Next Article in Journal
Immunoblot Criteria for Diagnosis of Lyme Disease: A Comparison of CDC Criteria to Alternative Interpretive Approaches
Previous Article in Journal
Avian Pathogenic Escherichia coli (APEC) in Broiler Breeders: An Overview
Previous Article in Special Issue
Brain Tuberculosis: An Odyssey through Time to Understand This Pathology
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Collected Thoughts on Mycobacterial Lipoarabinomannan, a Cell Envelope Lipoglycan

by
Jordi B. Torrelles
1,2,* and
Delphi Chatterjee
3,*
1
International Center for the Advancement of Research and Education (I • Care), Texas Biomedical Research Institute, San Antonio, TX 78227, USA
2
Population Health Program, Texas Biomedical Research Institute, San Antonio, TX 78227, USA
3
Mycobacteria Research Laboratories, Department of Microbiology, Immunology and Pathology, Colorado State University, Fort Collins, CO 80523, USA
*
Authors to whom correspondence should be addressed.
Pathogens 2023, 12(11), 1281; https://doi.org/10.3390/pathogens12111281
Submission received: 30 August 2023 / Revised: 18 October 2023 / Accepted: 24 October 2023 / Published: 26 October 2023
(This article belongs to the Special Issue Pathogenicity of Mycobacterium tuberculosis)

Abstract

:
The presence of lipoarabinomannan (LAM) in the Mycobacterium tuberculosis (Mtb) cell envelope was first reported close to 100 years ago. Since then, numerous studies have been dedicated to the isolation, purification, structural definition, and elucidation of the biological properties of Mtb LAM. In this review, we present a brief historical perspective on the discovery of Mtb LAM and the herculean efforts devoted to structurally characterizing the molecule because of its unique structural and biological features. The significance of LAM remains high to this date, mainly due to its distinct immunological properties in conjunction with its role as a biomarker for diagnostic tests due to its identification in urine, and thus can serve as a point-of-care diagnostic test for tuberculosis (TB). In recent decades, LAM has been thoroughly studied and massive amounts of information on this intriguing molecule are now available. In this review, we give the readers a historical perspective and an update on the current knowledge of LAM with information on the inherent carbohydrate composition, which is unique due to the often puzzling sugar residues that are specifically found on LAM. We then guide the readers through the complex and myriad immunological outcomes, which are strictly dependent on LAM’s chemical structure. Furthermore, we present issues that remain unresolved and represent the immediate future of LAM research. Addressing the chemistry, functions, and roles of LAM will lead to innovative ways to manipulate the processes that involve this controversial and fascinating biomolecule.

1. Introduction

The hallmark of Mycobacterium tuberculosis (Mtb) is a complex hydrophobic lipid-rich cell envelope [1]. The dominant features have been studied by electron microscopy [2] and biochemical fractionation studies [3]. The Mtb cell envelope is mainly composed of four distinct entities [3]. Surrounding the cytosol is the cytoplasmic membrane or innermost membrane, which is similar to other bacterial membranes in thickness and mainly composed of proteins and phospholipids such as phosphatidic acid (PA), diphosphatidylglycerol (PG), phosphatidylcholine (PC), phosphatidylethanolamine (PE), and the phosphatidyl-myo-inositol mannosides (PIMs). The cell envelope lipoarabinomannan (LAM) and lipomannan (LM) are lipoglycans found anchored in the plasma membrane via a phosphatidyl-myo-inositol moiety that is the structural basis of the PIMs. Beyond the cell membrane, peptidoglycan (PG) and its attached arabinogalactan (AG), which together form the AGP complex (AGP), are apparently located in the periplasmic space. Outermost is an outer membrane or “mycomembrane”, which consists of the mycolic acids that form a distinct layer [4]. This peripheral lipid-rich cell envelope and its distinct components have been defined as virulent factors due to their ability to trigger a host response detrimental to the host and favoring the establishment of an Mtb infection. A debate still exists on the exact location of the PIMs, LM, and LAM. The latter may be anchored in the plasma membrane during its biosynthesis but also exposed on the cell envelope surface. Evidence of the existence of “kettle holes” on the cell envelope poles, which could allow the surface exposure of plasma membrane anchored LAM [5,6,7,8], could explain this seeming contradiction. The concept of the exposure of LAM on the surface is supported by quantitative transmission electron microscopy studies indicating that LAM is mainly located in the bacillus poles/tips and within cell envelope surface depressions [9,10,11]. Meanwhile, the concept of LAM anchored to the plasma membrane is supported by the need for mechanical and chemical steps to extract LAM from the bacterial cell envelope [12] and by the fact that many steps in the LAM biosynthesis are associated with the plasma membrane [13,14].

2. Defining the Basic Structure of LAM: A Brief History

The volume of literature concerning the discovery of LAM, followed by its structural characterization and discovery of its immunological properties, is substantial. A timeline of the evolution of LAM structural studies is presented in Figure 1. Briefly, in 1925, Laidlaw and Dudley [15] used a mild alkaline extraction of a defatted human strain of Mtb to obtain a novel component that contained 31% pentose. Later, Heidelberger and Menzel using dilute acetic acid on defatted Mtb [16], as well as Laidlaw and Dudley and Mueller using culture filtrate, showed the presence of a serologically active polysaccharide rich in D-arabinose and D-mannose [17]. The presence of this polysaccharide was confirmed further in avian and bovine strains [18]. Over the next 10 years, several studies followed, all concluding the presence of a certain polysaccharide composed of D-mannose, D-arabinose, and an unidentified sugar acid that reacted as haptens.
The potential first evidence of a mycobacterial LAM-like structure was established in 1935 by Chargaff and Schaefer, who identified the polysaccharide, designated as fraction A, as being a soluble, dextrorotatory, weak acid containing 77% reducing sugars and 3% amino sugars, mainly composed of mannose and arabinose together with small amounts of inositol [19]. Since the time that this study was reported, many subsequent studies focused on defining the chemical and immunological properties of this polysaccharide, including its α-haemosensitin properties involved in the Middlebrook–Dubos hemagglutination reaction.
In 1977, a serologically active and LAM-related D-arabino-D-mannan from Mtb was characterized with a 2:1 ratio of arabinose:mannose. This arabinomannan consisted of α-(1 → 5)-linked D-arabinose residues and α-(1 → 6)- and (1 → 2)-linked D-mannose residues. Further methylation and enzymatic degradation studies using Arthrobacter sp. α-D-mannosidase and M-2 enzyme (D-arabinan hydrolase) provided evidence of the existence of short side chains built up from α-(1 → 2)-D-mannosidic linkages that were attached to an α-(1 → 6)-linked mannan backbone [20].
In 1979, Weber and Gray [21] isolated an acidic arabinomannan from M. smegmatis and defined the presence of 56 arabinosyl and 11 mannosyl residues, 2 phosphates, 6 mono-esterified succinates, and 4 ether-linked lactate groups. Subsequently, this acidic polysaccharide was separated into phosphorylated and non-phosphorylated forms with similar structures, wherein the main structural feature was the presence of chains of contiguous arabinofuranosyl residues linked α-(1 → 5).
Close to 50 years since the first reported evidence of a polysaccharide containing arabinose, mannose, inositol, and phosphorous in Mtb, in 1986, Hunter et al. structurally defined a family of major arabinose- and mannose-containing phosphorylated lipopolysaccharides isolated from Mtb and M. leprae [22]. LAM-B, as designated, eluting off an ion exchange column, in addition to arabinose and mannose, also contained glycerol and a myo-inositol 1-phosphate, as well as acylations with lactate, succinate, palmitate, and 10-methyloctadecanoate. Other studies followed that provided details of the arabinose-to-mannose ratios and the presence of palmitic, stearic, and tuberculostearic acids as the main fatty acids [23]. Brennan and colleagues, when characterizing the fully acetylated PIM family and their biosynthesis, realized that LAM was essentially a polyglycosylated extension of the PIMs [24,25]. Confirmation was provided in 1990 by Hunter and Brennan, who established the presence of the phosphatidyl-myo-inositol membrane anchor present in the PIMs in Mtb LAM and LM [26]. The most decisive feature of this study was the evidence for the presence of glycosidically linked diacylglycerol residues, in common with the phosphatidylinositol (PI) of the PIMs [24].
Chatterjee and colleagues, using enzymatic treatments, subsequently focused on the exact architecture of the arabinan domain, establishing that its internal regions consist of branched 3,5-linked α-D-Araf units with stretches of linear 5-linked α-D-Araf residues attached at both branch positions, whereas the non-reducing terminal segments of the LAM arabinan domain consist of either of the two arrangements, Ara4: β-D-Araf-(1 → 2)-α-D-Araf-(1 → 5)-α-D-Araf-(1 → 5)-α-D-Araf → or Ara6: [β-D-Araf-(1 → 2)-α-D-Araf-(1 → ]2 → (3 and 5)-α-D-Araf-(1 → 5)-α-D-Araf → [27] (Figure 2). Two subsequent studies by Chatterjee and colleagues further refined the LAM structure. The first focused on describing the structural environment of the PI anchor, showing clearly the presence of the 1-(sn-glycerol-3-phospho)-D-myo-inositol-2,6-bis-α-D-mannoside unit, indistinguishable from that derived from phosphatidyl-myo-inositol dimannoside (PIM2). The same study also demonstrated that the C-6 position of inositol is the site of attachment of the mannan core of LAM, which consists of an α(1 → 6)-linked backbone with several α(1 → 2)-side chains [28]. In a parallel study, Chatterjee and colleagues also demonstrated that the termini of LAM from virulent Mtb strains, unlike those from attenuated Mtb strains, are extensively capped with mannosyl (Manp) residues, either with a single α-D-Manp, a dimannoside (α-D-Manp-(1 → 2)-α-D-Manp) or a trimannoside (α-D-Manp-(1 → 2)-α-D-Manp-(1 → 2)-α-D-Manp). Thereby, the functionally important so-called mannose-capped LAM or Man-LAM was identified (Figure 3). Indeed, after this work, a large number of studies followed, uncovering a wide array of biological properties attributable to the various forms of Mtb LAM has, and how Mtb LAM compares to LAMs from other Mtb complexes and mycobacterial species [29,30,31,32,33]. In particular, these subsequent studies, using advanced HPLC resolution and MS analysis, demonstrated the enormous intrinsic heterogeneity and complexity of LAM at both the reducing and non-reducing ends, as well as variations in the fatty acyl content of the GPI anchor. Variations in the length and branching patterns of the D-mannan and D-arabinan domains were also observed. These, and potentially other short-chain acyl features, have now been established as attached to the arabinan domain. Further, diversity within the Man-caps, in particular, the presence of a 5-deoxy-5-methylthio-xylofuranose (MTX) residue, has been observed [34,35,36]. Finally, studies also depicted the effects of specific drugs (e.g., ethambutol) on the structure of LAM in drug-resistant Mtb strains and other mycobacterial species, as well as how the size of LAM influences Mtb/host cell interactions [30,37,38]. Current efforts are directed towards advancing the native structure of LAM isolated from tissues and/or specimens, as well as developing new tools for the structural characterization of LAM [34,39].

3. Evolution of LAM Structural Studies across Species

Our knowledge of the structure of LAM is constantly evolving [41,42], and although there are differences across species, the overall features are common to all, at least in all those investigated to date. Essentially, LAM throughout contains four structural domains: a mannosylated phosphatidyl myo-inositol (PI) anchor, a D-mannan core, a D-arabinan domain, and different capping motifs that contribute to species and strain diversity (reviewed in [43,44]). The mannan core consists of a chain of α-(1 → 6)-linked mannopyranose (Manp) residues, some of which are modified by the addition of α-(1 → 2)-linked Manp motifs, usually, but not always, as a single residue. An arabinan, composed of solely D-arabinofuranose (Araf) residues, is attached to the non-reducing end of the mannan core [42]. Capping motifs can be added at specific positions contributing to intra- and inter-species structural variability [41]. Fast-growing mycobacterial species predominately produce AraLAM (uncapped LAM) or PILAM (phosphoinositol capped, as defined for M. smegmatis) [45]. Slow-growing mycobacteria like Mtb and M. leprae produce LAM with α-(1 → 2)-linked Manp capping residues, giving a molecule referred to as ManLAM [8,30]. Within the Mtb complex group, variations regarding primarily the degree of terminal mannose capping present in ManLAM can range between 40–70% [27,28,46,47,48]. Some fast-growing and/or non-pathogenic mycobacteria also produce ManLAM; however, these differ in the Man content of the capping motifs. In addition to Manp capping, ManLAM from strains of the Mtb complex group also contains a unique residue –MTX on the terminal Manp caps [34,35,36,49]. Further variability comes from acylation of the arabinan, most commonly with succinylation (Figure 3), which can be found either in the internal [50] or termini [41] of the arabinan domain [51].
The extreme heterogeneity in LAM is evident from the broad diffuse band observed on a SDS-polyacrylamide gel electrophoresis (PAGE) analysis of LAM and LM [22] and its capacity to be separated in different isoforms [31], as well as from several recent matrix-assisted laser desorption/ionization—mass spectrometry (MALDI-MS) studies, which provided an indication of the mean distribution of true molecular mass. It has been shown that native LAM from M. bovis BCG and Mtb gives a broad peak centered at 17.3 kDa before deacylation and 16.7 kDa after deacylation, with a reported size distribution range of ±4 kDa depending on the studied strains [52]. With recent advances in MS instrumentation, a peak centered at m/z 14439.921, providing a molecular mass of approx. 15 kDa for LAM, has been reported (Chatterjee et al., personal communication) (Figure 4).

4. Biological Properties of Mtb LAM

Mtb LAM was widely studied in the 1990s. Studies started to reveal the immunological properties of LAM, initially as a potential candidate for ELISA-based diagnosis of TB and leprosy [53,54], as well as its capacity to inhibit the activation of macrophages, among other immunological features [55,56].
Early data published include LAM-induced abrogation of T-cell activation [57]; inhibition of various IFN-γ-induced functions including macrophage microbicidal and tumoricidal activity [56]; scavenging of potentially cytotoxic oxygen free radicals [58]; inhibition of protein kinase C activity [58]; and evocation of a large array of cytokines associated with macrophages such as TNF [59,60,61,62,63], GM-CSF, IL-1α, IL-1β, IL-6, and IL-10 [64,65]. In 1991, Chan and colleagues provided the first evidence of the role of Mtb LAM in downregulating macrophage effector functions by scavenging potentially cytotoxic oxygen free radicals, inhibiting protein kinase C activity, and blocking the transcriptional activation of IFN-γ inducible genes [58]. Related to the use of LAM by Mtb to recognize and infect host cells, it has been reported that both the mannose receptor (MR) and DC-SIGN in phagocytes recognize and bind to LAM, and this binding can result in efficient internalization of Mtb to its intracellular niche within host cells [9,66].
Of the many biological properties that Mtb ManLAM has exhibited, one has led us to understand how critical this molecule is for the persistence of Mtb within host cells. First, Schlesinger and colleagues showed that ManLAM is recognized by macrophages [67]. This finding was later supported by others [68]. Subsequently, Deretic and others showed the capacity of Mtb ManLAM to block phagosome acidification [69], and later, Schlesinger and colleagues showed how Mtb exploits the ManLAM/MR route [70] to gain entry into macrophages and survive [9], potentially generating the perfect niche for survival and subsequent immune responses deriving into granuloma formation and thus, contributing to the persistence of Mtb in a latent stage in the host. Another well-known characteristic of ManLAM is its intracellular processing and presentation towards CD1-restricted T cells, contributing to the host immune response against Mtb infection [31,71,72]. In this regard, liposomal delivery of ManLAM triggers ManLAM-specific T cells [73]. The ManLAM phosphatidyl-myo-inositol moiety plays a central role in its binding to CD1b, although the exact epitopes involved in this binding have not yet been defined [31]. Thus, ManLAM is shown to be a CD1b ligand and some of its in vivo properties are probably derived from the binding of CD1-restricted T cells [71]. Further details on LAM biological functions are described in detail elsewhere [43].

5. LAM as a Diagnosis Biomarker for TB Disease

Mtb LAM represents up to 15% of the bacterial mass [1,26,74,75]. This molecule is firmly but non-covalently attached to the inner membrane and extends to the exterior of the cell envelope [76], where it interacts as a potent virulence factor modulating host immune responses [77]. Importantly, the linear terminus Ara4: (β-D-Araf-(1 → 2)-α-D-Araf-(1→5)-α-D-Araf-(1 → 5)-α-D-Araf) (Figure 2) and a branched terminus Ara6: ([β-D-Araf-(1→2)-α-D-Araf-(1-)2 → 3, and →5]-α-D-Araf-(1→5)-α-D-Araf) (Figure 2) are shown to be the epitopes recognized by anti-LAM monoclonal antibodies (mAb) [40,74,78,79,80,81].
LAM has been validated to be present in variable concentrations in sputum, serum, and urine [82,83,84]. In recent years, several laboratories including ours have made significant advances toward developing urinary LAM-based diagnostics for active TB [40,79,85,86,87]. A point-of-care (POC) test that readily diagnoses active TB would reduce diagnostic delays, interrupt transmission with appropriate therapy, and address many of the current gaps in global TB control (Stop TB Partnership in collaboration with Imperial College London). The development of sensitive POC methods to detect LAM in non-invasive samples such as urine using immunoassays is currently stagnant due to the suboptimal sensitivity of the assays. Current methods are also limited in applicability to TB diagnosis only in people living with HIV and severe disease [84,88,89]. However, there are studies showing that LAM can be also detected in urine from active TB cases without HIV co-infection [85,90]. Recently, as an alternative to urine, groups have focused their attention on other non-invasive samples such as exhaled breath condensate (EBC), in which ManLAM seems to be present in significant amounts [91,92].
We and the others actively pursuing the field have hypothesized that these disappointing results with POC methods are due to the array of anti-LAM antibodies used in the current commercially available POC immunoassays, which are generally raised against one common Mtb laboratory strain [40,79,87,93]. Thus, Mtb lineage-specific and demographical strain anti-LAM antibodies may be required to develop more efficient LAM detection-based POC diagnostic tests for their use in TB-endemic countries.
We should also add that the structure of LAM differs among mycobacterial species and strains. In the context of non-typical mycobacteria (NTM) such as M. avium, from the few studies performed, the LAM structure differs mainly in the degree of mannose caps present. Although Mtb presents tri-, di-, and mono-mannose-capped LAM, M. avium LAM mainly contains mono-mannose-capped LAM [48,49]. From the diagnosis perspective, LAM is well-established as a biomarker for active TB, with high sensitivity and specificity by immunological assays. Recently, urinary LAM was considered as a biomarker for determining cystic fibrosis (CF) patients at low risk of NTM infection [94]. In this regard, the urine LAM test is reported to have high specificity (91–99%) but low sensitivity (9–39%) for pulmonary NTM in the Danish CF population [95]. We have reported that CF patients with recent positive NTM sputum cultures have detectable amounts of LAM in their urine (measured by Gas Chromatography–Mass Spectrometry), with a 100% concordance between a non-detectable quantity of urine LAM and a history of negative NTM sputum culture. We have also reported urinary LAM being a sensitive marker of treatment response in an individual with M. abscessus successfully treated with phage therapy [96]. However, we should emphasize that in the NTMs and M. abscessus cases studied, immunoassays using TB-LAM-specific antibodies failed to detect LAM in urine, even after several-fold concentrations. We have speculated that there could be several reasons for this—that LAM is present albeit in very low amounts, or there are fundamental differences between TB-LAM vs. NTM-LAM structures, or how these are presented. While detecting low LAM concentrations in diluted urine is part of the challenge, a key gap in our existing knowledge remains related to which LAM recognition motifs are most critical for highly sensitive LAM detection.
As of today, LAM from NTM has not been studied in depth. Thus, it is not apparent if any of the features such as Man caps, MTX, or acylation are present/absent in NTM-LAM. Using well-phenotyped collected CF samples [94], we have tested anti-LAM monoclonal antibodies using CF urine samples with LAM concentrations previously determined by GC-MS. Dot blot screening indicated that antibodies have only moderate activity towards NTM-LAM compared to TB-LAM, despite a fairly high concentration (~250 ng/mL) of LAM spiked in the urine samples. This suggests that either the epitopes of NTM-LAM vs. TB-LAM are different, or unlike in TB, during NTM infection LAM is cryptic and is somehow not available to the antibodies for recognition due to biofilm formation. In either case, we will need to generate specific monoclonal and polyclonal antibodies to be able to increase the sensitivity of any immunoassay in detecting NTM-LAM [81].

6. Concluding Remarks

The structure and biological functions of LAM are still widely studied to understand its importance in the context of Mtb pathogenesis and drug resistance. Over time, new protocols have been developed to extract, purify, and accurately characterize the LAM molecule. Current knowledge on the structure of LAM has resulted primarily from detailed studies on a few selected laboratory strains of Mtb [31,74], M. bovis bacillus Calmette–Guérin (BCG) [29], M. smegmatis [32], and M. kansasii [97], among a few others. Efforts have been invested in correlating unique structural features with aspects of the immunopathogenesis of TB [74]. An outcome of these efforts is the consensus that the mannose caps of LAM (ManLAM) constitute the single most important structural entity engaged in phagocytosis by phagocytes and subsequent events such as inhibition of phagosome/lysosome fusion and immunomodulation of host responses.
Indeed, the presence of LAM on the cell surface allows Mtb to mimic mammalian glycoproteins that are cleared subsequently from circulation by phagocytes using specific receptors such as the MR through the recognition of the LAM mannose caps. This phenomenon confers on Mtb a unique pathway of entry and survival within host cells. However, how LAM is being processed metabolically within host cells and its role in Mtb intracellular survival over a long period is, as yet, not well defined. In this context, it is speculated that the ratio of mannose-containing molecules (PIMs, LM, ManLAM, AM, and mannosylated glycoproteins) on the Mtb cell envelope may play a role in mitigating host immune responses, as well as contributing to the bacillus dormancy metabolic status within host cells [98]. However, how the degree of surface mannosylation on different Mtb strains correlates to their pathogenesis and their development of drug resistance is still unknown.
Although the overall structure and properties of LAM are conserved, its chemical composition varies among Mtb complex species and Mtb strains. In this regard, the human-adapted Mtb complex exhibits a strong phylogeographical population structure, with some Mtb lineages occurring globally and others showing a strong geographical restriction [99,100]. Among these, lineages L2 and L4 are widely distributed in the world, with L2 dominating in East Asia. L1 and L3 are found mainly in regions around the Indian Ocean, and L5 and L6 seem restricted to West Africa, whereas L7 seems to be exclusively found in Ethiopia. We posit that Mtb strains from different lineages will have a wide spectrum of virulence, differently modulating host immune responses, and thus, determining Mtb bacillary load in patients with pulmonary TB. In this context, it will be important to perform LAM phenotype mapping in Mtb strains that cause TB outbreaks in TB-endemic geographical areas, and further examine whether any epidemiologically relevant LAM structural characteristics are associated with these strains [41].
We end this review with some additional thoughts. We have introduced a historical perspective on how several laboratories including ours, starting with problems in the field of structural definition, have gradually become involved in the issues of biosynthesis, immunology, pathogenesis, and finally, in TB diagnosis. One of the vital contributions that we bring to such progress is the sensitivity towards precisely defined structural LAM attributes. Much has been learned about the chemistry and biology of LAM and yet, this knowledge is still evolving, as it is obvious from this review. As an example, unlike many other known glycans, LAM arabinan is a homopolymer with no repeating units, and a single donor decaprenylphosphoryl arabinose is identified to be donating all of the arabinofuranose residues [101]. Its location and distribution within the Mtb cell envelope remain unresolved.
Studies indicate that LAM is found in urine, serum, EBC, and lung biopsies from TB patients [34,91,92,93,102]. The fact that LAM is found in the serum of TB patients means that it probably affects a wide variety of host cell populations, consequently influencing both innate and adaptive immune responses during Mtb infection, dormancy, subclinical TB, and progression to active TB disease. Future studies on delineating the roles of LAM in immunopathogenesis need to be focused on in vivo studies whenever possible, with the hope that we will eventually delineate the LAM structural motifs that contribute as virulence factors and/or protective epitopes in vivo.
We believe that, due to their chemistry, LAM molecules are key glycans sitting at the crossroads of many critical biological processes. As such, we (the authors) strongly deem that the very final goal of LAM research should be to have an “atomic architecture” of the LAM structure in which its relevant features are finely described with their biological functions dissected. In the coming years, accomplishing this complex but fundamental task will have a significant impact on advancing the TB field.
LAM biosynthesis (a topic not addressed in this review) is not trivial and encompasses a complex network of enzymes with high energy demand. Thus, the necessity for Mtb to produce LAM to survive seems apparent. To this day, there are many biological properties described for LAM, but it is still unclear if the presence of LAM in the Mtb cell envelope ultimately benefits the bacterium or the host. In this regard, Mtb strains with a highly mannosylated cell envelope, including LAM, are thought to be adapted to the host, driving the dormant stage of the infection. Thus, a question that remains is: are the innate and adaptive immune responses against LAM protecting the bacterium, allowing Mtb infection to perpetuate? In this context, we have shown that upon minimal contact (15 min) with human alveolar lining fluid, the Mtb cell envelope is modified by alveolar homeostatic hydrolytic activities releasing cell envelope fragments into the milieu [103]. As a result, Mtb loses up to 65% of its LAM exposed on the envelope surface [103]. This released LAM could be binding with alveolar soluble innate components such as surfactant protein D (SP-D) and mannose-binding lectin (MBL), blocking innate responses that ultimately favor Mtb survival. This is yet to be proven.
Little is known about the role of LAM in the early events of Mtb infection driving the formation of the granuloma. As a future perspective in LAM research, there is a need for improved methods for studying the LAM structure in vivo. As an unquestionable hot topic, innovative, although certainly challenging, procedures, able to deliver the relevant information, would provide a realistic picture of the actual structure of LAM without any chemical alterations possibly related to the current isolation methods used. This would be a giant leap forward in elucidating any links between LAM’s specific structural characteristics, its real biological functions, and the disease arising from its detection by the host immune system.

Author Contributions

J.B.T. and D.C. made substantial contributions to the conception and design of this review, and gave their final approval of the version to be submitted. All authors have read and agreed to the published version of the manuscript.

Funding

External funding was provided by NIH/NIAID R01AI-037139 and NIH/NIAID R01AI-132680 to D.C; and partially by the Robert J. Kleberg Jr and Helen C. Kleberg Foundation to J.B.T. J.B.T. is part of the Interdisciplinary NextGen Tuberculosis Research Advancement Center (IN-TRAC) at Texas Biomed, which is supported by the National Institutes of Health/National Institute of Allergy and Infectious Diseases (NIH/NIAID) under the award number P30 AI-168439. The mass spectrometry (MS) was done on a Bruker ultrafleXtreme matrix-assisted laser desorption time-of-flight spectrometer (MALDI-TOF/TOF MS), which was funded by the National Science Foundation (NSF) under grant MRI-2117934 and supported by the Analytical Resources Core (RRID: SCR_021758).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The research reviewed in this article was possible only through the dedication, enthusiasm, and creativity of scores of coworkers, whose names are acknowledged in the publications cited from this laboratory. We wish to record our gratitude to Patrick J. Brennan (Hon.), University Distinguished Professor (Emeritus), Colorado State University (CSU). He introduced Delphi Chatterjee to the topic of LAM structure in the early 1990s and she in turn guided Jordi B. Torrelles in his studies on LAM. We thank Anita Amin and the Analytical Resources Core (ARC) facility at CSU for assistance with the MALDI-TOF/TOF MS studies of LAM.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Brennan, P.J.; Nikaido, H. The envelope of mycobacteria. Annu. Rev. Biochem. 1995, 64, 29–63. [Google Scholar] [CrossRef]
  2. Dahl, J.L. Electron microscopy analysis of Mycobacterium tuberculosis cell division. FEMS Microbiol. Lett. 2004, 240, 15–20. [Google Scholar] [CrossRef]
  3. Chiaradia, L.; Lefebvre, C.; Parra, J.; Marcoux, J.; Burlet-Schiltz, O.; Etienne, G.; Tropis, M.; Daffé, M. Dissecting the mycobacterial cell envelope and defining the composition of the native mycomembrane. Sci. Rep. 2017, 7, 12807. [Google Scholar] [CrossRef]
  4. Daffé, M.; Marrakchi, H. Unraveling the Structure of the Mycobacterial Envelope. Microbiol. Spectr. 2019, 7. [Google Scholar] [CrossRef]
  5. Dulberger, C.L.; Rubin, E.J.; Boutte, C.C. The mycobacterial cell envelope—A moving target. Nat. Rev. Microbiol. 2020, 18, 47–59. [Google Scholar] [CrossRef]
  6. Gaylord, H.; Brennan, P.J. Leprosy and the leprosy bacillus: Recent developments in characterization of antigens and immunology of the disease. Annu. Rev. Microbiol. 1987, 41, 645–675. [Google Scholar] [CrossRef]
  7. McNeil, M.R.; Brennan, P.J. Structure, function and biogenesis of the cell envelope of mycobacteria in relation to bacterial physiology, pathogenesis and drug resistance; some thoughts and possibilities arising from recent structural information. Res. Microbiol. 1991, 142, 451–463. [Google Scholar] [CrossRef]
  8. Chatterjee, D.; Khoo, K.-H. Mycobacterial lipoarabinomannan: An extraordinary lipoheteroglycan with profound physiological effects. Glycobiology 1998, 8, 113–120. [Google Scholar] [CrossRef] [PubMed]
  9. Kang, P.B.; Azad, A.K.; Torrelles, J.B.; Kaufman, T.M.; Beharka, A.; Tibesar, E.; DesJardin, L.E.; Schlesinger, L.S. The human macrophage mannose receptor directs Mycobacterium tuberculosis lipoarabinomannan-mediated phagosome biogenesis. J. Exp. Med. 2005, 202, 987–999. [Google Scholar] [CrossRef] [PubMed]
  10. Hett, E.C.; Rubin, E.J. Bacterial growth and cell division: A mycobacterial perspective. Microbiol. Mol. Biol. Rev. 2008, 72, 126–156. [Google Scholar] [CrossRef] [PubMed]
  11. Torrelles, J.B. Broadening our view about the role of Mycobacterium tuberculosis cell envelope components during infection: A battle for survival. In Understanding Tuberculosis—Analyzing the Origin of Mycobacterium tuberculosis Pathogenicity; Cardona, P.J., Ed.; Intech: Rijeka, Croatia, 2012; pp. 1–46. [Google Scholar]
  12. Shi, L.; Torrelles, J.B.; Chatterjee, D. Lipoglycans of Mycobacterium tuberculosis: Isolation, purification, and characterization. In Mycobacteria Protocols, 2nd ed.; Parish, T., Brown, A.C., Eds.; Humana Press: Totowa, NJ, USA, 2008; pp. 23–45. [Google Scholar]
  13. Besra, G.S.; Morehouse, C.B.; Rittner, C.M.; Waechter, C.J.; Brennan, P.J. Biosynthesis of mycobacterial lipoarabinomannan. J. Biol. Chem. 1997, 272, 18460–18466. [Google Scholar] [CrossRef] [PubMed]
  14. Kaur, D.; Guerin, M.E.; Skovierova, H.; Brennan, P.J.; Jackson, M. Chapter 2: Biogenesis of the cell wall and other glycoconjugates of Mycobacterium tuberculosis. Adv. Appl. Microbiol. 2009, 69, 23–78. [Google Scholar]
  15. Laidlaw, P.P.; Dudley, H.W. A specific pre-cipitating substance from tubercle bacilli. Br. J. Exp. Pathol. 1925, 6, 197–201. [Google Scholar]
  16. Heidelberger, M.; Menzel, A.E. Specific and non-specific cell polysaccharides of the humantype of tubercle bacillus, H37. Proc. Soc. Exp. Biol. Med. 1932, 29, 631–633. [Google Scholar] [CrossRef]
  17. Mueller, J.H. A chemical study of the specific elements of tuberculin: II. The preparation of resi¬due antigen from old tuberculin. J. Exp. Med. 1926, 43, 9–12. [Google Scholar] [CrossRef] [PubMed]
  18. Menzel, A.E.; Heidelberger, M. Specific and non-specific cell polysaccharides of an avian strain of tubercle bacillus. J. Biol. Chem. 1939, 127, 221–236. [Google Scholar] [CrossRef]
  19. Chargaff, E.; Schaefer, W. A specific polysaccharide from the Bacillus Calmette-Guerin (BCG). J. Biol. Chem. 1935, 112, 393–405. [Google Scholar] [CrossRef]
  20. Misaki, A.; Azuma, I.; Yamamura, Y. Structural and immunochemical studies on D-arabino-D-mannans and D-mannans of Mycobacterium tuberculosis and other Mycobacterium species. J. Biochem. 1977, 82, 1759–1770. [Google Scholar] [CrossRef]
  21. Weber, P.L.; Gray, G.R. Structural and immunochemical characterization of the acidic arabinomannan of Mycobacterium smegmatis. Carbohydr. Res. 1979, 74, 259–278. [Google Scholar] [CrossRef]
  22. Hunter, S.W.; Gaylord, H.; Brennan, P.J. Structure and antigenicity of the phosphorylated lipopolysaccharide antigens from the leprosy and tubercle bacilli. J. Biol. Chem. 1986, 261, 12345–12351. [Google Scholar] [CrossRef]
  23. Sugden, E.A.; Samagh, B.S.; Bundle, D.R.; Duncan, J.R. Lipoarbinomannan and Lipid-free arabinomannan antigens of Mycobacterium paratuberculosis. Infect. Immun. 1987, 55, 762–770. [Google Scholar] [CrossRef]
  24. Brennan, P.; Ballou, C.E. Biosynthesis of mannophosphoinositides by Mycobacterium phlei: Enzymatic acylation of the dimannophosphoinositides. J. Biol. Chem. 1968, 243, 2975–2984. [Google Scholar] [CrossRef] [PubMed]
  25. Brennan, P.J.; Ballou, C.E. Biosynthesis of Mannophosphoinositides by Mycobacterium phlei: The family of dimannosylphosphoinositides. J. Biol. Chem. 1967, 242, 3046–3056. [Google Scholar] [CrossRef]
  26. Hunter, S.W.; Brennan, P.J. Evidence for the presence of a phosphatidylinositol anchor on the lipoarabinomannan and lipomannan of Mycobacterium tuberculosis. J. Biol. Chem. 1990, 265, 9272–9279. [Google Scholar] [CrossRef] [PubMed]
  27. Chatterjee, D.; Bozic, C.M.; McNeil, M.; Brennan, P.J. Structural features of the arabinan component of the lipoarabinomannan of Mycobacterium tuberculosis. J. Biol. Chem. 1991, 266, 9652–9660. [Google Scholar] [CrossRef] [PubMed]
  28. Chatterjee, D.; Hunter, S.W.; McNeil, M.; Brennan, P.J. Lipoarabinomannan. Multiglycosylated form of the mycobacterial mannosylphophatidylinositols. J. Biol. Chem. 1992, 267, 6228–6233. [Google Scholar] [CrossRef]
  29. Prinzis, S.; Chatterjee, D.; Brennan, P.J. Structure and antigenicity of lipoarabinomannan from Mycobacterium bovis BCG. J. Gen. Microbiol. 1993, 139, 2649–2658. [Google Scholar] [CrossRef]
  30. Torrelles, J.B.; Khoo, K.H.; Sieling, P.A.; Modlin, R.L.; Zhang, N.; Marques, A.M.; Treumann, A.; Rithner, C.D.; Brennan, P.J.; Chatterjee, D. Truncated Structural Variants of Lipoarabinomannan in Mycobacterium leprae and an Ethambutol-resistant Strain of Mycobacterium tuberculosis. J. Biol. Chem. 2004, 279, 41227–41239. [Google Scholar] [CrossRef]
  31. Torrelles, J.B.; Sieling, P.A.; Zhang, N.; Keen, M.A.; McNeil, M.R.; Belisle, J.T.; Modlin, R.L.; Brennan, P.J.; Chatterjee, D. Isolation of a distinct Mycobacterium tuberculosis mannose-capped lipoarabinomannan isoform responsible for recognition by CD1b-restricted T cells. Glycobiology 2012, 22, 1118–1127. [Google Scholar] [CrossRef]
  32. Khoo, K.-H.; Dell, A.; Morris, H.R.; Brennan, P.J.; Chatterjee, D. Inositol phosphate capping of the nonreducing termini of lipoarabinomannan from rapidly growing strains of Mycobacterium. J. Biol. Chem. 1995, 270, 12380–12389. [Google Scholar] [CrossRef]
  33. Petzold, C.J.; Stanton, L.H.; Leary, J.A. Structural characterization of Lipoarabinomannans from Mycobacterium tuberculosis and Mycobacterium smegmatis by ESI mass spectrometry. J. Am. Soc. Mass Spectrom. 2005, 16, 1109–1116. [Google Scholar] [CrossRef] [PubMed]
  34. De, P.; Shi, L.; Boot, C.; Ordway, D.; McNeil, M.; Chatterjee, D. Comparative Structural Study of Terminal Ends of Lipoarabinomannan from Mice Infected Lung Tissues and Urine of a Tuberculosis Positive Patient. ACS Infect. Dis. 2020, 6, 291–301. [Google Scholar] [CrossRef]
  35. Treumann, A.; Xidong, F.; McDonnell, L.; Derrick, P.J.; Ashcroft, A.E.; Chatterjee, D.; Homans, S.W. 5-Methylthiopentose: A new substituent on lipoarabinomannan in Mycobacterium tuberculosis. J. Mol. Biol. 2002, 316, 89–100. [Google Scholar] [CrossRef] [PubMed]
  36. Turnbull, W.B.; Shimizu, K.H.; Chatterjee, D.; Homans, S.W.; Treumann, A. Identification of the 5-methylthiopentosyl substituent in Mycobacterium tuberculosis lipoarabinomannan. Angew. Chem. Int. Ed. Engl. 2004, 43, 3918–3922. [Google Scholar] [CrossRef] [PubMed]
  37. Torrelles, J.B.; Knaup, R.; Kolareth, A.; Slepushkina, T.; Kaufman, T.M.; Kang, P.B.; Hill, P.; Brennan, P.J.; Chatterjee, D.; Belisle, J.T.; et al. Identification of mycobacterium tuberculosis clinical isolates with altered phagocytosis by human macrophages due to a truncated lipoarabinomannan. J. Biol. Chem. 2008, 283, 31417–31428. [Google Scholar] [CrossRef] [PubMed]
  38. Zhang, N.; Torrelles, J.B.; McNeil, M.R.; Escuyer, V.E.; Khoo, K.H.; Brennan, P.J.; Chatterjee, D. The Emb proteins of mycobacteria direct arabinosylation of lipoarabinomannan and arabinogalactan via an N-terminal recognition region and a C-terminal synthetic region. Mol. Microbiol. 2003, 50, 69–76. [Google Scholar] [CrossRef]
  39. Al-Jourani, O.; Benedict, S.T.; Ross, J.; Layton, A.J.; van der Peet, P.; Marando, V.M.; Bailey, N.P.; Heunis, T.; Manion, J.; Mensitieri, F.; et al. Identification of D-arabinan-degrading enzymes in mycobacteria. Nat. Commun. 2023, 14, 2233. [Google Scholar] [CrossRef]
  40. Amin, A.G.; De, P.; Spencer, J.S.; Brennan, P.J.; Daum, J.; Andre, B.G.; Joe, M.; Bai, Y.; Laurentius, L.; Porter, M.D.; et al. Detection of lipoarabinomannan in urine and serum of HIV-positive and HIV-negative TB suspects using an improved capture-enzyme linked immuno absorbent assay and gas chromatography/mass spectrometry. Tuberculosis 2018, 111, 178–187. [Google Scholar] [CrossRef]
  41. De, P.; Amin, A.G.; Flores, D.; Simpson, A.; Dobos, K.; Chatterjee, D. Structural implications of lipoarabinomannan glycans from global clinical isolates in diagnosis of Mycobacterium tuberculosis infection. J. Biol. Chem. 2021, 297, 101265. [Google Scholar] [CrossRef]
  42. Angala, S.; Li, W.; Boot, C.M.; Jackson, M.; McNeil, M.R. Secondary Extended Mannan Side Chains and Attachment of the Arabinan in Mycobacterial Lipoarabinomannan. Commun. Chem. 2020, 3, 101. [Google Scholar] [CrossRef]
  43. Turner, J.; Torrelles, J.B. Mannose-capped lipoarabinomannan in Mycobacterium tuberculosis pathogenesis. Pathog. Dis. 2018, 76, fty026. [Google Scholar] [CrossRef]
  44. Nigou, J.; Gilleron, M.; Puzo, G. Lipoarabinomannans: From structure to biosynthesis. Biochimie 2003, 85, 153–166. [Google Scholar] [CrossRef]
  45. Palcekova, Z.; Angala, S.K.; Belardinelli, J.M.; Eskandarian, H.A.; Joe, M.; Brunton, R.; Rithner, C.; Jones, V.; Nigou, J.; Lowary, T.L.; et al. Disruption of the SucT acyltransferase in Mycobacterium smegmatis abrogates succinylation of cell envelope polysaccharides. J. Biol. Chem. 2019, 294, 10325–10335. [Google Scholar] [CrossRef]
  46. Chatterjee, D.; Lowell, K.; Rivoire, B.; McNeil, M.R.; Brennan, P.J. Lipoarabinomannan of Mycobacterium tuberculosis. Capping with mannosyl residues in some strains. J. Biol. Chem. 1992, 267, 6234–6239. [Google Scholar] [CrossRef]
  47. Chatterjee, D.; Khoo, K.H.; McNeil, M.R.; Dell, A.; Morris, H.R.; Brennan, P.J. Structural definition of the non-reducing termini of mannose-capped LAM from Mycobacterium tuberculosis through selective enzymatic degradation and fast atom bombardment-mass spectrometry. Glycobiology 1993, 3, 497–506. [Google Scholar] [CrossRef]
  48. Khoo, K.-H.; Tang, J.-B.; Chatterjee, D. Variation in mannose-capped terminal arabinan motifs of lipoarabinomannans from clinical isolates of Mycobacterium tuberculosis and Mycobacterium avium complex. J. Biol. Chem. 2001, 276, 3863–3871. [Google Scholar] [CrossRef]
  49. Angala, S.K.; McNeil, M.R.; Shi, L.; Joe, M.; Pham, H.; Zuberogoitia, S.; Nigou, J.; Boot, C.M.; Lowary, T.L.; Gilleron, M.; et al. Biosynthesis of the Methylthioxylose Capping Motif of Lipoarabinomannan in Mycobacterium tuberculosis. ACS Chem. Biol. 2017, 12, 682–691. [Google Scholar] [CrossRef]
  50. Angala, S.K.; Palcekova, Z.; Belardinelli, J.M.; Jackson, M. Covalent modifications of polysaccharides in mycobacteria. Nat. Chem. Biol. 2018, 14, 193–198. [Google Scholar] [CrossRef]
  51. Palcekova, Z.; Obregon-Henao, A.; De, K.; Walz, A.; Lam, H.; Philp, J.; Angala, S.K.; Patterson, J.; Pearce, C.; Zuberogoitia, S.; et al. Role of succinyl substituents in the mannose-capping of lipoarabinomannan and control of inflammation in Mycobacterium tuberculosis infection. PLoS Pathog. 2023, 19, e1011636. [Google Scholar] [CrossRef]
  52. Venisse, A.; Berjeaud, J.M.; Chaurand, P.; Gilleron, M.; Puzo, G. Structural features of lipoarabinomannan from Mycobacterium bovis BCG. Determination of molecular mass by laser desorption mass spectrometry. J. Biol. Chem. 1993, 268, 12401–12411. [Google Scholar] [CrossRef]
  53. Sada, E.; Brennan, P.J.; Herrera, T.; Torres, M. Evaluation of lipoarabinomannan for the serological diagnosis of tuberculosis. J. Clin. Microbiol. 1990, 28, 2587–2590. [Google Scholar] [CrossRef]
  54. Cho, S.N.; Shin, J.S.; Kim, J.D.; Chong, Y. Production of monoclonal antibodies to lipoarabinomannan-B and use in the detection of mycobacterial antigens in sputum. Yonsei Med. J. 1990, 31, 333–338. [Google Scholar] [CrossRef]
  55. Sibley, L.D.; Adams, L.B.; Krahenbuhl, J.L. Inhibition of interferon-gamma-mediated activation in mouse macrophages treated with lipoarabinomannan. Clin. Exp. Immunol. 1990, 80, 141–148. [Google Scholar] [CrossRef]
  56. Sibley, L.D.; Hunter, S.W.; Brennan, P.J.; Krahenbuhl, J.L. Mycobacterial lipoarabinomannan inhibits gamma interferon-mediated activation of macrophages. Infect. Immun. 1988, 56, 1232–1236. [Google Scholar] [CrossRef]
  57. Kaplan, G.; Gandhi, R.R.; Weinstein, D.E.; Levis, W.R.; Patarroyo, M.E.; Brennan, P.J.; Cohn, Z.A. Mycobacterium leprae antigen-induced suppression of T cell proliferation in vitro. J. Immunol. 1987, 138, 3028–3034. [Google Scholar] [CrossRef]
  58. Chan, J.; Fan, X.; Hunter, S.W.; Brennan, P.J.; Bloom, B.R. Lipoarabinomannan, a possible virulence factor involved in persistence of Mycobacterium tuberculosis within macrophages. Infect. Immun. 1991, 59, 1755–1761. [Google Scholar] [CrossRef]
  59. Moreno, C.; Mehlert, A.; Lamb, J. The inhibitory effects of mycobacterial lipoarabinomannan and polysaccharides upon polyclonal and monoclonal human T cell proliferation. Clin. Exp. Immunol. 1988, 74, 206–210. [Google Scholar]
  60. Moreno, C.; Taverne, J.; Mehlert, A.; Bate, C.A.W.; Brealey, R.J.; Meager, A.; Rook, G.A.W.; Playfair, J.H.L. Lipoarabinomannan from Mycobacterium tuberculosis induces the production of tumor necrosis factor from human and murine macrophages. Clin. Exp. Immunol. 1989, 76, 240–245. [Google Scholar]
  61. Barnes, P.F.; Chatterjee, D.; Brennan, P.J.; Rea, T.H.; Modlin, R.L. Tumor necrosis factor production in patients with leprosy. Infect. Immun. 1992, 60, 1441–1446. [Google Scholar] [CrossRef]
  62. Chatterjee, D.; Roberts, A.D.; Lowell, K.; Brennan, P.J.; Orme, I.M. Structural basis of capacity of lipoarabinomannan to induce secretion of tumor necrosis factor. Infect. Immun. 1992, 60, 1249–1253. [Google Scholar] [CrossRef]
  63. Adams, L.B.; Fukutomi, Y.; Krahenbuhl, J.L. Regulation of murine macrophage effector functions by lipoarabinomannan from mycobacterial strains with different degrees of virulence. Infect. Immun. 1993, 61, 4173–4181. [Google Scholar] [CrossRef]
  64. Barnes, P.F.; Chatterjee, D.; Abrams, J.S.; Lu, S.; Wang, E.; Yamamura, M.; Brennan, P.J.; Modlin, R.L. Cytokine production induced by Mycobacterium tuberculosis lipoarabinomannan: Relationship to chemical structure. J. Immunol. 1992, 149, 541–547. [Google Scholar] [CrossRef]
  65. Zhang, Y.; Rom, W.N. Regulation of the interleukin-1b (IL-1b) gene by mycobacterial components and lipopolysaccharide is mediated by two nuclear factor-IL6 motifs. Mol. Cell Biol. 1993, 13, 3831–3837. [Google Scholar]
  66. Maeda, N.; Nigou, J.; Herrmann, J.L.; Jackson, M.; Amara, A.; Lagrange, P.H.; Puzo, G.; Gicquel, B.; Neyrolles, O. The cell surface receptor DC-SIGN discriminates between Mycobacterium species through selective recognition of the mannose caps on lipoarabinomannan. J. Biol. Chem. 2003, 278, 5513–5516. [Google Scholar] [CrossRef]
  67. Schlesinger, L.S.; Hull, S.R.; Kaufman, T.M. Binding of the Terminal Mannosyl Units of Lipoarabinomannan from a Virulent Strain of Mycobacterium Tuberculosis to Human Macrophages. J. Immunol. 1994, 152, 4070–4079. [Google Scholar] [CrossRef]
  68. Venisse, A.; Fournié, J.-J.; Puzo, G. Mannosylated lipoarabinomannan interacts with phagocytes. Eur. J. Biochem. 1995, 231, 440–447. [Google Scholar] [CrossRef]
  69. Fratti, R.A.; Chua, J.; Vergne, I.; Deretic, V. Mycobacterium tuberculosis glycosylated phosphatidylinositol causes phagosome maturation arrest. Proc. Natl. Acad. Sci. USA 2003, 100, 5437–5442. [Google Scholar] [CrossRef]
  70. Schlesinger, L.S.; Kaufman, T.M.; Iyer, S.; Hull, S.R.; Marchiando, L.K. Differences in mannose receptor-mediated uptake of lipoarabinomannan from virulent and attenuated strains of Mycobacterium tuberculosis by human macrophages. J. Immunol. 1996, 157, 4568–4575. [Google Scholar] [CrossRef]
  71. Sieling, P.A.; Chatterjee, D.; Porcelli, S.A.; Prigozy, T.I.; Mazzaccaro, R.J.; Soriano, T.; Bloom, B.R.; Brenner, M.B.; Kronenberg, M.; Brennan, P.J.; et al. CD1-restricted T cell recognition of microbial lipoglycan antigens. Science 1995, 269, 227–230. [Google Scholar] [CrossRef]
  72. Prigozy, T.I.; Sieling, P.A.; Clemens, D.; Stewart, P.L.; Behar, S.M.; Porcelli, S.A.; Brenner, M.B.; Modlin, R.L.; Kronenberg, M. The mannose receptor delivers lipoglycan antigens to endosomes for presentation to T cells by CD1b molecules. Immunity 1997, 6, 187–197. [Google Scholar] [CrossRef]
  73. Kallert, S.; Zenk, S.F.; Walther, P.; Grieshober, M.; Weil, T.; Stenger, S. Liposomal delivery of lipoarabinomannan triggers Mycobacterium tuberculosis specific T-cells. Tuberculosis 2015, 95, 452–462. [Google Scholar] [CrossRef]
  74. Kaur, D.; Obregon-Henao, A.; Pham, H.; Chatterjee, D.; Brennan, P.J.; Jackson, M. Lipoarabinomannan of Mycobacterium: Mannose capping by a multifunctional terminal mannosyltransferase. Proc. Natl. Acad. Sci. USA 2008, 105, 17973–17977. [Google Scholar] [CrossRef]
  75. Alderwick, L.J.; Birch, H.L.; Mishra, A.K.; Eggeling, L.; Besra, G.S. Structure, function and biosynthesis of the Mycobacterium tuberculosis cell wall: Arabinogalactan and lipoarabinomannan assembly with a view to discovering new drug targets. Biochem. Soc. Trans. 2007, 35, 1325–1328. [Google Scholar] [CrossRef] [PubMed]
  76. Chatterjee, D.; Brennan, P.J. Glycosylated components of the mycobacterial cell wall;structure and function. In Microbial Glycobiology: Structures, Relevance and Applications, 1st ed.; Holst, O., Brennan, P.J., Itzstein, V.M., Eds.; Academic Press: Oxford, UK, 2009; pp. 147–167. [Google Scholar]
  77. Briken, V.; Porcelli, S.A.; Besra, G.S.; Kremer, L. Mycobacterial lipoarabinomannan and related lipoglycans: From biogenesis to modulation of the immune response. Mol. Microbiol. 2004, 53, 391–403. [Google Scholar] [CrossRef] [PubMed]
  78. Rivoire, B.; Ranchoff, B.; Chatterjee, D.; Gaylord, H.; Tsang, A.; Kolk, A.H.J.; Aspinall, G.O.; Brennan, P.J. Generation of monoclonal antibodies to the specific sugar epitopes of Mycobacterium avium complex serovars. Infect. Immun. 1989, 57, 3147–3158. [Google Scholar] [CrossRef]
  79. Ahmad, R.; Xie, L.; Pyle, M.; Suarez, M.F.; Broger, T.; Steinberg, D.; Ame, S.M.; Lucero, M.G.; Szucs, M.J.; MacMullan, M.; et al. A rapid triage test for active pulmonary tuberculosis in adult patients with persistent cough. Sci. Transl. Med. 2019, 11, eaaw8287. [Google Scholar] [CrossRef]
  80. Choudhary, A.; Patel, D.; Honnen, W.; Lai, Z.; Prattipati, R.S.; Zheng, R.B.; Hsueh, Y.C.; Gennaro, M.L.; Lardizabal, A.; Restrepo, B.I.; et al. Characterization of the Antigenic Heterogeneity of Lipoarabinomannan, the Major Surface Glycolipid of Mycobacterium tuberculosis, and Complexity of Antibody Specificities toward this Antigen. J. Immunol. 2018, 200, 3053–3066. [Google Scholar] [CrossRef] [PubMed]
  81. Corrigan, D.T.; Ishida, E.; Chatterjee, D.; Lowary, T.L.; Achkar, J.M. Monoclonal antibodies to lipoarabinomannan/arabinomannan—Characteristics and implications for tuberculosis research and diagnostics. Trends Microbiol. 2023, 31, 22–35. [Google Scholar] [CrossRef]
  82. Lawn, S.D.; Gupta-Wright, A. Detection of lipoarabinomannan (LAM) in urine is indicative of disseminated TB with renal involvement in patients living with HIV and advanced immunodeficiency: Evidence and implications. Trans. R. Soc. Trop. Med. Hyg. 2016, 110, 180–185. [Google Scholar] [CrossRef] [PubMed]
  83. Gupta-Wright, A.; Peters, J.A.; Flach, C.; Lawn, S.D. Detection of lipoarabinomannan (LAM) in urine is an independent predictor of mortality risk in patients receiving treatment for HIV-associated tuberculosis in sub-Saharan Africa: A systematic review and meta-analysis. BMC Med. 2016, 14, 53. [Google Scholar] [CrossRef]
  84. Shah, M.; Hanrahan, C.; Wang, Z.Y.; Dendukuri, N.; Lawn, S.D.; Denkinger, C.M.; Steingart, K.R. Lateral flow urine lipoarabinomannan assay for detecting active tuberculosis in HIV-positive adults. Cochrane Database Syst. Rev. 2016, 5, CD011420. [Google Scholar] [CrossRef] [PubMed]
  85. Amin, A.G.; De, P.; Graham, B.; Calderon, R.I.; Franke, M.F.; Chatterjee, D. Urine lipoarabinomannan in HIV uninfected, smear negative, symptomatic TB patients: Effective sample pretreatment for a sensitive immunoassay and mass spectrometry. Sci. Rep. 2021, 11, 2922. [Google Scholar] [CrossRef] [PubMed]
  86. Amin, A.G.; De, P.; Graham, B.; Jensen, B.L.; Moreau, E.; Chatterjee, D. Overcome low levels of detection limit and choice of antibody affects detection of lipoarabinomannan in pediatric tuberculosis. PLoS ONE 2022, 17, e0275838. [Google Scholar] [CrossRef]
  87. Sigal, G.B.; Pinter, A.; Lowary, T.L.; Kawasaki, M.; Li, A.; Mathew, A.; Tsionsky, M.; Zheng, R.B.; Plisova, T.; Shen, K.; et al. A Novel Sensitive Immunoassay Targeting the 5-Methylthio-d-Xylofuranose-Lipoarabinomannan Epitope Meets the WHO’s Performance Target for Tuberculosis Diagnosis. J. Clin. Microbiol. 2018, 56, e01338-18. [Google Scholar] [CrossRef] [PubMed]
  88. Zhang, A.; Jumbe, E.; Krysiak, R.; Sidiki, S.; Kelley, H.V.; Chemey, E.K.; Kamba, C.; Mwapasa, V.; Garcia, J.I.; Norris, A.; et al. Low-cost diagnostic test for susceptible and drug-resistant tuberculosis in rural Malawi. Afr. J. Lab. Med. 2018, 7, 690. [Google Scholar] [CrossRef]
  89. Lawn, S.D.; Kerkhoff, A.D.; Nicol, M.P.; Meintjes, G. Underestimation of the true specificity of the urine lipoarabinomannan (LAM) point-of-care diagnostic assay for HIV-associated tuberculosis. J. Acquir. Immune Defic. Syndr. 2015, 69, e144. [Google Scholar] [CrossRef]
  90. Paris, L.; Magni, R.; Zaidi, F.; Araujo, R.; Saini, N.; Harpole, M.; Coronel, J.; Kirwan, D.E.; Steinberg, H.; Gilman, R.H.; et al. Urine lipoarabinomannan glycan in HIV-negative patients with pulmonary tuberculosis correlates with disease severity. Sci. Transl. Med. 2017, 9, eaal2807. [Google Scholar] [CrossRef]
  91. Nabeemeeah, F.; Sabet, R.; Moloantoa, T.; Waja, Z.; Pretorius, Z.; Majoro, K.; Letutu-Xaba, M.; Vilaplana, C.; Nigou, J.; Martinson, N. Exhaled breath specimens subjected to point-of-care lipoarabinomannan testing. Int. J. Tuberc. Lung Dis. 2023, 27, 703–705. [Google Scholar] [CrossRef]
  92. Mosquera-Restrepo, S.F.; Zuberogoïtia, S.; Gouxette, L.; Layre, E.; Gilleron, M.; Stella, A.; Rengel, D.; Burlet-Schiltz, O.; Caro, A.C.; Garcia, L.; et al. A Mycobacterium tuberculosis fingerprint in human breath allows tuberculosis detection. Nat. Commun. 2022, 13, 7751. [Google Scholar] [CrossRef]
  93. Broger, T.; Tsionksy, M.; Mathew, A.; Lowary, T.L.; Pinter, A.; Plisova, T.; Bartlett, D.; Barbero, S.; Denkinger, C.M.; Moreau, E.; et al. Sensitive electrochemiluminescence (ECL) immunoassays for detecting lipoarabinomannan (LAM) and ESAT-6 in urine and serum from tuberculosis patients. PLoS ONE 2019, 14, e0215443. [Google Scholar] [CrossRef]
  94. De, P.; Amin, A.G.; Graham, B.; Martiniano, S.L.; Caceres, S.M.; Poch, K.R.; Jones, M.C.; Saavedra, M.T.; Malcolm, K.C.; Nick, J.A.; et al. Urine lipoarabinomannan as a marker for low-risk of NTM infection in the CF airway. J. Cyst. Fibros. 2020, 19, 801–807. [Google Scholar] [CrossRef]
  95. Qvist, T.; Johansen, I.S.; Pressler, T.; Hoiby, N.; Andersen, A.B.; Katzenstein, T.L.; Bjerrum, S. Urine lipoarabinomannan point-of-care testing in patients affected by pulmonary nontuberculous mycobacteria--experiences from the Danish Cystic Fibrosis cohort study. BMC Infect. Dis. 2014, 14, 655. [Google Scholar] [CrossRef]
  96. Nick, J.A.; Dedrick, R.M.; Gray, A.L.; Vladar, E.K.; Smith, B.E.; Freeman, K.G.; Malcolm, K.C.; Epperson, L.E.; Hasan, N.A.; Hendrix, J.; et al. Host and pathogen response to bacteriophage engineered against Mycobacterium abscessus lung infection. Cell 2022, 185, 1860–1874.e1812. [Google Scholar] [CrossRef]
  97. Guerardel, Y.; Maes, E.; Briken, V.; Chirat, F.; Leroy, Y.; Locht, C.; Strecker, G.; Kremer, L. Lipomannan and lipoarabinomannan from a clinical isolate of Mycobacterium kansasii: Novel structural features and apoptosis-inducing properties. J. Biol. Chem. 2003, 278, 36637–36651. [Google Scholar] [CrossRef] [PubMed]
  98. Glass, L.N.; Swapna, G.; Chavadi, S.S.; Tufariello, J.M.; Mi, K.; Drumm, J.E.; Lam, T.T.; Zhu, G.; Zhan, C.; Vilcheze, C.; et al. Mycobacterium tuberculosis universal stress protein Rv2623 interacts with the putative ATP binding cassette (ABC) transporter Rv1747 to regulate mycobacterial growth. PLoS Pathog. 2017, 13, e1006515. [Google Scholar] [CrossRef] [PubMed]
  99. Coscolla, M.; Gagneux, S. Consequences of genomic diversity in Mycobacterium tuberculosis. Semin. Immunol. 2014, 26, 431–444. [Google Scholar] [CrossRef] [PubMed]
  100. Gagneux, S.; DeRiemer, K.; Van, T.; Kato-Maeda, M.; de Jong, B.C.; Narayanan, S.; Nicol, M.; Niemann, S.; Kremer, K.; Gutierrez, M.C.; et al. Variable host-pathogen compatibility in Mycobacterium tuberculosis. Proc. Natl. Acad. Sci. USA 2006, 103, 2869–2873. [Google Scholar] [CrossRef]
  101. Mikusova, K.; Huang, H.; Yagi, T.; Holsters, M.; Vereecke, D.; D’Haeze, W.; Scherman, M.S.; Brennan, P.J.; McNeil, M.R.; Crick, D.C. Decaprenylphosphoryl arabinofuranose, the donor of the D-arabinofuranosyl residues of mycobacterial arabinan, is formed via a two-step epimerization of decaprenylphosphoryl ribose. J. Bacteriol. 2005, 187, 8020–8025. [Google Scholar] [CrossRef]
  102. Brock, M.; Hanlon, D.; Zhao, M.; Pollock, N.R. Detection of mycobacterial lipoarabinomannan in serum for diagnosis of active tuberculosis. Diagn. Microbiol. Infect. Dis. 2020, 96, 114937. [Google Scholar] [CrossRef]
  103. Arcos, J.; Sasindran, S.J.; Fujiwara, N.; Turner, J.; Schlesinger, L.S.; Torrelles, J.B. Human lung hydrolases delineate Mycobacterium tuberculosis-macrophage interactions and the capacity to control infection. J. Immunol. 2011, 187, 372–381. [Google Scholar] [CrossRef]
Figure 1. Schematic Representation of the evolution of Mtb LAM structural studies. Approximate timeline for last 100 years starting in 1925.
Figure 1. Schematic Representation of the evolution of Mtb LAM structural studies. Approximate timeline for last 100 years starting in 1925.
Pathogens 12 01281 g001
Figure 2. Schematic representation of ManLAM from Mtb. The antiLAM monoclonal antibodies (mAbs) CS-35, A194-01, and CHCS9-08 are shown to react with the LAM arabinan terminal arrangements (Ara4 and Ara6, respectively). These three mAbs are widely used in immunoassays for TB diagnosis. The cartoon is based on a screening of 12 synthetic arabinan glycoconjugates by Dr. Todd Lowary [40]. In order to identify which motifs of LAM are being recognized by anti-LAM antibodies, an exhaustive digestion of LAM with commercially available α-mannosidase (derived from Jack Beans) is performed. This digestion removes the mannose caps of LAM (depicted as grey hexagons). An additional digestion using an endo-arabinanase (in-house isolated from Cellulomonas) releases arabinan fragments from the D-arabinan domain, mainly Ara4, Ara5, and Ara6 fragments. These enzymatic digestions allowed the identification of the Ara4 and Ara6 motifs as the ones being recognized by monoclonal antibodies against LAM and were used in the development of TB diagnosis downstream.
Figure 2. Schematic representation of ManLAM from Mtb. The antiLAM monoclonal antibodies (mAbs) CS-35, A194-01, and CHCS9-08 are shown to react with the LAM arabinan terminal arrangements (Ara4 and Ara6, respectively). These three mAbs are widely used in immunoassays for TB diagnosis. The cartoon is based on a screening of 12 synthetic arabinan glycoconjugates by Dr. Todd Lowary [40]. In order to identify which motifs of LAM are being recognized by anti-LAM antibodies, an exhaustive digestion of LAM with commercially available α-mannosidase (derived from Jack Beans) is performed. This digestion removes the mannose caps of LAM (depicted as grey hexagons). An additional digestion using an endo-arabinanase (in-house isolated from Cellulomonas) releases arabinan fragments from the D-arabinan domain, mainly Ara4, Ara5, and Ara6 fragments. These enzymatic digestions allowed the identification of the Ara4 and Ara6 motifs as the ones being recognized by monoclonal antibodies against LAM and were used in the development of TB diagnosis downstream.
Pathogens 12 01281 g002
Figure 3. Dominant arabinan termini of Mtb lipoarabinomannan from culture. Only non-reducing ends are shown. The terminal structures evolved after extensive enzymatic degradation of LAM followed by liquid chromatography with tandem mass spectrometry (LC/MS-MS) analyses [41]. Abbreviations: Ara = D-Arabiniose, Man = D-Mannose, Suc = Succinates.
Figure 3. Dominant arabinan termini of Mtb lipoarabinomannan from culture. Only non-reducing ends are shown. The terminal structures evolved after extensive enzymatic degradation of LAM followed by liquid chromatography with tandem mass spectrometry (LC/MS-MS) analyses [41]. Abbreviations: Ara = D-Arabiniose, Man = D-Mannose, Suc = Succinates.
Pathogens 12 01281 g003
Figure 4. MALDI-TOF analysis of LAM from Mtb. The mass spectrometry (MS) was done on Bruker ultrafleXtreme matrix-assisted laser desorption time-of-flight spectrometry (MALDI-TOF/TOF MS), indicating a molecular mass for ManLAM of approximately 15 kDa. Heterogeneity in mass of 162 m/z corresponds to a hexose and 100 m/z corresponds to succinates.
Figure 4. MALDI-TOF analysis of LAM from Mtb. The mass spectrometry (MS) was done on Bruker ultrafleXtreme matrix-assisted laser desorption time-of-flight spectrometry (MALDI-TOF/TOF MS), indicating a molecular mass for ManLAM of approximately 15 kDa. Heterogeneity in mass of 162 m/z corresponds to a hexose and 100 m/z corresponds to succinates.
Pathogens 12 01281 g004
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

Torrelles, J.B.; Chatterjee, D. Collected Thoughts on Mycobacterial Lipoarabinomannan, a Cell Envelope Lipoglycan. Pathogens 2023, 12, 1281. https://doi.org/10.3390/pathogens12111281

AMA Style

Torrelles JB, Chatterjee D. Collected Thoughts on Mycobacterial Lipoarabinomannan, a Cell Envelope Lipoglycan. Pathogens. 2023; 12(11):1281. https://doi.org/10.3390/pathogens12111281

Chicago/Turabian Style

Torrelles, Jordi B., and Delphi Chatterjee. 2023. "Collected Thoughts on Mycobacterial Lipoarabinomannan, a Cell Envelope Lipoglycan" Pathogens 12, no. 11: 1281. https://doi.org/10.3390/pathogens12111281

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