogler, H. Arthrichitin. A new cell wall active metabolite from Arthrinium phaeospermum. J. Org. Chem. 1996, 61, 6591–6593. [Google Scholar] [CrossRef] [PubMed]
  163. Fostel, J.M.; Montgomery, D.A.; Shen, L.L. Characterization of DNA topoisomerase I from Candida albicans as a target for drug discovery. Antimicrob. Agents Chemother. 1992, 36, 2131–2138. [Google Scholar] [CrossRef] [PubMed]
  164. Fostel, J.; Montgomery, D. Identification of the aminocatechol A-3253 as an in vitro poison of DNA topoisomerase I from Candida albicans. Antimicrob. Agents Chemother. 1995, 39, 586–592. [Google Scholar] [CrossRef] [PubMed]
  165. Levitz, S.M. Aspergillus vaccines: Hardly worth studying or worthy of hard study? Sabouraudia 2016, 55, 103–108. [Google Scholar] [CrossRef]
Figure 1. A general mechanism of the Fungal infection and pathogenesis in Humans.
Figure 1. A general mechanism of the Fungal infection and pathogenesis in Humans.
Microorganisms 11 00671 g001
Figure 2. A diagrammatic representation of cellular innate response and humoral response against fungal infections.
Figure 2. A diagrammatic representation of cellular innate response and humoral response against fungal infections.
Microorganisms 11 00671 g002
Table 1. Common Human Pathogenic Fungi.
Table 1. Common Human Pathogenic Fungi.
Fungal PathogenDiseaseOrgans Affected
AspergillusAspergillosisUsually occurs in people with lung diseases or weakened immune systems.
CandidaCandidiasisCommonly affects the mouth and throat, vagina, or the bloodstream.
CryptococcusCryptococcosisInfect the lungs and/or central nervous system (cryptococcal meningitis).
PneumocystisPneumocystis pneumoniaInfect mainly lungs when the host immune system becomes compromised.
MucormycetesAggressive infectionsOrgans such the brain, spleen, heart, and skin by spreading locally and through the circulation.
HistoplasmaHistoplasmosisInfect the lungs and/or central nervous system (cryptococcal meningitis).
Table 2. Load of common life-threatening fungal infections.
Table 2. Load of common life-threatening fungal infections.
Fungal InfectionNumber AffectedCase Fatality RateEstimated Deaths
Cryptococcal meningitis223,000 in AIDS15–20% USA
>50% developing world
180,000 in AIDS
Pneumocystis pneumonia>400,000 in AIDS
>100,000 in non-AIDS
~15% in AIDS with best treatment
~50% in non-AIDS
>200,000 in AIDS
>50,000 non-AIDS
Disseminated histoplasmosis~100,00015–30%, if diagnosed and treated>80,000
Invasive aspergillosis>1,000,000~30% mortality in leukemia in HIC>500,000
Invasive candidiasis>750,000~40% mortality treated>350,000
Chronic pulmonary aspergillosis>3,000,000~15–40% mortality in HIC
~15% mortality in the developed world
>450,000 in non-hospitalized populations
Severe asthma with fungal sensitization (SAFS)>6,500,000<1% but no good figures350,000–489,000 asthma deaths
~50% related to SAFS
Fungal keratitis1.0–1.4 millionBlinding > 60%>600,000 blind eyes
Total~13,500,000 >1,600,000
Table 3. PRR Recognition of Fungal PAMPs.
Table 3. PRR Recognition of Fungal PAMPs.
Host PRRPathogensFungal PAMPs
CLR
Dectin-1
Candidaspp., A. fumigates, Pneumocystisspp., Coccidioidesspp., Fonsecaeaspp., T. rubrumβ (1,3)- glucan
Dectin-2Candidaspp., Malasseziaspp., A. fumigates, P. brasiliensis, Pneumocystisspp., C. posadasii, F. pedrosoi, H. capsulatum, T. rubrum, C. neoformansα-mannan, high mannose structures
Dectin-3C. albicansα-mannan
MincleC. albicans, A. fumigatus, Pneumocystisspp., Fonsecaeaspp., Malasseziaspp.α-mannose
MclC. albicans, B. dermatitidisα-mannose
DC-SIGNC. albicans, A. fumigates, T. rubrum, P. brasiliensisMannose, surface carbohydrate in extracellular vesicles
MRC. neoformans, P. carinii, C. immitis, P. brasiliensis, H. capsulatum, T. rubrumMannose
Mannose receptorC. albicans, C. neoformans, A. fumigates, P. brasiliensisMannan (N-linked), mannoproteins, mannan, gp43
Galectin-3C. albicansβ-1,2-mannosides
Scarf1/CDC36C. albicans, C. neoformansβ (1,3)- glucan
TLR
TLR2
C. albicans, Alternaria, A. fumigatusPhospholipomannans, α-glucans zymosan
TRL3A. fumigatusdsRNA
TLR4C. albicans, A. fumigatusα-, β-glucan, and galactomannan N
TLR6Candidaspp. Zymosan, unmethylated DNA with CpG motif
TLR7Candidaspp. ssRNA
TLR9C. neoformans, C. albicans, A. fumigatusZymosan, CpG-oligodeoxynucleotides
NLR
NLRP3
C. albicans, A. fumigatus, C. neoformans, Malasseziaspp., P. brasiliensis, S. schenckii, H. capsulatumUnknown
NLRP4C. albicansUnknown
NLRP10C. albicansUnknown
NOD1A. fumigatusUnknown
NOD2C. parapsilosis, A. fumigatusChitin
RLR MDA5C. albicansUnknown
Table 4. The list of some antifungal vaccines (Nami et al. (2019) with some modifications) [92].
Table 4. The list of some antifungal vaccines (Nami et al. (2019) with some modifications) [92].
PathogenAntigenVaccine TypeUnderlying Immune MechanismReference
CandidiasisAls3p
Als1p
Recombinant protein (NDV-3)IgG, IL17 A, IFN-γ, and IgA1,[93,94,95,96,97,98]
SAP2Recombinant proteinProtective antibodies[95,99]
aspartyl proteinase protein, Sap2p (Secreted fraction)RecombinantAntibodies[99]
Tet-NRG1Recombinant and live attenuatedElicit immunity mediated by T cells [100,101]
PCA-2 strain of Candida albicansLive-attenuated type Elevated activity of macrophages and polymorphonuclear leukocyte[102]
Surface proteins of the cell wallProtein subunit vaccine typeImmunoglobulins- and Th17 cytokine-mediated response [103]
Mannose derivatives of Candida albicansProtein conjugate vaccine with mannan derivativesImmunoglobulin responses[104]
Laminarin (Lam) β-polysaccharidesLam diphtheria toxoid CRM197 conjugatePassive immunoglobulin responses[82,83,104,105,106,107]
C. dubliniensis mannan and human serum albuminConjugate vaccine type Immunoglobulin (IgG, IgA)-mediated and Th1-mediated responses[108]
Fructose bisphosphate aldolase (Fba) (cytosolic and cell wall peptidesSubunit vaccine typeImmunoglobulin response [109,110]
Heat-killed C. albicans (HK-CA)Recombinant vaccine typeAntibody (IgG, IgA)- Th1[111]
Cell surface protein named Hyr 1 from C. albicans. Recombinant vaccine type (using N-terminal part) Immunoglobulin response[112]
Combination of peptide epitopes and β-mannan conjugatesSubunit/conjugate vaccine typeImmunoglobulin-mediated and Th1- mediated responses[113]
AspergillosisCrude culture of Aspergillus fumigatus filtrate Sonicated and filtered fractions subunit-based vaccineTh1 cells-mediated IFN-γ and IL-2 cytokine-induced response [114]
Asp f3Incomplete Freund’s adjuvantImmunoglobulin- and T cell-mediated response[115]
Viable conidia of Aspergillus fumigatusSonicated and filtered fractions subunit-based vaccineNot yet Described [116]
Aspergillus fumigatus hyphal sonicate (HS)RecombinantImmunoglobulin-mediated and T cell-mediated responses[117]
Heat-killed Saccharomyces sp.Live-attenuated vaccine types Mostly cell-mediated immune response mediated by Th1, Th2, and Th17 cytokines [117,118]
Cell wall epitope (p41 of gluconate, Crfl1) from Aspergillus fumigatusSubunitMHC 2-mediated T cell-based protection and protection against lethal infection with C. albicans in cross-section[119]
Asp 16 fRecombinant vaccine type Th1[120]
Asp 3 fRecombinant vaccine type Th1[117]
Various proteins Gel1p,
Mep1p, Crf1p, Sod1p, Dpp5p, RNUp,
Pep1p, Cat1p, Polysaccharides: β1–3 glucan, β1–3 glucan, GM, glycolipids: GSL, LGM
Recombinant/Subunit type vaccineTh1[98]
BlastomycosisAdhesin BAD1 geneWhole organism/live attenuated vaccine Cytotoxic T cell-mediated immunity [121]
Paracoccidioidomycosis (PCM)gp 43 (P10)DNA vaccine (pcDNA3-P10)T regulatory cells-based immunological memory[122]
gp 43 (P10)Recombinant protein vaccine Th1 cell immunity with concurrent role of IL-12 and IFN-γ cytokines[123]
P10- FliC fusion proteinRecombinantTh1[124]
rPb27RecombinantImmunoglobulin-based humoral immunity [125]
Heat shock protein (Hsp 60)RecombinantHelper cell-mediated immunity [126]
Mycobacterium leprae derived HSP65Recombinant DNATh1[127]
CryptococcosisGXMConjugate/soluble antigenic fractionsAnti-GMX antibodies (active immunization)[128]
GalXMSubunit/conjugate Vaccine Immunoglobulin (IgG, IgA)-mediated immunity [129]
Serotype A and Matα of C. neoformans strain H99γ Live attenuatedTh- and cytotoxic T cell-mediated immunity[130]
Mutant C. neoformans strain lacking the enzyme steryl glucosidase 1 named (Δsgl1)Live attenuated recombinantTh- and cytotoxic T cell-mediated immunity[131]
CneF (culture filtrate Ags), mannoproteinSubunit/recombinantImmunoglobulin (IgG, IgA)-mediated and Th1-mediated responses[132]
GXMGXM–protein conjugateHigh-titer IgG responses[133]
LaminaranSubunit (algal βglucan based)Passive immunity[134]
PneumocystisKexin genesKexin-CD40 L DNA vaccineElevated IgG titers[135]
P55 proteinRecombinant proteinTh1-Th2 responses[136]
gp120Recombinant proteinImmunoglobulin (IgG, IgA)-mediated and Th1-mediated responses[137]
HistoplasmosisEthylenediamine extract from cell wall in aqueous formulationInactivated filtrated antigen combined with soluble antigenic fractionsNd[138]
Ribosomes or live yeast cells of H. capsulatumLive attenuated vaccine typeLymphoid cell-mediated immunity[139]
Histone H2B–like proteinLive attenuated and recombinant type Immunoglobulin (IgG, IgA)-mediated and Th1-mediated responses[140]
HSP-60Recombinant type vaccineTh1[141]
HIS-62Recombinant protein vaccine Cellular immune response[142]
80 kDa antigenRecombinant vaccineImmunoglobulin (IgG, IgA)-mediated and Th1-mediated responses[143]
Sec31 homologueRecombinantT cell-mediated[144]
H antigenRecombinant antigenTh1.Th2/CD8+[145]
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

Rabaan, A.A.; Alfaraj, A.H.; Alshengeti, A.; Alawfi, A.; Alwarthan, S.; Alhajri, M.; Al-Najjar, A.H.; Al Fares, M.A.; Najim, M.A.; Almuthree, S.A.; et al. Antibodies to Combat Fungal Infections: Development Strategies and Progress. Microorganisms 2023, 11, 671. https://doi.org/10.3390/microorganisms11030671

AMA Style

Rabaan AA, Alfaraj AH, Alshengeti A, Alawfi A, Alwarthan S, Alhajri M, Al-Najjar AH, Al Fares MA, Najim MA, Almuthree SA, et al. Antibodies to Combat Fungal Infections: Development Strategies and Progress. Microorganisms. 2023; 11(3):671. https://doi.org/10.3390/microorganisms11030671

Chicago/Turabian Style

Rabaan, Ali A., Amal H. Alfaraj, Amer Alshengeti, Abdulsalam Alawfi, Sara Alwarthan, Mashael Alhajri, Amal H. Al-Najjar, Mona A. Al Fares, Mustafa A. Najim, Souad A. Almuthree, and et al. 2023. "Antibodies to Combat Fungal Infections: Development Strategies and Progress" Microorganisms 11, no. 3: 671. https://doi.org/10.3390/microorganisms11030671

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