Open access peer-reviewed chapter

Fungal Immunology: Mechanisms of Host Innate Immune Recognition and Evasion by Pathogenic Fungi

Written By

Faisal Rasheed Anjum, Sidra Anam, Muhammad Luqman, Ameena A. AL-surhanee, Abdullah F. Shater, Muhammad Wasim Usmani, Sajjad ur Rahman, Muhammad Sohail Sajid, Farzana Rizvi and Muhammad Zulqarnain Shakir

Submitted: 06 October 2021 Reviewed: 27 October 2021 Published: 31 January 2022

DOI: 10.5772/intechopen.101415

From the Edited Volume

Fungal Reproduction and Growth

Edited by Sadia Sultan and Gurmeet Kaur Surindar Singh

Chapter metrics overview

381 Chapter Downloads

View Full Metrics

Abstract

For a fungal pathogen to successfully infect, colonize and spread inside a susceptible host, it must have overcome the host immune responses. The early recognition of the fungal pathogen-associated molecular patterns (PAMPS) by the host’s pattern recognition receptors (PRRs) results in the establishment of anti-fungal immunity. Although, our immune system has evolved several processes to combat these pathogens both at the innate and adaptive immune levels. These organisms have developed various escape strategies to evade the recognition by the host\'s innate immune components and thus interfering with host immune mechanisms. In this chapter, we will summarize the major PRRs involved in sensing fungal PAMPS and most importantly the fungal tactics to escape the host\'s innate immune surveillance and protective mechanisms.

Keywords

  • PAMPs
  • PRRs
  • innate immunity
  • escape mechanisms
  • pathogenic fungi

1. Introduction

Pathogenic fungi are an important cause of morbidity and mortality in humans particularly in immune-compromised individuals [1, 2]. The most common risk factors for the increased incidence of fungal infections in immunocompromised individuals are cancer therapy, use of corticosteroids and neutropenia [3, 4, 5]. Sporadic occurrence of fungal infections has also been described in immunocompetent individuals that have undergone any traumatic inoculation such as the use of catheters or surgeries [6, 7]. Fungal pathogens show a considerable variation in their biology and disease pathogenesis and may include opportunistic fungi, i.e., Aspergillus fumigatus (A. fumigatus), and Fusariumspp. as well as some commensals such as Candida albicans (C. albicans).

The human innate immune system is the first line of defense and plays a pivotal role in the body’s defense on confrontation to the invading pathogens. One of the fundamental responses towards the infectious agents including fungi is the inflammatory response that is launched immediately by the host body following an immunological insult. This inflammatory response drives the antigens specific adaptive immune response such as activation of antigen-specific lymphocytes against the invading pathogens. The innate immune system recognizes a particular set of conserved surface molecules exhibited by the pathogens called pathogens-associated molecular patterns (PAMPs). Host cell pattern recognition receptors (PRRs) detect microbial PAMPs and trigger the intracellular signaling pathways that lead to the production of cytokines, reactive oxygen species (ROS), reactive nitrogen species (RNS), and lipid mediators [8, 9]. Toll-like receptors (TLRs) and C-type lectin receptors (CLRs) are the most common PRRs characterized for the detection of fungal PAMPs. Microbial detection by these PRRs results in a cascade of signaling events that eventually result in the production of inflammatory mediators, phagocytosis and induction of adaptive immune response [10]. However, fungal pathogens have adapted simple yet innovative strategies to evade and/or counteract the host innate immune responses thus resulting in the establishment of a successful infection inside the host. In the current chapter, we have made a comprehensive understanding of the major innate immune receptors involved in the detection of the fungal pathogens as well as the strategies employed by the pathogenic fungi to evade and therefore enhance their viability inside the host during infection.

Advertisement

2. Innate immune recognition of fungal PAMPs by host PRRs

2.1 Role of TLRs in recognition of pathogenic fungi

TLRs are a family of receptors that share structural homology with the Toll receptor (first described in the Drosophila). To date, 13 types of human and 10 types of murine TLRs have been discovered [11, 12, 13]. Generally, TLRs are comprised of extracellular and intracellular domains. The extracellular domain is rich in leucine repeats whereas the intracellular domain shares homology with the Toll/IL-1 receptor (TIR) domain. The TIR domain recruits the adapter proteins such as MyD88, TRIF, TRAM, and TIRAP followed by the initiation of intracellular signaling pathways which eventually result in the activation of different transcription factors, i.e., NF-κB, AP-1, IRFs (IRF3/7), and MAP kinases. These transcription factors lead to the expression of cytokines and co-stimulatory molecules [11].

Both TLR2 and TLR4 are involved in the innate immune recognition of fungal PAMPs (Table 1) (Figure 1) [10]. TLR4 has been described to recognize the fungal-derived mannans. Recognition of Saccharomyces cerevisiae (S. cerevisiae) and C. albicans derived mannans by human monocytes have been attributed to a mechanism dependent on TLR4 and CD14 with Lipopolysaccharide Binding Protein (LBP) amplifying this mechanism [14]. Further investigation in this regard reveals that recognition of mannans is brought by the cooperation of TLR4 with Mannose Receptor (MR) with TLR4 recognizing the O-linked mannans and MR recognizing the N-linked mannans [15]. TRL4 is required for the innate immune recognition of rhamnomannans isolated from P. boydii. Rhamnomannans trigger cytokine production from macrophages via TLR4 activation [16]. Similarly, TLR4 also detects glucuronoxylomannans (GXM) from Cryptococcus neoformans (C. neoformans) suggesting the vital role of TLR4 in innate immune recognition of mannose-containing polysaccharides [17].

PRRsFungal pathogenFungal PAMPs
1. TLRs
TLR4O-linked mannansC. albicans
MannansSacchromycesspp.
RhamnomannansP. boydii
PhospholipomannansC. albicans
GlucuronuxylomannansC. neoformans
Unknown/α-, β-glucan, galactomannanA. fumigatus
UnknownP. brasilensis
TLR2/TLR6PhospholipomannansC. albicans
GlucuronuxylomannansC. neoformans
TLR2/TLR1GlucuronuxylomannansC. neoformans
TLR2Α 1,4 glucansP. boydii
UnknownA. fumigatus (conidia and hyphae form
2. CLRs
MRN-linked mannansC. albicans
MannansP. Carinii
MannoproteinsC. neoformans
A. fumigatus
gp43P. brasilensis
Dectin-2α-mannansC. albicans
A. fumigatus
Dectin-1Β (1, 3)- glucansC. albicans
A. fumigatus
Sacchromycesspp.
DG-SIGNGalactomannansA. fumigatus
MannansC. albicans
Unknown/surface carbohydrates in extracellular vesiclesP. brasilensis
MinclePolysaccharides containing α-mannosyl residuesMalasseziaspp.
C. albican
CD14MannansSacchromycesspp.
UnknownA. fumigatus
Α (1,4) glucansP. boydii
3. NNLRs
NLRP3Unknown/ β (1,3)-glucansC. albicans
A. fumigatus
C. neoformans
P. brasilensis

Table 1.

List of major PRRs involved in recognition of various fungal-derived PAMPs.

Figure 1.

Major receptors involved in the innate immune recognition of fungal polysaccharides and glycoconjugates (Barreto-Bergter and Figueiredo, 2014).

Alike TLR4, TLR2 is also involved in the recognition of the fungal molecules. TLR2 triggers the activation of NF-κB and subsequent release of cytokines from the macrophages in response to phospholipomannan (a cell wall lipoglycan isolated from C. albicans). On the other hand, both TLR4 and TLR6 respond partially to phospholipomannan [18]. Moreover, TLR2 is responsible for the detection of glucogen (i.e., α-1,6-branched α-1,4-glucans) [19]. TLR2/TLR1 and TLR2/TLR6 heterodimers are described to be important receptors in the detection of GXM isolated from the capsules of C. gatii and C. neoformans [20].

It has been observed that cytokine production from macrophages and dendritic cells is mediated by TLR2 and CD14 in response to P. boydii- derived α-glucans [21]. Polysaccharides extracted from medicinal fungi (Ganoderma lucidum and Cordyceps sinensis) possess immune-modulatory and anticancer activities. These polysaccharides also trigger cytokine production and B cells activation through TLR2 and TLR4. Although, direct binding of fungal polysaccharides with the TLR2 and TLRs has been described [22]. The exact underlying mechanisms by which TLR2 and TLR4 interact and recognize fungal polysaccharides and other glycoconjugates containing mannose are still poorly defined. Fungal polysaccharides such as α-glucans and mannans are structurally different from the TLR2- and TLR4-prototypical agonists. There could be a possibility that complex fungal polysaccharides or cell wall components may present some uncharacterized glycolipids anchored to their structures that could serve as true TLR2- and/or TLR4- agonists. However, to find these answers, sensitive analytical techniques such as mass spectromery and nuclear magnetic resonance are required in combination with other complementary approaches, i.e., selective inhibition of investigating molecules, use of chemically defined ligands, availability of genetic models deficient in synthesizing the specific fungal molecules. Both TLR4 and TLR2 interact directly with bacterial lipid a and lipopeptides, respectively. Moreover, crystallographic studies of lipid A-TLR4 complex and lipopeptide-TLR1/TLR2 or lipopeptide-TLR6/TLR2 complex have demonstrated a physical interaction between these ligands and their respective receptors [23, 24, 25]. It was observed that fatty acid chains present in the bacterial ligands interact and bind with the hydrophobic pockets in the extracellular domains of the receptors complex. This suggests that TLR4 and TLR2 interact with hydrophilic ligands such as fungal polysaccharides in a way that is distinct from that of classical bacterial ligands. Thus, knowing the structural basis for interactions between fungal polysaccharides and TLR2 or TLR4 will be very helpful in understanding how distinct structures, i.e., LPS (lipopolysaccharide), lipopeptides and mammalian endogenous molecules (hyaluronic acid, carboxy alkyl pyrroles, heme) are recognized by these receptors [26].

2.2 Role of CLRs in recognition of pathogenic fungi

CLRs are a group of proteins receptors involved in the detection of fungal glycoconjugates and are characterized by the presence of two motifs; the EPN motif and QPD motifs both of which drive the specificity of CTLRs towards carbohydrate moieties. The EPN motif helps in the recognition of mannose, N-acetylglucosamine, glucose and L-fucose whereas the QPD motif is involved in the recognition of N-acetylgalactosamine and galactose [27, 28, 29, 30, 31]. The major CTLRs implicated in the recognition of fungal molecules are Dectin-1, Dectin-2, Mannose receptors (MR), Mincle, DC-SIGN, CD-12, and CD-11b/CD18 (Table 1) (Figure 1).

2.2.1 Mannose receptor (MR)

The mannose receptor is a type I transmembrane receptor that contains an N-terminal cysteine-rich domain and a type II fibronectin domain. The extracellular component of MR comprises of eight CTLDs (C-type lectin domains) whereas the intracellular component possesses a motif required for endocytic signaling [32]. MR is capable of recognizing fungal PAMPs that contain either mannose, glucose, N-acetylglucosamine [32, 33, 34], or sulfated galactose and sulfated N-acetylglucosamine (Table 1) (Figure 1). Recognition of sulfated glycoconjugates is mediated by the cysteine-rich domain of MR and is independent of CTLDs [35, 36]. On the other hand, recognition of mannose-based fungal PAMPs is dependent on the activity of CTLDs (mostly 4–8) [34]. Several fungal pathogens can be detected by MR including C. albicans and Pneumocystis carinii [15, 37, 38]. MR is considered a phagocytic receptor due to its involvement in the phagocytosis of pathogenic fungi (Table 1) [37, 38]. However, the expression of MR on non-phagocytic cells questions its designation as a professional phagocytic receptor [39]. The role of MR in the recognition and phagocytosis of pathogenic fungi containing mannose ligands is well established [37, 38]. There is a possibility that role of MR as a phagocytic receptor might be cell type-specific.

Besides its role in endocytosis, MR also contributes to the production of cytokines in response to glycoconjugates comprising mannans [15, 40, 41]. On recognizing the C. albicans derived mannans, MR induces the release of TNF-α from macrophages [15]. MR is also considered to be involved in sensing of mannosylated glycoconjugates such as those derived from mycobacteria and Pichia pastoris indicating that in addition to C. albicans derived mannans, MR can also recognize mannosylated glycoconjugates presented from other pathogenic fungi [42, 43]. The mechanism of cytokine induction by MR in response to fungal glycoconjugates is still undefined. MR also contains a short cytoplasmic tail that lacks motifs involved in intracellular signaling and eventually cytokine production. There could be a possibility that MR confers cytokine production in association with other CLRs or TLRs.

2.2.2 Dectin-2

A transmembrane protein that was first characterized in a cell line derived from Langerhans cells. Dectin-2 is comprised of a short cytoplasmic domain and an extracellular domain with CLTD present in the COOH-terminal region. Activation of Dectin-2 by the fungal ligands leads to the production of eicosanoids and cytokines [44, 45, 46]. Usually, macrophages, some dendritic cells and IL-6/IL-23-stimulated neutrophils express Dectin-2 receptors [45, 47, 48]. The EPN motif present in the extracellular domain of Dectin-2 recognizes fungal glycoconjugates containing mannose and fucose (Table 1) (Figure 1) [47]. The binding of Dectin-2 to zymosan requires Ca2+, however, higher concentrations of mannose, fucose, glucose, galactose and N-acetylglucosamine can inhibit this binding. Dectin-2 shows a higher binding affinity towards synthetic carbohydrates that are extensively mannosylated. However, the binding affinity decreases with the decrease in mannosylated residues [49]. Dectin-2 receptor recognizes C. albicans by binding to its α-mannans [46, 50]. In case of Malasseziaspp., this recognition is mediated through the binding of Dectin-2 with a glycoprotein containing O-linked α-1,2-mannobiose [51]. Besides this, in cooperation with MCL, Dectin-2 also has been described to promote recognition of C. albicans hyphae through binding to α-mannans followed by the formation of heterodimer complex. This cooperation between Dectin-2 and MCL results in higher sensitivity in recognizing the C. albicans and eventually leads to amplified leukocyte responses towards the fungal pathogens [50].

2.2.3 Dectin-1

Dectin-1 is a type II transmembrane receptor that recognizes and binds to the molecules containing β (1,3)-glucans. It is expressed by many cell types including macrophages, dendritic cells, eosinophils and neutrophils [52, 53, 54]. Dectin-1 mediated signaling results in the production of cytokines [55], maturation of dendritic cells [56] and production of ROS [57]. This suggests that Dectin-1 is an important PRR in recognizing β (1,3)-glucans followed by leukocytes activation and induction of adaptive immunity. However, in contrast to many other CTLRs, Dectin-1 binding to β-glucans is not dependent on Ca2+ [58, 59]. Dectin-1 possesses higher specificity towards β (1,3)-glucans having β (1,6)-branches. On the other hand, Dectin-1 is unable to bind with mannans, pullulans, β1,6-glucans or β (1,3)/(β1,4)-glucans [58]. In contrast to TLRs which recognize soluble ligands, activation of Dectin-1 is dependent upon its clustering by the β-glucans molecules followed by exclusion of tyrosine phosphatases (CD45 and CD48) and phosphorylation of hemi-ITAM motif present in the cytoplasmic tail of Dectin-1 [60, 61]. The hemi-ITAM recruits Syk kinases and initiates the upstream signaling pathway leading to the activation of NF-κB and NFAT [60, 62]. Usually, alveolar macrophages (AMs), resident peritoneal macrophages and dendritic cells present Dectin-1 dependent responses towards β-glucans while bone marrow-derived macrophages do not. However, Dectin-1 dependent responses can be promoted in non-responding cells such as bone marrow-derived macrophages in the presence of IFN-γ and GM-CSF thus suggesting that responses mediated by Dectin-1 are flexible [63, 64].

2.2.4 Mincle

Also known as CLEC4E is a type II transmembrane protein that was first identified as a macrophage-expressed gene dependent on the activity of the NF-IL6 transcription factor [65]. It is composed of a short cytoplasmic tail with an extracellular domain containing a CLTD. Mincle triggers cell signaling by recruiting the FcRγ chain which leads to the activation of NFAT and NF-κB and eventually induces transcription of cytokines [66]. Like other CTLRs, Mincle also plays an important role in the recognition of fungal molecules (Table 1) (Figure 1) [67, 68]. Soluble Mincle has been described to interact with C. albicans [68]. Moreover, in response to C. albicans infection, Mincle is required for TNF-α production from the macrophages [67]. Mincle also detects Malasseziaspp. (human commensal fungi) and activates NFAT mediated cytokine transcription in cell lines. However, an impaired cytokine production and leukocyte recruitment was observed in Clec4e-/- macrophages in response to Malasseziaspp. [69]. Currently, two glycolipids ligands have been identified from C. albicans that bind with Mincle. One is a polar glycolipid (comprising one dimannosyl-10-hydroxy-octadecanoic acid and two mannosyl-10-hydroxy-octadecanoic acids). The second ligand is a glyceroglycolipid containing the disaccharide gentiobiose joined to a glycerol backbone acylated with C14 and C18 fatty acids [51]. Although, Mincle ligands for other species of fungi have not been identified yet. However, it seems that Mincle must be specific in binding and recognizing glycolipid molecules among other pathogenic fungi [67, 68].

2.2.5 DC-sign

A type II transmembrane receptor with extracellular domain containing one CRD in its COOH- terminal and an and extracellular stalk. The stalk is comprised of seven residues that aid in DC-SIGN oligomerization [70, 71]. The CRD of DC-SIGN contains an EPF motif. DC-SIGN binds to glycoconjugates containing mannans and fucosylated carbohydrates in the presence of Ca2+ [72]. Both macrophages and dendritic cells express DC-SIGN. It is an endocytic receptor that can recognize and internalize several fungal pathogens followed by their release in the endosomal vesicles [73, 74]. In response to mannose-containing ligands, DC-SIGN has been described to enhance the cytokine production induced by the TLRs whereas fucosylated ligands amplify the IL-10 while inhibiting the production of proinflammatory cytokines. Mannosylated lipoarabinomannans (ManLAM) mediated activation of DC-SIGN results in inhibition of dendritic cells maturation by the LPS [75]. Thus, some pathogens infecting the dendritic cells can escape the immune activity by activating and inhibiting the DC-SIGN mediated maturation of dendritic cells. Thus, activation of DC-SIGN seems to exhibit complex effects such as internalizing the pathogens, triggering cytokine production and restricting the maturation of dendritic cells [76].

DC-SIGN has been involved in the recognition of many fungal pathogens including C. albicans (Table 1) (Figure 1). DC-SIGN is involved in the internalization and delivery of C. albicans to the phagolysosome of dendritic cells [77]. The binding and internalization of C. albicans is associated with the presence of N-linked mannans as a decreased binding was observed in C. albicans strains deficient in N-linked mannansylation [38]. However, dendritic cells do not exhibit any decreased binding affinity to the C. albicans strains lacking N-linked mannansylation. It appears that the binding of C. albicans through DC-SIGN in dendritic cells requires N-linked mannans whereas other glycoconjugates such as phosphomannans, O-linked mannans, or terminal β (1,2) mannosides are dispensable. There is no doubt regarding the role of DC-SIGN in recognition of C. alibican but it’s the MR that in cooperation with DC-SIGN, contributes majorly to C. albicans internalization by the dendritic cells [38]. DC-SIGN is also an important PRR for recognition of A. fumigatus conidia by macrophages and dendritic cells. Unlike the C. albicans, MR is not required for the recognition of A. fumigatus by the human dendritic cells, however, this recognition by DC-SIGN has been described to be inhibited by the purified mannans and galacto-mannans [78]. Soluble DC-SIGN detects the glycoconjugates such as mannans, monosacchrides comprising mannans and Lewis antigen structures in a Ca2+ dependent manner [72, 74, 79]. Although, DC-SIGN is involved in the recognition of fungal pathogens, the underlying phenomena of modulation of macrophage and dendritic cell-mediated immune responses by DC-SIGN in response to fungi are still undefined.

2.3 CD11b/CD18 (MAC-1, CR3)

CD11b/CD18 also recognized as CD18 is a heterodimer receptor comprising of type I protein chains; αM chain (CD11b) and the common chain CD18 both of which are attached non-covalently. CD11b/CD18 is expressed by many of the leukocytes such as neutrophils, eosinophils, monocytes, macrophages and NK cells [80]. CD11b/CD18 helps in the adhesion of leukocytes to the activated endothelium and phagocytic receptors for antigens opsonized with iC3b [81]. In addition, CD11b/CD18 is also involved in the detection of β (1,3)-glucans. The αM chain of CD11b/CD18 possesses two distinct domains; the I-domain and a lectin domain. The I-domain binds ICAM-1, iC3b and fibrinogen whereas the lectin domain recognizes the fungal glycoconjugates such as β (1,3)-glucans, glucose, mannose and N-acetyl-D-glucosamine [82]. CD11b/CD18 triggers ROS production from neutrophils and macrophages in response to S. cerevisiae and zymosans [83]. Although, CD11b/CD18 is actively involved in the recognition of β (1,3)-glucans, however, some controversy exists in some experimental settings regarding its role in the identification of β (1,3)-glucans along with Dectin-1 mediated responses. The differences in experimental settings could be attributed to the observed disparities in results and can be attributed to many factors such as the use of distinct ligands (i.e., soluble vs. particulate β-glucan structures) [63, 84, 85], heterogeneity of β-glucan structures (both zymosan and fungi are heterogeneous and also contains carbohydrates and lipids in addition to mannans) [83], presence of the serum [85] and variability in the cell populations (neutrophils vs. macrophages) [84] used in the experiments. In conclusion, we can say that both CD11b/CD18 and Dectin-1 are involved in the recognition of β-glucan structures and their activation must lead to the induction of immune responses towards the fungal pathogens. Besides recognizing β-glucans, CD11b/CD18 also acts as an internalization receptor for mycobacterial PIM2 (a mycobacterial glycoconjugate-coated beads) [86] suggesting that CD11b/CD18 also work as a receptor for other fungal molecules other than β-glucan.

2.4 CD14

A glycosylphosphatidylinositol-anchored protein receptor was initially considered as an LPS binder. The Cd14 receptor is comprised of an extracellular domain containing cysteine-rich residues that form a horseshoe-like conformation [87, 88, 89]. Although, the CD14 receptor does not contain intracellular regions, however in cooperation with TLR2/MD receptors, it confers a high degree of sensitivity towards LPS [89]. CD14 has also been recognized as a co-receptor involved in TLR2- [90], TLR3- [91], TLR7- and TLR9-mediated detection of ligands [92]. Similar to the LPS, detection of mannans derived from C. albicans and S. cerevisiae also depends on CD14, LBP and TLR4 [14]. CD14 can also detect other fungal glycoconjugates, i.e., β (1,3)-glucans [21] and carbohydrate and therefore, act as an important receptor for innate immune recognition of P. boydii [93] and A. fumiatus [93]. However, the structural basis for the recognition of carbohydrate ligands by CD14 is still undefined. Also, the direct binding of CD14 with mannans and α-glucans has not been elucidated. There is a possibility that CD14 must be binding to these ligands via hydrophilic cleft. As CD14 also acts as a receptor for TLR ligands, we can speculate that CD14 must be promoting intracellular signaling first by binding to these carbohydrate ligands followed by their loading onto TLR2 or TLR4.

Advertisement

3. Fungal strategies of host innate immune evasion

3.1 Shielding of stimulatory PAMPs

Protecting the pathogen’s inflammatory PAMPs from recognition by the host’s PRRs is one of the most significant escape mechanisms employed by the microbes [94, 95]. PRRs, which are found in various cellular components of primitive immune cells, are capable to identify recurrent pathogenic structures called PAMPs [96]. The host usually responds via phagocytic processes to establish the immediate antifungal mechanisms in response to fungal PAMPs. It is also accompanied by antimicrobial and pro-inflammatory responses launched through the activation of various intracellular signaling pathways that lead to the cytokine’s and chemokine’s gene transcription [97]. The prime objective of this response is to limit the disease while capturing as well as presenting the antigen to activate the adaptive immunity [98, 99]. NOD-like receptors (NLRs), TLRs, CTLRs and RIG-I-like, are the four groups of PRRs that vary in regards to ligand identification, signal transduction, as well as subcellular localization. Dendritic cells (DCs) and other myeloid cells exhibit the majority of PRRs, which are known for activating innate immune responses. PRR signaling, on the other hand, may regulate the progression of innate immune responses by the secretion of cytokines that helps in the polarization of CD4+ cells [100]. CLRs are the main class of receptors that identify fungus, according to multiple investigations, whereas NRLs and TRLs play leading functions. Microbes may hide such that they are often overlooked by the immune system [101]. Polysaccharides as well as many other components of the cell wall are often layered and serve physiological and architectural roles in the cell wall. The structure of the cell wall layers of fungus is critical in serological identification [102]. Cell wall elements (i.e., chitin, mannan, and glucans) are also included among the fungal PAMPs. Most fungi contain chitin and also α (1,3)-glucan-based internal skeletal layer of the cell wall, which is linked to certain cell wall glycoproteins and polysaccharides [102]. Many fungal species may alter glucans and chitin to decrease host recognition, thus avoiding immune activation. Antagonism or synergism of receptor activation may result in a variety of diverse pathways of inflammatory processes. In vivo, a variety of fungal ligands have been exhibited in varying proportions, resulting in the activation of various PRRs [103]. Dectin-1 has a specific key for detecting hyphal infections via the identification of β-glucans, a fundamental component of the hyphal cell wall [94, 95]. C. albicans (a polymorphic fungus) may exhibit a transition between yeast and filamentous types, depending on environmental conditions. The bud scars on the Candida wall, which are exposed during budding, reveal the usually hidden β-glucan, and are predisposed to Dectin-1 identification. Usually, β-glucans of C. albicans are hidden from identification via Dectin-1 and outer wall elements during hyphal development [104, 105]. Dectin-1 also enhances the fungicidal activities of human neutrophils, which seem to be key effector cells throughout the fight against hyphal morphology [106]. The failure of Dectin-1 to identify the fungi as a result of β-glucan protection in hyphal development shields these bigger morphologies by avoiding internalization. In addition, hyphal types elicit protective T-helper cell type 2 (Th2) immunological responses in DCs rather than a Th1 immune response [107, 108]. Consequently, immune cells respond differentially to hyphae and yeast.

Similar to Dectin-1, Dectin-2 and Dectin-3 are also integral membrane proteins that belong to the CLRs family. Dectin-1 detects glucans, while Dectin-2/3 identifies mannans [109]. These may produce heterodimer complexes that provide greater sensitivity to host tissues as well as a high potential for binding to mannans [50]. While investigating the functions of Dectin-2 in C. albicans related-diseases, it was observed that the mice lacking Dectin-2 were more vulnerable to infection. In addition, phagocytosis and cytokine production was also decreased in these mice [110].

The α (1,3)-glucans present in the outermost layer of the cell wall helps in the pathogenicity of Histoplasma capsulatum (H. capsulatum) by hiding its immune-stimulatory β-glucans. This is analogous to how external mannans protect β-glucan from Dectin-1 recognition in Candida. Further evidence was observed in H. Capsulatum variants without α (1,3)-glucans which resulted in an increased TNF-α production. On the other hand, a reduction in Dectin-1 (necessary for β-glucans) expression, suppresses TNF-α levels. On switching into its infectious yeast stage, Paracoccidioides brasiliensis (P. brasiliensis) changes its β(1, 3)-glucans to α(1, 3)-glucans [111]. This is due to the reason that α (1, 3)-glucans are less likely to be identified by the host PRRs and therefore essential in fungal evasion of the immune system. C. neoformans masks the surface of PAMPs by producing GXM, which inhibits the production of IL-1β and pro-inflammatory TNF-α [112]. Both the pigment DHN-melanin and the RodA protein create a hydrophobic coating on A. fumigatus conidia and cover the glucans in order to avoid TLR activation. Although, quiescent conidia do not cause macrophages to secrete cytokines during germination, and the surface of RodA is destroyed. Furthermore, proteins that are recognized by PRRs are exposed to dendritic cells and macrophages, promoting the expression of the co-stimulatory molecule and cytokine production [113]. Dectin-1 redundant function may be explained by the lack of numerous glucans on the surface of resting Aspergillus conidia. Dectin-1 suppression on alveolar macrophages has little effect on the phagocytosis of Aspergillus that may be influenced by conidia germination [114, 115]. The spherule external wall glycoprotein (SOWgp) of Coccidioides posadasii (a respiratory fungal pathogen) is involved in its escape from innate immune recognition. The fungal cells secrete a metalloproteinase (Mep1) during endospore development, which metabolizes SOWgp [116]. Because SOWgp is downregulated during endospore production, fungal cells are therefore capable of avoiding phagocytosis and death during the susceptible spore-forming stage [117]. The capacity of the fungi to live in various morphotypes and to transiently shift from one form to another during infection is one strategy of shielding the host’s immune system [118, 119]. These polymorphic phases are linked to phenotypic switching and result in evasion from cellular PRRs. This ability has most likely developed to help fungi survive in a variety of environments.

The C. neoformans capsule obscures the α (1,3)-glucans and mannan of the basal cell wall. Macrophages can easily recognize and phagocytose the acapsular mutant strains of C. neoformans and both glucans and mannose receptors are involved in this identification [120]. However, the capsule acts as a shield and masks the recognition of fungal PAMPs from the phagocytic receptor. Generally, TLRs detect the capsule and initiate an inflammatory response that is essential for limiting fungal infections. Further evidence in this regard has been provided by TLR2-deficient mice that have shown increased susceptibility to C. neoformans infections [121]. Contrary to yeast-form, blastoconidia of C. albicans, as well as hyphal form, are capable of evading the innate immune recognition by the Dectin-1 [104]. These blastoconidia activate the TLR2 and TLR4 in the ancillary monocytes and peritoneal macrophages. Such hyphal forms are not detected by the TLR4 and cause tissue-invasive infection [122]. Phenotypic change during germination could be a crucial survival strategy for several fungal pathogens. TLR4-mediated responses are diminished after the germination of A. fumigatus while its conidia elicit TLR2- as well as TLR4-mediated responses, in the tissue invasion. Mostly conidia germinate specifically into hyphae when TLR4 production is reduced, resulting in less intense proinflammatory cytokine production [123]. Proinflammatory responses mediated by TLR4 are essential in the prevention of invasive infections (aspergillosis) [124]. The stealth mode is not always successful and a fungal pathogen will usually be detected by the host in a certain way. Therefore, microorganisms frequently discover new strategies to exploit the host recognition networks and manipulate them for establishing a successful infection. These organisms may have chemicals on their surfaces or release compounds that trigger regulatory systems specifically. Throughout this way, the pathogen may either directly suppress or develop kinds of immune responses that aren’t typically efficient against the pathogen [101].

3.2 Modulation of inflammatory signals

In respect of anti-inflammatory cytokine impact, TLR2 stimulation differentiates from TLR4 activation, with proinflammatory cytokine production being lower following TLR2 stimulation than the TLR4 stimulation [125]. TLR4 agonists selectively produced Th1-inducing cytokine signals in DCs, whereas TLR2 activation generate a more strong anti-inflammatory Th2 reaction [126]. Because each effector’s arm elicits a different immune reaction, the equilibrium between Th1/Th2 reactions is thought to be important in deciding the severity of infection [127, 128]. The Th1 pathway generates pro-inflammatory cytokines such as IFN-γ, which stimulate cell-mediated immune mechanisms such as cytotoxicity and phagocyte activation. Th1 pathway is essential in the fight against intracellular and fungal infections. The Th2 pathway is characterized by cytokines such as IL-4, IL-5, as well as IL-10 and promotes a humoral response while suppressing the Th1-dependent effector functions [129]. Th2 cytokines may decrease monocyte anti-hyphal activity as well as lead to oxidative burst amid antifungal reactions [128]. TLR2-deficient macrophages have improved anti-candidal abilities [130], and TLR2 macrophages in mice are significantly more tolerant to widespread C. albicans related diseases [124, 130, 131]. As a result, the hyphal forms of C. albicans (tissue-invasive) and A. fumigatus are likely to shift the equilibrium towards that Th2 pathway by avoiding TLR4 stimulation in favor of TLR2 activation. C. neoformans has also been found to possess immunosuppressive properties. The primary virulence component of C. neoformans is indeed the GXM, which is a strong activator of the IL-10 (anti-inflammatory cytokine) and a pro-Th2 cytokine mediator in human monocytes [132, 133]. The melanin pigment produced frequently by pathogenic filamentous fungi has been associated with fungal pathogenicity and its immunomodulatory impact in C. neoformans has been investigated. Melanized C. neoformans variants result in increased pulmonary IL-4 levels thus driving the host cells to switch towards Th2 response [134]. Blastomyces dermatitidis (B. dermatitidis) that causes systemic and pulmonary mycosis, may result in comparative immunosuppression by reducing the synthesis of the TNF-α (a pro-inflammatory cytokine) [135]. The binding of Blastomyces surface adhesins to the complement receptor III on macrophages results in suppression of TNF-α synthesis which otherwise could be harmful to B. dermatitidis’ existence [136].

3.3 Shedding of decoy components

Several innate immune evasion strategies have been identified for Pneumocystis jiroveci (P. jiroveci), an opportunistic fungi that usually infects AIDS patients. Glycoprotein A (gpA) complex is the main protein antigen present on the membrane of Pneumocystis. It is highly glycosylated with glucose, mannose, as well as galactose-containing carbohydrate moieties [137]. Mannose receptors present on the alveolar monocytes recognize these structures. However, Pneumocystis escapes this recognition by MR on alveolar macrophages and impedes its phagocytic activity by premature release of its gpA glycoprotein as a decoy [138]. Furthermore, Pneumocystis has also been described to deplete MR from the membrane of AMs, preventing non-opsonic absorption by MR (surface-expressed) [139].

3.4 Persistence in the intracellular environments

Several fungal pathogens have developed the potential to avoid the phagocytic activity of macrophages. For example, C. neoformans can phenotypically shift to a mucilaginous colony type generating a significantly bigger capsulated polysaccharide GXM with modified biochemical and biophysical characteristics thus limiting AM phagocytic effectiveness [140]. When certain variants fail to escape host identification, phagocytosis by macrophages does not necessarily result in death and the ending of the life cycle as some fungi may survive the harsh environment inside the phagolysosome. Some C. albicansspp. can withstand intracellular death and produce hyphal structures and thus escape the macrophages [141]. C. albicans have an extremely specialized anti-nitric oxide (NO) defense mechanism including the NO-scavenging flavohemoglobin genetic traits that convert NO to less toxic substances when comes into contact with reactive nitrogen molecules like NO and oxygen free radicals generated by monocytes/macrophages [142]. Comprehensive morphologic investigations have shown that Candida could produce germ tubes, proliferate, and ultimately escape the host cell despite phagocytosis by macrophages [143]. Phagocytosis provokes C. albicans within macrophages to switch into self-preservation mode, which includes a delayed growth rate, carbon utilization, as well as an oxidative stress reaction to thrive in the hostile environment inside macrophages [144, 145] suggesting that the phagocytic activity solely might not be sufficient to clear the infection from the host. During persistent infection, the fungi have been shown to survive and reproduce inside the phagocytic cells [146, 147]. To escape intracellular death, C. neoformans cause aberrant lysosomal transport and significant cytoplasmic vacuolation in the host cell, leading to host cell disintegration [148]. Similarly, H. capsulatum, is also capable of surviving inside macrophages for longer durations following primary infection and become activated as the immune responses are diminished [149, 150]. Histoplasma is supposed to prevent phagolysosome formation and proactively regulate the phagosomal pH following phagocytosis to maximize its survival inside phagosomes [151, 152]. Furthermore, Histoplasma can prevent the production of toxic superoxide radicals that are harmful to its survival inside macrophages [153].

3.5 Complement evasion

The complement system is a dynamic mechanism that plays a significant part in innate immunity and antibody-mediated protection against pathogenic microbes [154]. Several foreign antigens including fungal PAMPs, cellular debris, as well as antigen–antibody complexes can activate a series of complement pathways [98, 155]. Excessive tissue damage and inflammation by the complement system are avoided by the regulatory molecules of the complement system [156]. The complement system is split into three pathways; classical, alternative and lectin pathway. The activation of all these pathways varies in regards to associated components but all pathways submerge by producing the same group of effector molecules, i.e., opsonization and formation of membrane attack complex (MAC) [96]. All complement mechanisms contribute to the production of C3 convertase as well as the C3b fraction, which in turn promotes the synthesis of C5 convertase. C5 convertase cleaves the C5 factor into C5a and C5b. The distal complement components are formed as a result of a succession of accumulation and polymerization processes, as well as the mobilization of terminal complement elements such as C6, C7, C8, and C9. The terminal complement components form MAC causing cell lysis by inserting C9 into the lipid membrane layer [157, 158, 159]. Pathogenic organisms, on the other hand, have adopted different approaches to evade complement attacks, such as binding to regulatory complement proteins by secreting proteases or evading opsonization. For example, Aspergillusspp. and C. albicans release proteins on their membranes that bind to complement proteins to prevent being eliminated by the complement system. These proteins, when linked to the hyphal surface, block the complement cascade, allowing the fungi to avoid the complement attack [156]. Aspergillus and Candida are recognized to activate complement by depositing C3 on the fungal membrane, which facilitates opsonization and the synthesis of the chemoattractant (C5a), which recruits leukocytes to the infected area [160, 161, 162]. Pigmentation on the conidial surface of A. fumigatus has been demonstrated to influence pathogenicity by reducing C3 protein accumulation and neutrophil activity [163]. Transcription factors such as Factor H-like protein 1 (FHL1), Factor H, as well as C4 binding protein (C4BP) for the signaling pathway, keep the complement system in balance against abnormal activation. A. fumigatus and C. albicans both have been described to bind FHL-1, C4BP, and Factor H, on their membrane to evade the complement cascade [164, 165, 166]. Furthermore, the dense yeast cell wall is impervious to immediate lysis by the MAC [161]. Complement is far more than a “defensive” mechanism against infections. It has a role in inflammatory responses, cellular response regulation, and cell–cell interactions, all of which are important for cell differentiation and initial growth [167]. Currently, two complement-targeted drugs for non-fungal illnesses have been approved in the health center: eculizumab (an anti-C5 antibody) and different formulations of C1 esterase inhibitor (C1-INH). Several other drugs that target distinct elements of the complement cascade are all in different phases of trials [167, 168, 169, 170].

Advertisement

4. Conclusion

Our understanding regarding the innate immune recognition of pathogenic fungi by the corresponding fungal PAMPs is still poor. Moreover, the fungal ligands involved in the activation of host PRRs remain largely unknown for several pathogenic fungi. Characterization of fungal PAMPs and their recognition by the host PRRs can provide a comprehensive understanding of pathogenesis and immunity to the pathogenic fungi. In addition, characterization of these fungal ligands and their activation of respective PRRs is essential not only to discover new therapeutic approaches against fungal infections particularly in immune-compromised patients but also to develop novel adjuvants for enhancing the prophylactic immune responses against pathogenic fungi.

Advertisement

Acknowledgments

The author would like to acknowledge all the co-authors for their contribution.

Advertisement

Conflict of interest

No competing financial interest exists.

References

  1. 1. Singh N. Trends in the epidemiology of opportunistic fungal infections: predisposing factors and the impact of antimicrobial use practices. Clinical Infectious Diseases. 2001;33(10):1692-1696
  2. 2. Segal BH, Walsh TJ. Current approaches to diagnosis and treatment of invasive aspergillosis. American Journal of Respiratory and Critical Care Medicine. 2006;173(7):707-717
  3. 3. Morgan J et al. Incidence of invasive aspergillosis following hematopoietic stem cell and solid organ transplantation: interim results of a prospective multicenter surveillance program. Medical Mycology. 2005;43(Supplement_1):S49-S58
  4. 4. Chamilos G et al. Invasive fungal infections in patients with hematologic malignancies in a tertiary care cancer center: an autopsy study over a 15-year period (1989-2003). Haematologica. 2006;91(7):986-989
  5. 5. Marr KA et al. Epidemiology and outcome of mould infections in hematopoietic stem cell transplant recipients. Clinical Infectious Diseases. 2002;34(7):909-917
  6. 6. Pasqualotto A, Denning D. Post-operative aspergillosis. Clinical Microbiology and Infection. 2006;12(11):1060-1076
  7. 7. Meersseman W et al. Galactomannan in bronchoalveolar lavage fluid: a tool for diagnosing aspergillosis in intensive care unit patients. American Journal of Respiratory and Critical Care Medicine. 2008;177(1):27-34
  8. 8. Akira S, Uematsu S, Takeuchi O. Pathogen recognition and innate immunity. Cell. 2006;124(4):783-801
  9. 9. Medzhitov R. Recognition of microorganisms and activation of the immune response. Nature. 2007;449(7164):819-826
  10. 10. van de Veerdonk FL et al. Host–microbe interactions: Innate pattern recognition of fungal pathogens. Current Opinion in Microbiology. 2008;11(4):305-312
  11. 11. Takeuchi O, Akira S. Pattern recognition receptors and inflammation. Cell. 2010;140(6):805-820
  12. 12. Hidmark A, von Saint Paul A, Dalpke AH. Cutting edge: TLR13 is a receptor for bacterial RNA. The Journal of Immunology. 2012;189(6):2717-2721
  13. 13. Oldenburg M et al. TLR13 recognizes bacterial 23S rRNA devoid of erythromycin resistance–forming modification. Science. 2012;337(6098):1111-1115
  14. 14. Tada H et al. Saccharomyces cerevisiae-and Candida albicans-derived mannan induced production of tumor necrosis factor alpha by human monocytes in a CD14-and Toll-like receptor 4-dependent manner. Microbiology and Immunology. 2002;46(7):503-512
  15. 15. Netea MG et al. Immune sensing of Candida albicans requires cooperative recognition of mannans and glucans by lectin and Toll-like receptors. The Journal of Clinical Investigation. 2006;116(6):1642-1650
  16. 16. Figueiredo RT et al. TLR4 recognizes Pseudallescheria boydii conidia and purified rhamnomannans. Journal of Biological Chemistry. 2010;285(52):40714-40723
  17. 17. Shoham S et al. Toll-like receptor 4 mediates intracellular signaling without TNF-α release in response to Cryptococcus neoformans polysaccharide capsule. The Journal of Immunology. 2001;166(7):4620-4626
  18. 18. Ibata-Ombetta S et al. Candida albicans phospholipomannan promotes survival of phagocytosed yeasts through modulation of bad phosphorylation and macrophage apoptosis. Journal of Biological Chemistry. 2003;278(15):13086-13093
  19. 19. Kakutani R et al. Essential role of Toll-like receptor 2 in macrophage activation by glycogen. Glycobiology. 2012;22(1):146-159
  20. 20. Fonseca FL et al. Immunomodulatory effects of serotype B glucuronoxylomannan from Cryptococcus gattii correlate with polysaccharide diameter. Infection and Immunity. 2010;78(9):3861-3870
  21. 21. Bittencourt VCB et al. An α-glucan of Pseudallescheria boydii is involved in fungal phagocytosis and toll-like receptor activation. Journal of Biological Chemistry. 2006;281(32):22614-22623
  22. 22. Hsu T-L et al. Profiling carbohydrate-receptor interaction with recombinant innate immunity receptor-Fc fusion proteins. Journal of Biological Chemistry. 2009;284(50):34479-34489
  23. 23. Jin MS et al. Crystal structure of the TLR1-TLR2 heterodimer induced by binding of a tri-acylated lipopeptide. Cell. 2007;130(6):1071-1082
  24. 24. Kang JY et al. Recognition of lipopeptide patterns by Toll-like receptor 2-Toll-like receptor 6 heterodimer. Immunity. 2009;31(6):873-884
  25. 25. Park BS et al. The structural basis of lipopolysaccharide recognition by the TLR4–MD-2 complex. Nature. 2009;458(7242):1191-1195
  26. 26. Figueiredo RT, Carneiro LA, Bozza MT. Fungal surface and innate immune recognition of filamentous fungi. Frontiers in Microbiology. 2011;2:248
  27. 27. Drickamer K. Engineering galactose-binding activity into a C-type mannose-binding protein. Nature. 1992;360(6400):183-186
  28. 28. Kolatkar AR, Weis WI. Structural basis of galactose recognition by C-type animal lectins (∗). Journal of Biological Chemistry. 1996;271(12):6679-6685
  29. 29. Kolatkar AR et al. Mechanism of N-acetylgalactosamine binding to a C-type animal lectin carbohydrate-recognition domain. Journal of Biological Chemistry. 1998;273(31):19502-19508
  30. 30. Zelensky AN, Gready JE. The C-type lectin-like domain superfamily. The FEBS Journal. 2005;272(24):6179-6217
  31. 31. Sancho D, Sousa CR e. Signaling by myeloid C-type lectin receptors in immunity and homeostasis. Annual Review of Immunology. 2012;30:491-529
  32. 32. Martinez-Pomares L. The mannose receptor. Journal of Leukocyte Biology. 2012;92(6):1177-1186
  33. 33. Taylor ME, Drickamer K. Structural requirements for high affinity binding of complex ligands by the macrophage mannose receptor. Journal of Biological Chemistry. 1993;268(1):399-404
  34. 34. Taylor ME, Bezouska K, Drickamer K. Contribution to ligand binding by multiple carbohydrate-recognition domains in the macrophage mannose receptor. Journal of Biological Chemistry. 1992;267(3):1719-1726
  35. 35. Liu Y et al. Crystal structure of the cysteine-rich domain of mannose receptor complexed with a sulfated carbohydrate ligand. Journal of Experimental Medicine. 2000;191(7):1105-1116
  36. 36. Leteux C et al. The cysteine-rich domain of the macrophage mannose receptor is a multispecific lectin that recognizes chondroitin sulfates A and B and sulfated oligosaccharides of blood group Lewisa and Lewisx types in addition to the sulfated N-glycans of lutropin. Journal of Experimental Medicine. 2000;191(7):1117-1126
  37. 37. Ezekowitz R et al. Uptake of Pneumocystis carinii mediated by the macrophage mannose receptor. Nature. 1991;351(6322):155-158
  38. 38. Cambi A et al. Dendritic cell interaction with Candida albicans critically depends on N-linked mannan. Journal of Biological Chemistry. 2008;283(29):20590-20599
  39. 39. Le Cabec V et al. The human macrophage mannose receptor is not a professional phagocytic receptor. Journal of Leukocyte Biology. 2005;77(6):934-943
  40. 40. Tachado SD et al. Pneumocystis-mediated IL-8 release by macrophages requires coexpression of mannose receptors and TLR2. Journal of Leukocyte Biology. 2007;81(1):205-211
  41. 41. van de Veerdonk FL et al. The macrophage mannose receptor induces IL-17 in response to Candida albicans. Cell Host & Microbe. 2009;5(4):329-340
  42. 42. Kang PB et al. The human macrophage mannose receptor directs Mycobacterium tuberculosis lipoarabinomannan-mediated phagosome biogenesis. The Journal of Experimental Medicine. 2005;202(7):987-999
  43. 43. Torrelles JB, Azad AK, Schlesinger LS. Fine discrimination in the recognition of individual species of phosphatidyl-myo-inositol mannosides from Mycobacterium tuberculosis by C-type lectin pattern recognition receptors. The Journal of Immunology. 2006;177(3):1805-1816
  44. 44. Sato K et al. Dectin-2 is a pattern recognition receptor for fungi that couples with the Fc receptor γ chain to induce innate immune responses. Journal of Biological Chemistry. 2006;281(50):38854-38866
  45. 45. Barrett NA et al. Dectin-2 recognition of house dust mite triggers cysteinyl leukotriene generation by dendritic cells. The Journal of Immunology. 2009;182(2):1119-1128
  46. 46. Saijo S et al. Dectin-2 recognition of α-mannans and induction of Th17 cell differentiation is essential for host defense against Candida albicans. Immunity. 2010;32(5):681-691
  47. 47. Ariizumi K et al. Cloning of a second dendritic cell-associated C-type lectin (dectin-2) and its alternatively spliced isoforms. Journal of Biological Chemistry. 2000;275(16):11957-11963
  48. 48. Robinson MJ et al. Dectin-2 is a Syk-coupled pattern recognition receptor crucial for Th17 responses to fungal infection. Journal of Experimental Medicine. 2009;206(9):2037-2051
  49. 49. McGreal EP et al. The carbohydrate-recognition domain of Dectin-2 is a C-type lectin with specificity for high mannose. Glycobiology. 2006;16(5):422-430
  50. 50. Zhu L-L et al. C-type lectin receptors Dectin-3 and Dectin-2 form a heterodimeric pattern-recognition receptor for host defense against fungal infection. Immunity. 2013;39(2):324-334
  51. 51. Ishikawa T et al. Identification of distinct ligands for the C-type lectin receptors Mincle and Dectin-2 in the pathogenic fungus Malassezia. Cell Host & Microbe. 2013;13(4):477-488
  52. 52. Brown GD et al. Dectin-1 is a major β-glucan receptor on macrophages. The Journal of Experimental Medicine. 2002;196(3):407-412
  53. 53. Taylor PR et al. The β-glucan receptor, dectin-1, is predominantly expressed on the surface of cells of the monocyte/macrophage and neutrophil lineages. The Journal of Immunology. 2002;169(7):3876-3882
  54. 54. Willment JA et al. The human β-glucan receptor is widely expressed and functionally equivalent to murine Dectin-1 on primary cells. European Journal of Immunology. 2005;35(5):1539-1547
  55. 55. Brown GD et al. Dectin-1 mediates the biological effects of β-glucans. The Journal of Experimental Medicine. 2003;197(9):1119-1124
  56. 56. Yoshitomi H et al. A role for fungal β-glucans and their receptor Dectin-1 in the induction of autoimmune arthritis in genetically susceptible mice. Journal of Experimental Medicine. 2005;201(6):949-960
  57. 57. Gantner BN et al. Collaborative induction of inflammatory responses by dectin-1 and Toll-like receptor 2. Journal of Experimental Medicine. 2003;197(9):1107-1117
  58. 58. Adams EL et al. Differential high-affinity interaction of dectin-1 with natural or synthetic glucans is dependent upon primary structure and is influenced by polymer chain length and side-chain branching. Journal of Pharmacology and Experimental Therapeutics. 2008;325(1):115-123
  59. 59. Brown GD, Gordon S. A new receptor for β-glucans. Nature. 2001;413(6851):36-37
  60. 60. Rogers NC et al. Syk-dependent cytokine induction by Dectin-1 reveals a novel pattern recognition pathway for C type lectins. Immunity. 2005;22(4):507-517
  61. 61. Goodridge HS et al. Activation of the innate immune receptor Dectin-1 upon formation of a ‘phagocytic synapse’. Nature. 2011;472(7344):471-475
  62. 62. Goodridge HS, Simmons RM, Underhill DM. Dectin-1 stimulation by Candida albicans yeast or zymosan triggers NFAT activation in macrophages and dendritic cells. The Journal of Immunology. 2007;178(5):3107-3115
  63. 63. Rosas M et al. The induction of inflammation by dectin-1 in vivo is dependent on myeloid cell programming and the progression of phagocytosis. The Journal of Immunology. 2008;181(5):3549-3557
  64. 64. Goodridge HS et al. Differential use of CARD9 by dectin-1 in macrophages and dendritic cells. The Journal of Immunology. 2009;182(2):1146-1154
  65. 65. Matsumoto M et al. A novel LPS-inducible C-type lectin is a transcriptional target of NF-IL6 in macrophages. The Journal of Immunology. 1999;163(9):5039-5048
  66. 66. Yamasaki S et al. Mincle is an ITAM-coupled activating receptor that senses damaged cells. Nature Immunology. 2008;9(10):1179-1188
  67. 67. Wells CA et al. The macrophage-inducible C-type lectin, mincle, is an essential component of the innate immune response to Candida albicans. The Journal of Immunology. 2008;180(11):7404-7413
  68. 68. Bugarcic A et al. Human and mouse macrophage-inducible C-type lectin (Mincle) bind Candida albicans. Glycobiology. 2008;18(9):679-685
  69. 69. Yamasaki S et al. C-type lectin Mincle is an activating receptor for pathogenic fungus, Malassezia. Proceedings of the National Academy of Sciences. 2009;106(6):1897-1902
  70. 70. Geijtenbeek TB et al. Identification of DC-SIGN, a novel dendritic cell–specific ICAM-3 receptor that supports primary immune responses. Cell. 2000;100(5):575-585
  71. 71. Mitchell DA, Fadden AJ, Drickamer K. A novel mechanism of carbohydrate recognition by the C-type lectins DC-SIGN and DC-SIGNR: Subunit organization and binding to multivalent ligands. Journal of Biological Chemistry. 2001;276(31):28939-28945
  72. 72. Appelmelk BJ et al. Cutting edge: carbohydrate profiling identifies new pathogens that interact with dendritic cell-specific ICAM-3-grabbing nonintegrin on dendritic cells. The Journal of Immunology. 2003;170(4):1635-1639
  73. 73. Kwon DS et al. DC-SIGN-mediated internalization of HIV is required for trans-enhancement of T cell infection. Immunity. 2002;16(1):135-144
  74. 74. Guo Y et al. Structural basis for distinct ligand-binding and targeting properties of the receptors DC-SIGN and DC-SIGNR. Nature Structural & Molecular Biology. 2004;11(7):591-598
  75. 75. Gringhuis SI et al. Carbohydrate-specific signaling through the DC-SIGN signalosome tailors immunity to Mycobacterium tuberculosis, HIV-1 and Helicobacter pylori. Nature Immunology. 2009;10(10):1081-1088
  76. 76. van Kooyk Y, Geijtenbeek TB. DC-SIGN: Escape mechanism for pathogens. Nature Reviews Immunology. 2003;3(9):697-709
  77. 77. Cambi A et al. The C-type lectin DC-SIGN (CD209) is an antigen-uptake receptor for Candida albicans on dendritic cells. European Journal of Immunology. 2003;33(2):532-538
  78. 78. Serrano-Gómez D et al. Dendritic cell-specific intercellular adhesion molecule 3-grabbing nonintegrin mediates binding and internalization of Aspergillus fumigatus conidia by dendritic cells and macrophages. The Journal of Immunology. 2004;173(9):5635-5643
  79. 79. Van Liempt E et al. Molecular basis of the differences in binding properties of the highly related C-type lectins DC-SIGN and L-SIGN to Lewis X trisaccharide and Schistosoma mansoni egg antigens. Journal of Biological Chemistry. 2004;279(32):33161-33167
  80. 80. Ross GD. Regulation of the adhesion versus cytotoxic functions of the Mac-1/CR3/α M β 2-lntegrin glycoprotein. Critical Reviews in Immunology. 2000;20(3):1-30
  81. 81. Holers VM. Complement and its receptors: new insights into human disease. Annual Review of Immunology. 2014;32:433-459
  82. 82. Thornton BP et al. Analysis of the sugar specificity and molecular location of the beta-glucan-binding lectin site of complement receptor type 3 (CD11b/CD18). The Journal of Immunology. 1996;156(3):1235-1246
  83. 83. van Bruggen R et al. Complement receptor 3, not Dectin-1, is the major receptor on human neutrophils for β-glucan-bearing particles. Molecular Immunology. 2009;47(2-3):575-581
  84. 84. Qi C et al. Differential pathways regulating innate and adaptive antitumor immune responses by particulate and soluble yeast-derived β-glucans. Blood, The Journal of the American Society of Hematology. 2011;117(25):6825-6836
  85. 85. Bose N et al. Binding of soluble yeast β-glucan to human neutrophils and monocytes is complement-dependent. Frontiers in Immunology. 2013;4:230
  86. 86. Villeneuve C et al. Mycobacteria use their surface-exposed glycolipids to infect human macrophages through a receptor-dependent process. Journal of Lipid Research. 2005;46(3):475-483
  87. 87. Kim J-I et al. Crystal Structure of CD14 and Its Implications for Lipopolysaccharide Signaling*♦. Journal of Biological Chemistry. 2005;280(12):11347-11351
  88. 88. Kelley SL et al. The crystal structure of human soluble CD14 reveals a bent solenoid with a hydrophobic amino-terminal pocket. The Journal of Immunology. 2013;190(3):1304-1311
  89. 89. Granucci F, Zanoni I. Role of CD14 in host protection against infections and in metabolism regulation. Frontiers in Cellular and Infection Microbiology. 2013;3:32
  90. 90. Schröder NW et al. Lipopolysaccharide binding protein binds to triacylated and diacylated lipopeptides and mediates innate immune responses. The Journal of Immunology. 2004;173(4):2683-2691
  91. 91. Lee H-K et al. Double-stranded RNA-mediated TLR3 activation is enhanced by CD14. Immunity. 2006;24(2):153-163
  92. 92. Baumann CL et al. CD14 is a coreceptor of Toll-like receptors 7 and 9CD14: Coreceptor of TLR7 and TLR9. The Journal of Experimental Medicine. 2010;207(12):2689-2701
  93. 93. Wang J et al. Involvement of CD14 and toll-like receptors in signalling by Aspergillus hyphae. British Journal of Surgery. 2001;88(8):1138-1150
  94. 94. Latgé JP et al. Specific molecular features in the organization and biosynthesis of the cell wall of Aspergillus fumigatus. Medical Mycology. 2005;43(Supplement_1):S15-S22
  95. 95. Gow NA et al. Immune recognition of Candida albicans β-glucan by dectin-1. The Journal of Infectious Diseases. 2007;196(10):1565-1571
  96. 96. Janeway C et al. Basic concepts in immunology. In: Immunobiology the Immune System in Health and Disease. 5th ed. New York: Garland Publishing; 2001. pp. 1-34
  97. 97. Shizuo A, Takeda K. Toll-like receptor signaling. Nature Reviews. Immunology. 2004;4(7):499-511
  98. 98. Chai LY et al. Fungal strategies for overcoming host innate immune response. Medical Mycology. 2009;47(3):227-236
  99. 99. Bachiega TF et al. Participation of dectin-1 receptor on NETs release against Paracoccidioides brasiliensis: Role on extracellular killing. Immunobiology. 2016;221(2):228-235
  100. 100. Plato A, Hardison SE, Brown GD. Pattern recognition receptors in antifungal immunity. Seminars in Immunopathology. Springer Berlin Heidelberg. 2015;37(2):97-106
  101. 101. Underhill DM. Escape mechanisms from the immune response. In: Immunology of Fungal Infections. Dordrecht: Springer; 2007. pp. 429-442
  102. 102. Erwig LP, Gow NA. Interactions of fungal pathogens with phagocytes. Nature Reviews Microbiology. 2016;14(3):163-176
  103. 103. Levitz SM. Innate recognition of fungal cell walls. PLoS Pathogens. 2010;6(4):e1000758
  104. 104. Gantner BN, Simmons RM, Underhill DM. Dectin-1 mediates macrophage recognition of Candida albicans yeast but not filaments. The EMBO Journal. 2005;24(6):1277-1286
  105. 105. Heinsbroek SE, Brown GD, Gordon S. Dectin-1 escape by fungal dimorphism. Trends in Immunology. 2005;26(7):352-354
  106. 106. Kennedy AD et al. Dectin-1 promotes fungicidal activity of human neutrophils. European Journal of Immunology. 2007;37(2):467-478
  107. 107. Romani L, Bistoni F, Puccetti P. Fungi, dendritic cells and receptors: a host perspective of fungal virulence. Trends in Microbiology. 2002;10(11):508-514
  108. 108. d'Ostiani CF et al. Dendritic cells discriminate between yeasts and hyphae of the fungus Candida albicansImplications for initiation of T helper cell immunity in vitro and in vivo. Journal of Experimental Medicine. 2000;191(10):1661-1674
  109. 109. Saijo S, Iwakura Y. Dectin-1 and Dectin-2 in innate immunity against fungi. International Immunology. 2011;23(8):467-472
  110. 110. Ifrim DC et al. The role of dectin-2 for host defense against disseminated candidiasis. Journal of Interferon & Cytokine Research. 2016;36(4):267-276
  111. 111. Borges-Walmsley MI et al. The pathobiology of Paracoccidioides brasiliensis. Trends in Microbiology. 2002;10(2):80-87
  112. 112. Vecchiarelli A et al. Downregulation by cryptococcal polysaccharide of tumor necrosis factor alpha and interleukin-1 beta secretion from human monocytes. Infection and Immunity. 1995;63(8):2919-2923
  113. 113. Aimanianda V et al. Erratum: Surface hydrophobin prevents immune recognition of airborne fungal spores. Nature. 2010;465(7300):966-966
  114. 114. Steele C et al. The beta-glucan receptor dectin-1 recognizes specific morphologies of Aspergillus fumigatus. PLoS Pathogens. 2005;1(4):e42
  115. 115. Slesiona S et al. Persistence versus escape: Aspergillus terreus and Aspergillus fumigatus employ different strategies during interactions with macrophages. PLoS One. 2012;7(2):e31223
  116. 116. Hung C-Y et al. A metalloproteinase of Coccidioides posadasii contributes to evasion of host detection. Infection and Immunity. 2005;73(10):6689-6703
  117. 117. Hung C-Y et al. A parasitic phase-specific adhesin of Coccidioides immitis contributes to the virulence of this respiratory fungal pathogen. Infection and Immunity. 2002;70(7):3443-3456
  118. 118. Romani L. Immunity to fungal infections. Nature Reviews Immunology. 2004;4(1):11-24
  119. 119. Hogan LH, Klein BS, Levitz SM. Virulence factors of medically important fungi. Clinical Microbiology Reviews. 1996;9(4):469-488
  120. 120. Cross C, Bancroft G. Ingestion of acapsular Cryptococcus neoformans occurs via mannose and beta-glucan receptors, resulting in cytokine production and increased phagocytosis of the encapsulated form. Infection and Immunity. 1995;63(7):2604-2611
  121. 121. Yauch LE et al. Involvement of CD14, toll-like receptors 2 and 4, and MyD88 in the host response to the fungal pathogen Cryptococcus neoformans in vivo. Infection and Immunity. 2004;72(9):5373-5382
  122. 122. van der Graaf CA et al. Differential cytokine production and Toll-like receptor signaling pathways by Candida albicans blastoconidia and hyphae. Infection and Immunity. 2005;73(11):7458-7464
  123. 123. Netea MG et al. Aspergillus fumigatus evades immune recognition during germination through loss of toll-like receptor-4-mediated signal transduction. The Journal of Infectious Diseases. 2003;188(2):320-326
  124. 124. Bellocchio S et al. The contribution of the Toll-like/IL-1 receptor superfamily to innate and adaptive immunity to fungal pathogens in vivo. The Journal of Immunology. 2004;172(5):3059-3069
  125. 125. Hirschfeld M et al. Signaling by toll-like receptor 2 and 4 agonists results in differential gene expression in murine macrophages. Infection and Immunity. 2001;69(3):1477-1482
  126. 126. Re F, Strominger JL. Toll-like receptor 2 (TLR2) and TLR4 differentially activate human dendritic cells. Journal of Biological Chemistry. 2001;276(40):37692-37699
  127. 127. Netea MG et al. From the Th1/Th2 paradigm towards a Toll-like receptor/T-helper bias. Antimicrobial Agents and Chemotherapy. 2005;49(10):3991-3996
  128. 128. Stevens DA. Th1/Th2 in aspergillosis. Medical Mycology. 2006;44(Supplement_1):S229-S235
  129. 129. Jankovic D, Liu Z, Gause WC. Th1-and Th2-cell commitment during infectious disease: asymmetry in divergent pathways. Trends in Immunology. 2001;22(8):450-457
  130. 130. Blasi E et al. Biological importance of the two Toll-like receptors, TLR2 and TLR4, in macrophage response to infection with Candida albicans. FEMS Immunology and Medical Microbiology. 2005;44(1):69-79
  131. 131. Netea MG et al. Toll-like receptor 2 suppresses immunity against Candida albicans through induction of IL-10 and regulatory T cells. The Journal of Immunology. 2004;172(6):3712-3718
  132. 132. Vecchiarelli A et al. Purified capsular polysaccharide of Cryptococcus neoformans induces interleukin-10 secretion by human monocytes. Infection and Immunity. 1996;64(7):2846-2849
  133. 133. Chiapello LS et al. Immunosuppression, interleukin-10 synthesis and apoptosis are induced in rats inoculated with Cryptococcus neoformans glucuronoxylomannan. Immunology. 2004;113(3):392-400
  134. 134. Mednick AJ, Nosanchuk JD, Casadevall A. Melanization of Cryptococcus neoformans affects lung inflammatory responses during cryptococcal infection. Infection and Immunity. 2005;73(4):2012-2019
  135. 135. Finkel-Jimenez B et al. The WI-1 adhesin blocks phagocyte TNF-α production, imparting pathogenicity on Blastomyces dermatitidis. The Journal of Immunology. 2001;166(4):2665-2673
  136. 136. Brandhorst TT et al. Exploiting type 3 complement receptor for TNF-α suppression, immune evasion, and progressive pulmonary fungal infection. The Journal of Immunology. 2004;173(12):7444-7453
  137. 137. Stringer JR, Keely SP. Genetics of surface antigen expression in Pneumocystis carinii. Infection and Immunity. 2001;69(2):627-639
  138. 138. Pop SM, Kolls JK, Steele C. Pneumocystis: Immune recognition and evasion. The International Journal of Biochemistry & Cell Biology. 2006;38(1):17-22
  139. 139. Fraser IP et al. Pneumocystis carinii enhancessoluble mannose receptor production by macrophages. Microbes and Infection. 2000;2(11):1305-1310
  140. 140. Fries BC et al. Phenotypic switching of Cryptococcus neoformans occurs in vivo and influences the outcome of infection. The Journal of Clinical Investigation. 2001;108(11):1639-1648
  141. 141. Tavanti A et al. Candida albicans isolates with different genomic backgrounds display a differential response to macrophage infection. Microbes and Infection. 2006;8(3):791-800
  142. 142. Ullmann BD et al. Inducible defense mechanism against nitric oxide in Candida albicans. Eukaryotic Cell. 2004;3(3):715-723
  143. 143. Káposzta R et al. Rapid recruitment of late endosomes and lysosomes in mouse macrophages ingesting Candida albicans. Journal of Cell Science. 1999;112(19):3237-3248
  144. 144. Lorenz MC, Bender JA, Fink GR. Transcriptional response of Candida albicans upon internalization by macrophages. Eukaryotic Cell. 2004;3(5):1076-1087
  145. 145. Richardson MD, Smith H. Resistance of virulent and attenuated strains of Candida albicans to Intracellular Killing by Human and Mouse Phagocytes. The Journal of Infectious Diseases. 1981;144(6):557-564
  146. 146. Goldman DL et al. Persistent Cryptococcus neoformans pulmonary infection in the rat is associated with intracellular parasitism, decreased inducible nitric oxide synthase expression, and altered antibody responsiveness to cryptococcal polysaccharide. Infection and Immunity. 2000;68(2):832-838
  147. 147. Lee SC et al. Cryptococcus neoformans survive and replicate in human microglia. Laboratory Investigation; A Journal of Technical Methods and Pathology. 1995;73(6):871-879
  148. 148. Feldmesser M, Tucker S, Casadevall A. Intracellular parasitism of macrophages by Cryptococcus neoformans. Trends in Microbiology. 2001;9(6):273-278
  149. 149. Woods JP et al. Pathogenesis of Histoplasma capsulatum. Seminars in Respiratory Infections. 2001;16(2):91-101
  150. 150. Porta A, Maresca B. Host response and Histoplasma capsulatum/macrophage molecular interactions. Medical Mycology. 2000;38(6):399-406
  151. 151. Strasser JE et al. Regulation of the macrophage vacuolar ATPase and phagosome-lysosome fusion by Histoplasma capsulatum. The Journal of Immunology. 1999;162(10):6148-6154
  152. 152. Eissenberg LG, Goldman WE, Schlesinger PH. Histoplasma capsulatum modulates the acidification of phagolysosomes. Journal of Experimental Medicine. 1993;177(6):1605-1611
  153. 153. Eissenberg LG, Goldman WE. Histoplasma capsulatum fails to trigger release of superoxide from macrophages. Infection and Immunity. 1987;55(1):29-34
  154. 154. Kozel TR. Complement and its role in fungal diseases. In: Human Fungal Pathogens. Springer; 2004. pp. 193-205
  155. 155. Collette JR, Lorenz MC. Mechanisms of immune evasion in fungal pathogens. Current Opinion in Microbiology. 2011;14(6):668-675
  156. 156. Zipfel PF, Skerka C. Complement regulators and inhibitory proteins. Nature Reviews Immunology. 2009;9(10):729-740
  157. 157. Speth C et al. Complement and fungal pathogens: An update. Mycoses. 2008;51(6):477-496
  158. 158. Speth C, Rambach G. Complement attack against Aspergillus and corresponding evasion mechanisms. Interdisciplinary Perspectives on Infectious Diseases. 2012;2012
  159. 159. Luo S et al. Complement and innate immune evasion strategies of the human pathogenic fungus Candida albicans. Molecular Immunology. 2013;56(3):161-169
  160. 160. Speth C et al. The role of complement in invasive fungal infections. Mycoses. 2004;47(3-4):93-103
  161. 161. Kozel TR. Activation of the complement system by pathogenic fungi. Clinical Microbiology Reviews. 1996;9(1):34-46
  162. 162. Hector R, Yee E, Collins M. Use of DBA/2N mice in models of systemic candidiasis and pulmonary and systemic aspergillosis. Infection and Immunity. 1990;58(5):1476-1478
  163. 163. Tsai H-F et al. The developmentally regulated alb1 gene of Aspergillus fumigatus: Its role in modulation of conidial morphology and virulence. Journal of Bacteriology. 1998;180(12):3031-3038
  164. 164. Meri T et al. The yeast Candida albicans binds complement regulators factor H and FHL-1. Infection and Immunity. 2002;70(9):5185-5192
  165. 165. Vogl G et al. Immune evasion by acquisition of complement inhibitors: The mould Aspergillus binds both factor H and C4b binding protein. Molecular Immunology. 2008;45(5):1485-1493
  166. 166. Behnsen J et al. The opportunistic human pathogenic fungus Aspergillus fumigatus evades the host complement system. Infection and Immunity. 2008;76(2):820-827
  167. 167. Mastellos D, Lambris JD. Complement: More than a ‘guard’against invading pathogens? Trends in Immunology. 2002;23(10):485-491
  168. 168. Ricklin D, Lambris JD. Complement in immune and inflammatory disorders: Pathophysiological mechanisms. The Journal of Immunology. 2013;190(8):3831-3838
  169. 169. Morgan BP, Harris CL. Complement, a target for therapy in inflammatory and degenerative diseases. Nature Reviews Drug Discovery. 2015;14(12):857-877
  170. 170. Reis ES et al. Applying complement therapeutics to rare diseases. Clinical Immunology. 2015;161(2):225-240

Written By

Faisal Rasheed Anjum, Sidra Anam, Muhammad Luqman, Ameena A. AL-surhanee, Abdullah F. Shater, Muhammad Wasim Usmani, Sajjad ur Rahman, Muhammad Sohail Sajid, Farzana Rizvi and Muhammad Zulqarnain Shakir

Submitted: 06 October 2021 Reviewed: 27 October 2021 Published: 31 January 2022