Next Article in Journal
Molecular Frontiers in Melanoma: Pathogenesis, Diagnosis, and Therapeutic Advances
Next Article in Special Issue
The Response of Growth and Transcriptome Profiles of Tea Grey Blight Disease Pathogen Pestalotiopsis theae to the Variation of Exogenous L-Theanine
Previous Article in Journal
The Effects of Body Fat Reduction through the Metabolic Control of Steam-Processed Ginger Extract in High-Fat-Diet-Fed Mice
Previous Article in Special Issue
Selection and Validation of Reference Genes for Gene Expression in Bactericera gobica Loginova under Different Insecticide Stresses
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Bioinformatic Analysis of Sulfotransferases from an Unexplored Gut Microbe, Sutterella wadsworthensis 3_1_45B: Possible Roles towards Detoxification via Sulfonation by Members of the Human Gut Microbiome

1
Biodesign Center for Fundamental and Applied Microbiomics, Arizona State University, Tempe, AZ 85281, USA
2
School of Mathematical and Natural Sciences, Arizona State University, Glendale, AZ 85306, USA
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2024, 25(5), 2983; https://doi.org/10.3390/ijms25052983
Submission received: 19 January 2024 / Revised: 27 February 2024 / Accepted: 28 February 2024 / Published: 4 March 2024
(This article belongs to the Collection Feature Papers in Molecular Microbiology)

Abstract

:
Sulfonation, primarily facilitated by sulfotransferases, plays a crucial role in the detoxification pathways of endogenous substances and xenobiotics, promoting metabolism and elimination. Traditionally, this bioconversion has been attributed to a family of human cytosolic sulfotransferases (hSULTs) known for their high sequence similarity and dependence on 3′-phosphoadenosine 5′-phosphosulfate (PAPS) as a sulfo donor. However, recent studies have revealed the presence of PAPS-dependent sulfotransferases within gut commensals, indicating that the gut microbiome may harbor a diverse array of sulfotransferase enzymes and contribute to detoxification processes via sulfation. In this study, we investigated the prevalence of sulfotransferases in members of the human gut microbiome. Interestingly, we stumbled upon PAPS-independent sulfotransferases, known as aryl-sulfate sulfotransferases (ASSTs). Our bioinformatics analyses revealed that members of the gut microbial genus Sutterella harbor multiple asst genes, possibly encoding multiple ASST enzymes within its members. Fluctuations in the microbes of the genus Sutterella have been associated with various health conditions. For this reason, we characterized 17 different ASSTs from Sutterella wadsworthensis 3_1_45B. Our findings reveal that SwASSTs share similarities with E. coli ASST but also exhibit significant structural variations and sequence diversity. These differences might drive potential functional diversification and likely reflect an evolutionary divergence from their PAPS-dependent counterparts.

Graphical Abstract

1. Introduction

Sulfonation is a key detoxification pathway, where endogenous compounds and foreign substances undergo a transformation to increase their water solubility, aiding in metabolism and elimination [1]. This process has traditionally been attributed to a diverse array of human cytosolic sulfotransferases [2,3,4,5,6,7,8,9,10,11,12,13,14,15,16,17,18,19,20]. However, recent discoveries have identified enzymes capable of cholesterol sulfonation within the Bacteroides genus, a prominent member of the human gut microbiota [21,22]. This suggests the potential presence of additional sulfotransferases in the gut microbiome. While microbial sulfotransferases have been previously studied, they lack the extensive characterization seen with human enzymes [23,24,25,26]. Sulfotransferases generally facilitate a biochemical reaction in which a sulfo group is transferred from a donor to an acceptor molecule (Figure 1).
Extensive research has been conducted on human sulfotransferases. To date, 14 cytosolic human sulfotransferases (hSULTs) have been identified and characterized (Table S1) [2,3,4,5,6,7,8,9,10,11,12,13,14,15,16,17,18,19,20]. These enzymes exhibit a high degree of sequence similarity, and all utilize 3′-phosphoadenosine 5′-phosphosulfate (PAPS) as the common sulfo donor. However, their sulfo acceptor preferences differ, giving each hSULT a unique function [27]. This variability is mostly present in the substrate-binding regions of the enzymes [28]. Specifically, the substrate binding loops of hSULTs demonstrate significant diversity [27]. The 14 distinct isoforms of human sulfotransferases (hSULTs) have tissue-specific expression and are responsible for the sulfonation of numerous small molecules, both endogenous and exogenous [28]. These enzymes are typically characterized by their dependency on PAPS for the sulfotransferase activity and are recognized by the presence of PAPS binding motifs [29]. In addition, hSULTs are categorized under the large class of aryl sulfotransferases (EC 2.8.2.1) [26]. Human cytosolic sulfotransferases facilitate sulfo transfer reactions, utilizing either a nonsequential (random Bi Bi) mechanism, where the substrates bind independently without intermediate formation, or a sequential (ordered Bi Bi) mechanism, where the binding of one substrate facilitates the attachment of the other [20,30,31]. PAPS-dependent aryl sulfotransferases are not exclusive to humans but are also found across various eukaryotic and prokaryotic species [7,9,11,17,21,22]. The widespread and prevalent nature of these enzymes is not fully understood. One hypothesis is that these enzymes might be involved in detoxification of molecules that are present at varying levels in the environments of these diverse life forms [30].
In the late 1980s, a series of studies demonstrated that the intestinal flora exhibited sulfation activity on small phenolic compounds [23,24,26]. These investigations revealed a new class of previously unknown microbial sulfotransferases, distinctive in its utilization of sulfo donors other than the mammalian default, 3′-phosphoadenosine 5′-phosphosulfate (PAPS). Interestingly, these microbial enzymes were capable of utilizing a variety of nonphysiological phenolic sulfate esters as donors, with p-nitrophenol sulfate (pNPS) being predominantly used.
Due to the differences in donor specificities, reaction mechanisms, and kinetic profiles, these enzymes were designated as aryl-sulfate sulfotransferases (ASSTs), classified under EC 2.8.2.22 [32]. Beak et al. conducted studies on the distribution of ASSTs, showing their presence across both gram-positive and gram-negative bacterial species [33]. Furthermore, ASSTs were observed to exist both with and without signal peptides [33]. Cumulatively, these studies show that there are a many variations in microbial ASSTs. The widespread prevalence of ASSTs and the current gap in understanding their functional importance demonstrate the need for further study.
The integral role of gut microbiota in metabolizing a range of xenobiotics and endogenous substances is already well-recognized [34]. As mentioned earlier, a common route for the metabolism of these compounds is via sulfonation. Motivated by these findings, we have delved into the genomic landscape of gut microbes, specifically searching for genes annotated as aryl-sulfate sulfotransferases. Interestingly, we discovered that Sutterella, a genus within the human gut microbiome, contains various members that harbor multiple genes annotated for these sulfotransferases. It is very unusual to find more than two annotated metabolic genes predicted to produce the same type of protein.
Sutterella species have been isolated from human fecal samples [35,36,37,38,39,40,41] and implicated in various human health conditions, including ulcerative colitis [42], autism spectrum disorders (ASD) [43,44], bacteremia [45], multiple sclerosis [46], and autoimmune based thyroid disease [47]. Additionally, Sutterella is known to degrade immunoglobulin A (IgA), suggesting a pro-inflammatory role [48]. Interestingly, in the context of ulcerative colitis, a decline in Sutterella wadsworthensis has been associated with drug-free remission [42]. Given its widespread presence and links to several human health concerns, we decided to investigate and understand the presence of annotated asst genes and their products (ASST proteins) in the genome of S. wadsworthensis. Our bioinformatics study has revealed a predominant presence of multiple aryl-sulfate sulfotransferase (ASST) enzymes harboring members within the genus Sutterella. These enzymes exhibit considerable sequence and structural homogeneity, albeit with enough variability that can have functional consequences. Additionally, while sharing similarities with prokaryotic ASSTs, such as those in E. coli, Sutterella’s sulfotransferases appear to have diverged evolutionarily from the PAPS-dependent class of sulfotransferases.

2. Results and Discussion

2.1. Distribution of Aryl-Sulfate Sulfotransferase (asst) Genes in Human Gut Microbiota

Aryl-sulfate sulfotransferases (ASSTs) are enzymes found extensively across various genera of gut microbiota. Figure 2 illustrates the prevalence of annotated asst genes within diverse gut microbial genera. The data depict the presence of these genes across different species and strains within each genus. The analysis omits numerous E. coli strains due to their high abundance and to avoid skewing the data representation.
Our research indicates a broad distribution of asst genes among gut microbes. Interestingly, a majority of these microbes possess only one or two annotated asst genes per organism. Genera such as Lactobacillus, Bifidobacterium, Enterobacter, and Citrobacter are predominant, collectively harboring 60% of all asst gene annotations identified in gut microbes. Despite the apparent ubiquity of these genes, the underlying reasons for their prevalence in human gut microbiota remain elusive.

2.2. Variability of Annotated Aryl-Sulfate Sulfotransferase (asst) Genes in the Genus Sutterella

Our search through IMG/M database has uncovered multiple annotated aryl-sulfate sulfotransferases within the genus Sutterella. The bar graph in Figure 3 and Table S2 details the distribution of these genes among different species and strains of Sutterella. Notably, Sutterella wadsworthensis exhibits the largest number of annotated asst genes, with individual strains showing variability ranging from 8 to 17 genes.
Furthermore, Sutterella megalosphaeroides and Sutterella sp. KLE1602 each harbor a total of 11 annotated asst genes. Close behind, Sutterella massiliensis has 10 such genes. Distinct strains of Sutterella seckii, ASD393 and ASD3426, comprise 10 and 7 annotated asst genes, respectively. Other species, including Sutterella faecalis, Sutterella sp. KCTC 15688, and Sutterella parvirubra, display between 5 and 7 genes. Contrastingly, several strains, such as AM11-39, AF15-44LB, AF15-45LB, and AM18-8-1 of Sutterella sp., contain only 1 to 2 annotated asst genes. However, species like Sutterella morbirenis, Sutterella sanguinus, Sutterella stercoricanis, and Sutterella timonensis have no detectable annotated asst genes.
This variation within the same genus raises intriguing questions: Why do certain Sutterella species have a multitude of asst genes while others have none? Addressing these questions will be possible only after the development of advanced genetic manipulation tools specifically tailored for the Sutterella genus.

2.3. Bioinformatic Analysis of Annotated asst Genes in S. wadsworthensis 3_1_45B

Given that S. wadsworthensis has the highest count of annotated genes encoding aryl-sulfate sulfotransferases (assts), we focused on a detailed characterization of these genes and their corresponding enzyme products from the strain S. wadsworthensis 3_1_45B, which contains 17 identified asst genes. The comprehensive list and characteristics of these 17 genes and enzymes can be found in Table S3. For asst genes, we examined the percent GC content and the gene neighborhoods.

2.4. GC Content Variation in asst Genes

We first analyzed the percent GC content of the asst genes from S. wadsworthensis 3_1_45B, with data sourced from the IMG/M database. By plotting the GC content, we aimed to determine any variations or similarities that might exist within asst genes. The GC content of various S. wadsworthensis strains is generally around 55% [45]. The mean GC content of the asst genes was found to be 55.3%, with a median of 55%, aligning closely with the previously reported values for a GC content of S. wadsworthensis [45]. However, the GC content of the 17 annotated asst genes did display some diversity, ranging from a minimum of 49% to a maximum of 61% (Figure 4). It has been seen that the overall base composition of highly expressed genes can affect bacterial fitness. Additionally, mRNAs with higher GC content display higher stability [49]. GC content can also affect gene expression and regulation. This could also help organism to adapt to various environmental conditions. Thus, GC content variation in the annotated asst genes might have physiological relevance.

2.5. Analysis of Gene Neighborhoods for asst Genes in S. wadsworthensis 3_1_45B

Investigating the gene neighborhood is critical for understanding the physiological roles of specific genes and identifying potential operon structures, which may elucidate their functions. Interestingly, in E. coli, the asst gene is part of an operon associated with the formation of disulfide bonds [29]. Accordingly, such analyses are important in shedding light on the functions of the annotated sulfotransferases. Figure 5 illustrates the comprehensive analysis of gene neighborhoods of asst genes in S. wadsworthensis 3_1_45B.
An analysis of the seventeen annotated asst genes has revealed that seven of these have the lysR gene in close vicinity (Figure 5 and Table 1). The lysR gene encodes LysR family transcription regulator. LysR-type transcriptional regulators (LTTRs) play dual roles as activators or repressors of gene expression [50]. In E. coli, for instance, the LysR-type regulator CysB governs the expression of genes crucial for sulfate assimilation and organic sulfur utilization [51]. This regulatory control extends to sulfate starvation responses in P. putida as well [52]. Furthermore, recent discoveries have identified a novel LysR type transcriptional regulator (LTTR) from A. baumannii, which is involved in controlling the expression of genes involved in the uptake and reduction of various sulfur compounds [53]. LTTRs are prolific within prokaryotes and are known to regulate a multitude of bacterial functions, such as stress response, antibiotic resistance, motility, quorum sensing, degradation of aromatic compounds, and biosynthesis of amino acids [54].
Given that the ASST proteins are predicated to catalyze sulfo transfer reactions, it is plausible that LysR may influence the expression of ASSTs. Our findings indicate that lysR is a predominant gene in the vicinity of genes encoding SwASST proteins. Other genes located near assts are listed in Table 1, providing a comprehensive view of the gene neighborhood.
In the vicinity of asst genes, particularly those encoding SwASSTs 1, 5, and 13, we frequently identified the gene for the fumarate reductase flavoprotein subunit. This protein is integral to the process of anaerobic respiration [55], where sulfate can serve as one of the terminal electron acceptors. In addition, certain gut microbes are known to use naturally occurring organosulfonates, such as taurine, as terminal electron acceptors [56]. Given that aryl-sulfate sulfotransferases (ASSTs) are involved in the production of organosulfonates through the transfer of sulfates between organic molecules [23,24,26], it raises the question of whether fumarate reductase could facilitate the transfer of electrons to these organosulfonates. Could there be a coordinated expression between ASSTs and fumarate reductase, suggesting a metabolic linkage? This presents an interesting hypothesis, which requires experimental validation. Beyond the two genes commonly found near various asst genes, our analysis has identified additional genes situated in close proximity to single and specific asst genes encoding SwASSTs.
For instance, in the gene neighborhood of asst1, there is a gene that encodes a putative zinc metalloprotease. These enzymes are ubiquitous and versatile, often associated with pathogenicity and virulence in pathogenic microbes, though they are also present in nonpathogenic species. The presence of zinc metalloprotease is speculated to confer an evolutionary advantage to the microorganisms that retain it [57]. However, the connection between an ASST protein and zinc metalloprotease is unclear.
In the vicinity of the asst2 gene, there is a gene that encodes a putative ATP-binding cassette transporter (ABC transporter). ABC transporters, present in both eukaryotes and prokaryotes, perform a wide range of functions. Among their roles in microbial organisms is the transportation of sulfate, sulfonates, and sulfate esters [58,59]. Further research may reveal whether this ABC transporter is responsible for the translocation of the products generated by the activity of SwASST2.
Similar to the LysR family, the LuxR family transcriptional regulators can also function as either activators or repressors of gene expression, predominantly managing genes associated with quorum sensing. However, their regulatory influence extends beyond quorum sensing to other critical microbial biological functions [60]. One such case is a novel LuxR-type regulator involved in catechol metabolism within the prominent gut microbe, E. lenta [61]. A gene for LuxR family transcriptional regulator is also found adjacent to the asst3 gene, which encodes SwASST3. Whether it serves as a regulator for the expression of the asst gene, potentially influencing the metabolism of sulfated molecules in Sutterella spp., is a question that remains open for investigation.
Adjacent to the asst3 encoding SwASST3 are two other genes in addition to the gene that produces LuxR regulator (Table 1). The first is an annotated gene that encodes molybdate ABC transporter. Previous studies have demonstrated the competitive interaction between sulfate ions and molybdate for uptake via similar transport systems [62,63], an observation that has been made in the small intestine of sheep [64]. This raises the question: might there be a similar competitive mechanism affecting the transport of the sulfated products created by SwASST3 and molybdate?
The other gene in proximity to asst3 encodes adenylosuccinate lyase, an enzyme implicated in purine biosynthesis via the de novo pathway. Its enzymatic activity is essential for the synthesis of AMP [65]. Any possible connection between an ASST protein and an adenylosuccinate lyase has not been explored before.
Surrounding the gene encoding SwASST4, two notable genes are present (Table 1). The gene for ornithine carbamoyltransferase, which modulates ornithine levels, is one such gene. Microbes capable of breaking down ornithine demonstrate a competitive advantage [66]. This raises the question: could this enzyme also confer a survival benefit to Sutterella wadsworthensis in a healthy gut, and might SwASST4 play a role in this dynamic? Further investigation is needed to explore these possibilities. An additional gene adjacent to asst4 is a gene for oligopeptidase A, part of the M03A family of metallopeptidases, known for their ability to cleave small peptides starting with alanine or glycine [67]. There are no reports of any interaction between an oligopeptidase A and an ASST.
In the vicinity of the gene (asst6) encoding SwASST6, lies a gene for a LysR family transcription regulator, previously discussed (Table 1). Another neighboring gene encodes a carbonic anhydrase, pivotal for maintaining CO2, HCO3, and H+ balance within microbes and facilitating crucial exchange of these molecules with various metabolic pathways. Gut microbial carbonic anhydrases significantly differ from those in the host [68], suggesting that their activity levels can impact the survival of gut microbiota [69].
The gene neighborhood of asst8 includes an ATP-dependent Clp protease gene (Table 1). Proteases like these are essential for the removal of damaged or misfolded proteins. Because of their critical role in protein degradation pathways, these proteins are imperative for microbial physiology [70].
The gene encoding SwASST9 is associated with two distinct genes within its local genomic landscape (Table 1). In addition to the previously noted LysR family transcription regulator, this neighborhood also includes a gene for the outer membrane transport protein, TonB. TonB-dependent transporters are known to leverage the proton motive force to enable the translocation of substances across the outer membrane. Specifically, TonB plays a critical role in importing nutrients, particularly large polysaccharides derived from the diet, into gut microbes. The considerable size of these molecules necessitates an energy-dependent mechanism for membrane passage, a process facilitated by TonB-dependent transporters [71,72].
In addition to the above mentioned lysR family transcription regulator gene, three other annotated genes surrounding asst10 gene that encodes SwASST10 (Table 1). These genes are for translation elongation factor P, deoxycytidylate deaminase, and an amidohydrolase. The translation elongation factor P is instrumental in microbial protein synthesis, particularly for polypeptides comprising polyproline sequences [73]. The deoxycytidylate deaminase plays a crucial role in the synthesis of the thymidine nucleotide, a building block of DNA [74]. The amidohydrolase, a versatile enzyme, is capable of cleaving a variety of bonds, such as carbon–oxygen, phosphorous–oxygen, phosphorous–sulfur, carbon–nitrogen, carbon–sulfur, and carbon–chlorine, with known involvement in the metabolism of xenobiotics [75].
Adjacent to the gene encoding SwASST11, pyruvate carboxylase has been identified, as listed in Table 1. This enzyme plays a pivotal role in anaplerosis, the process of replenishing intermediates of the tricarboxylic acid (TCA) cycle that are consumed by various biosynthetic pathways. Pyruvate carboxylase specifically catalyzes the formation of oxaloacetate, a key intermediate in the TCA cycle. Moreover, carboxylases are integral to carbon fixation reactions, facilitating the incorporation of inorganic carbon into organic molecules [76].
Surrounding the gene encoding SwASST14, three significant genes have been annotated: histidinol dehydrogenase, ATP phosphoribosyltransferase, and a phosphate:Na+ symporter, as noted in Table 1 and Figure 5. The biosynthesis of histidine is a well-conserved pathway found across various life forms, with the exception of mammals [77]. ATP phosphoribosyltransferase and histidinol dehydrogenase are key enzymes in this pathway, representing the first and the final steps, respectively [78]. ATP phosphoribosyltransferase initiates the pathway by condensing 1-(5-phospho-D-ribosyl)-ATP with a diphosphate molecule. The phosphate:Na+ symporter may facilitate the uptake of phosphate, the substrate required for this first reaction in histidine biosynthesis. The potential involvement of SwASST14 in histidine biosynthesis or its regulation poses an intriguing question for future research endeavors.
Located in the genomic vicinity of the gene encoding SwASST15 is a cadmium-translocating P-type ATPase (Table 1). P-type ATPases play a critical role in cellular homeostasis by regulating the intracellular concentrations of both essential and toxic transition metals. They achieve this by actively expelling toxic metals from the cells, thereby mitigating metal toxicity [79].
Lastly, there are two annotated genes adjacent to the gene encoding SwASST16 (Table 1). One is the previously mentioned LysR family transcription regulator, and the other encodes flavocytochrome c. Flavocytochromes c are typically produced in large quantities under anaerobic conditions [80]. They are prominent in sulfate-reducing microbes, where soluble flavocytochrome c proteins transfer electrons to various acceptors, including sulfate, thiosulfate, sulfur, nitrate, and fumarate [81]. The involvement of flavocytochrome c proteins in the reduction of organosulfates has not been established in the literature. For this reason, the likelihood of SwASST16 reaction products, which are organosulfates, acting as direct terminal electron acceptors for flavocytochrome c seems minimal. Nonetheless, it would be intriguing to explore whether these organosulfates participate indirectly in a broader metabolic network that includes both ASST and flavocytochrome c proteins, thereby contributing to the electron transfer to organosulfates through more complex pathways.
Our gene neighborhood analysis has indicated that asst genes from S. wadsworthensis 3_1_45B typically do not integrate into recognizable operons. Instead, they appear to exist independently within the genome, as depicted in Figure 5. This finding leads us to hypothesize that the asst genes generally function as isolated units rather than as part of operonic clusters. Nevertheless, to fully understand the biological relevance of this arrangement, experimental investigations into the interactions between SwASSTs and the products of the neighboring genes are warranted.

2.6. Bioinformatic Analysis of Predicted ASST Proteins from S. wadsworthensis 3_1_45B

For the predicted SwASST proteins, we assessed protein length, presence of signal peptides, probable cellular localization, presence of transmembrane regions, amino acid sequence conservation, and structural variability.
Protein length variation among ASST enzymes of S. wadsworthensis 3_1_45B. The data presented in Figure 6 and Table S3 detail the amino acid lengths of the 17 identified ASST enzymes from S. wadsworthensis 3_1_45B. The majority of these enzymes consist of approximately 600 amino acids. Specifically, the longest ASST comprises 615 amino acids, while the shortest contains 449 amino acids. 16 of the 17 ASST proteins have lengths within a narrow range of 605 to 615 amino acids, exhibiting consistency in size among these enzymes.

2.7. Analysis of Signal Peptides in SwASST Enzymes

Given that a well-characterized ASST from E. coli has a signal peptide [29], we conducted a sequence analysis to determine the prevalence and types of signal peptides in SwASSTs from S. wadsworthensis 3_1_45B. Our findings reveal that 16 of the 17 annotated SwASSTs are predicted to have signal peptides, suggesting their translocation across the cytoplasmic membrane (Figure 7A and Table S3). The analysis of these signal peptides was carried out using SignalP 6.0, confirming the annotations provided by the IMG/M database. Upon examining the lengths of these signal peptides, the average length was found to be 26 amino acids, with a median of 25.5 amino acids. Among the 16 SwASSTs with signal peptides, 10 (SwASSTs 2, 4, 7, 8, 9, 10, 12, 13, 15, 17) harbor the Sec-type signal peptides associated with the general secretion pathway, while 6 (SwASSTs 1, 3, 5, 6, 14, 16) contain the Tat-type signal peptides linked to the twin-arginine translocation pathway (Figure 7B and Table 2). In addition, there is significant variability in the amino acid sequences of both Sec and Tat type signal peptides across the different annotated SwASSTs (Figure 8A,B).
Amino acid sequences of signal peptides from 17 annotated SwASSTs with the description of the type of secretary pathways (Sec or Tat). Underlined motifs are for Tat type signal peptides harboring conserved twin-arginine residues.
Sec and Tat translocases are common mechanisms for the transfer of proteins across the cytoplasmic membrane. The Sec system mediates protein transport across the membrane in an unfolded state, whereas the Tat system transports proteins in a folded state [82]. These proteins have distinct N-terminal signal peptides, with the Sec pathway typically featuring a positively charged N-terminus, a hydrophobic core, and polar C-terminal residues (Figure 8A and Table 2). The Tat pathway signal peptides are more conserved, particularly the twin-arginine residues (Figure 8B). However, SwASST14 and SwASST16 diverge from this pattern, presenting a lysine (K) in place of the first arginine (R) in the conserved region—a rare variant with potential implications for translocation efficiency, as reported in literature [83,84]. This observation suggests that SwASST14 and 16 may have a lower translocation efficiency than other SwASSTs with canonical twin-arginine motifs. Additionally, both Sec and Tat signal peptides frequently contain the conserved A-X-A motif at the C-terminus, a feature commonly found in N-terminal signal peptides [85].
Proteins secreted through the Sec or Tat pathways may localize to the periplasm, integrate into the cytoplasmic membrane, or be exported outside the cell via other secretion systems, especially in gram-negative bacteria [86]. Periplasmic and extracellular proteins typically follow a SecB-mediated route, with extracellular proteins requiring an additional translocation step across the outer membrane via Type II or Type V secretion systems (T2SS or T5SS) [86]. Genomic analysis of S. wadsworthensis 3_1_45B revealed the presence of T2SS components, suggesting that some SwASSTs might be T2SS substrates. Table 3 combines our findings regarding SwASSTs harboring the Sec or Tat signal peptides and proposes their potential cellular destinations.

2.8. Sequences Similarity Analysis of SwASSTs

For the multiple sequence alignments, E. coli ASST (EcASST) served as a reference sequence because it has been studied well in the context of structure–function relationships [29]. Utilizing the Clustal Omega tool from UniProt for multiple sequence alignment, we observed that all the key active site residues of EcASST, crucial for catalysis are all well conserved across all 17 SwASSTs, especially residues corresponding to His-252 (H252), His-356 (H356), Asn-358 (N358), Arg-374 (R374), and His-436 (H436) [29] (Figure 9). His-436 is the most important catalytic residue that becomes transiently sulfated during the catalysis by E. coli sulfotransferase. This happens due to two half-reactions that occur during the ASST catalysis. In the first half-reaction, a donor adds a sulfate group to the active site histidine residue. In the second half reaction, this sulfate group from a histidine residue is transferred to the acceptor molecule, which completes the catalytic cycle. The multiple sequence alignment of all 17 SwASST enzymes confirms that this catalytic histidine residue is perfectly conserved, suggesting a similar catalytic mechanism may be at play within these enzymes (Figure 9).
This alignment revealed an average sequence similarity of 62% among all 17 SwASSTs. The range of similarity spanned from a minimum of 53% to a maximum of 75.5%, with the median and mode being 62% and 61.25%, respectively (Figure 10A). The degree of similarity indicates that while the SwASST enzymes share a common structural framework, there is sufficient variation to suggest they may have specialized functions. Interestingly, SwASST2 and SwASST8 exhibited the greatest sequence similarity at 75.5%, whereas SwASST1 and SwASST3 shared the least similarity at 53%.
Additionally, our sequence alignments show that SwASSTs harbor more sequence similarities with each other than with EcASST. There are residues that are absolutely conserved in SwASSTs but are completely different in EcASST (Figure S1). For example, there are two conserved asparagine residues in all SwASSTs (1st one around 160–163 in and 2nd one around 161–164 in all SwASSTs except in the shorter SwASST11 where these are located at positions 28 and 29), which are not conserved in EcASST. EcASST has histidine and glycine residues in this place. Similarly, a threonine (T166) is replaced by an absolutely conserved glycine (164–167) residue in all SwASSTs. There are more such examples that can be seen in extended sequence alignment of these proteins. Based on sequence alignment, it is apparent that SwASST3 and 16 are more closely related to EcASST protein (Figure 10B, phylogenetic tree).

2.9. Structural Variations in SwASST Proteins

To understand differences at the structure level, structures of all 17 SwASSTs were generated using Alphafold2. The resulting confidence levels of each model, measured by the local distance difference test (lDDT) score, are plotted against their respective amino acid positions, which were generated by Alphafold2 (Figure S2) [87]. Generally, the lDDT confidence plots show decreased prediction reliability around residue positions 200, 350, and 550 for most SwASSTs. However, SwASST3, SwASST12, and SwASST16 exhibit less pronounced dips in confidence around residue 550. SwASST11, being smaller, displays a distinct confidence profile with no notable decrease in prediction reliability around the residues corresponding to position 550 in the other SwASSTs, such as in SwASSTs 3, 12, and 16. These observations highlight subtle structural differences among the SwASST proteins, possibly reflecting variations in their functional attributes. Structural alignment of SwASSTs, conducted using PyMol, revealed a shared common fold among these enzymes (Figure 11). The ability to superimpose these structures, shown in Figure 12, further supports their structural convergence. Additionally, an average root-mean-square deviation (RMSD) of 0.52 Å across all 17 aligned structures suggests a high degree of similarity [88]. While the overall secondary structure elements, such as alpha helices and beta sheets, are consistent across the majority of SwASSTs, SwASST11 deviates slightly, displaying two fewer beta sheets at the N-terminus, likely due to its shorter length. Additionally, a beta sheet spanning residues 318–323 in SwASST1 is oriented in the reverse direction compared to its counterparts in other SwASST enzymes.
As can be seen in Figure 12, most areas of SwASSTs are completely superimposable and very well aligned. However, there are some regions where we see a lower degree of structural alignment (Figure 13, blue and green areas). It is interesting to note that the areas with variable structural alignments can have either high or low amino acid sequence conservation. One such region is at the N-terminus with the conserved motif V/s/t-W-N-N-P-X-G-G-A-L/m/v-E-W (Figure 13, blue cluster, Figure S3), displays considerable variation in the loop positions among the aligned SwASST structures. Contrastingly, at the C-terminus, which is the other variable region (Figure 13, green cluster, Figure S3), the level of structural alignment diminishes relative to the rest of the SwASST structures. This region is characterized by low sequence conservation and exhibits the most significant variability in secondary structural elements across the SwASST family.
Sequence similarity studies and phylogenetic analyses indicate that SwASST3 is evolutionarily closest to E. coli ASST (EcASST), while SwASST6 is the most divergent (Figure 10B). Structural comparisons of EcASST [29] with SwASST3 and SwASST6 exhibit a substantial degree of alignment across the three structures (Figure 14). However, a loop region spanning residues 151–164 present in EcASST is absent in SwASST3 and SwASST6. Additionally, SwASST6 uniquely features an alpha helix and a loop between residues 555 and 568. We also included a ligand-bound structure of EcASST in our alignment (Figure 14). In the ligand-bound structure of EcASST (green), the active site binds a molecule of para-nitrophenol (pNP), a reaction product of EcASST with the sulfate donor para-nitrophenyl sulfate (pNPS). Figure 15 presents a detailed view of the active site, showing the ligand bound within EcASST and the catalytic histidine residues from all three structures. While the histidine residues from the two E. coli structures align perfectly with each other, the histidine residues from the SwASSTs align amongst themselves but show slight positional differences when compared to the catalytic histidine residue of EcASST. This observation suggests subtle variations in the active site residue positioning in SwASSTs relative to EcASST.

2.10. Different Classes of Sulfotransferases from the Members of the Human Gut Microbiome and Their Comparison to Human Sulfotransferases

Sulfotransferases have historically been categorized based on their dependency on PAPS as either PAPS-dependent, using PAPS as a sulfo donor, or PAPS-independent [29]. Human sulfotransferases, typically around 300 amino acids in length, exclusively use PAPS, whereas microbial aryl-sulfate sulfotransferases (ASSTs), also known as PAPS-independent sulfotransferases, are generally larger. Recent findings, however, have identified PAPS-dependent sulfotransferases in gut commensals; these microbial enzymes are longer than their human counterparts, averaging around 370 amino acids, yet smaller than ASSTs [21,22]. To understand the prevalence of sulfotransferase classes, we explored the genomes of dominant human gut microbes, using protein blast searches with either SwASST1 (a PAPS-independent aryl-sulfate sulfotransferase) or BT_0416 (a PAPS-dependent cholesterol sulfotransferase from Bacteroides thetaiotaomicron) as references. The results demonstrate a wide distribution of both sulfotransferase types across gut microbiome members (Figure S4). Specifically, the Bacteroides genus predominantly carries PAPS-dependent sulfotransferases, with Parabacteroides following closely (Figure S5), whereas PAPS-independent sulfotransferases (aryl-sulfate sulfotransferases) are abundant in Sutterella (Figure S6). Sequence (Figure S7) and structural (Figure S8) alignments were performed to elucidate the differences between these enzymes. The sequence similarity between SwASST1 and other PAPS-dependent sulfotransferases (BT_0416, hSULT1A1, and hSULT2B1) is very low. Within the PAPS-dependent class, however, there is higher sequence conservation. Structurally, there is a significant divergence between PAPS-independent (SwASST1) and PAPS-dependent sulfotransferases (Figure S8A). Human PAPS-dependent sulfotransferases align closely with each other and are completely superimposable (Figure S8B). When aligning all three PAPS-dependent sulfotransferases, including two from humans (hSULT1A1, and hSULT2B1) and one from gut microbes (BT_0416), some overlapping regions are apparent, but the structures are not completely superimposable (Figure S8C), particularly a large alpha-helix in BT_0416 that does not align with the others. A striking contrast is observed in the secondary structure composition between these two types of sulfotransferases. PAPS-dependent sulfotransferases are rich in alpha helices (Figure S8B,C), whereas PAPS-independent SwASSTs feature beta sheets as their dominant structural elements (Figure 12). This contrast suggests a divergent evolutionary path for these two classes of sulfotransferases.

3. Materials and Methods

3.1. Members of the Human Gut Microbiome Harboring Annotated asst Genes

IMG/M database (Department of Energy, Berkeley, California, USA) was used to search for the prevalent human gut microbes [89,90] that contain annotated genes for asst that are predicted to produce the protein aryl-sulfate sulfotransferases (ASSTs). A search of IMG/M database with either enzyme name, aryl-sulfate sulfotransferases or the enzyme ID, EC 2.8.2.22, produced lists of genomes containing predicted asst genes. By selecting genome name under the filter column, lists of predicted ASSTs for each specific genus were collected. The total number of sulfotransferase genes present in a genus was counted by adding annotated genes from all the species under that genus.

3.2. Annotated asst Genes from the Genus Sutterella

By selecting genus Sutterella in the IMG/M database, we were able to collect different species and strains harboring annotated asst genes. From this, Sutterella wadsworthensis 3_1_45B was selected due to the number of annotated sulfotransferases found in its genome. Locus tags, gene IDs, GC content, and gene neighborhoods were retrieved for all annotated asst genes via IMG/M database.

3.3. Properties of Predicted ASST Proteins from Sutterella wadsworthensis 3_1_45B

Amino acid sequences for all 17 predicted ASST proteins from S. wadsworthensis 3_1_45B were obtained from the IMG/M database and UniProt and were aligned to confirm that amino acid sequences retrieved from both sources are exactly the same.

3.4. Analysis of Signal Peptides

Amino acid sequences for all predicted sulfotransferases were analyzed with SignalP 6.0 using preset parameters to gain insights into the presence and variability of signal peptides. These collected datasets were also verified with IMG/M database. In addition to providing information about the presence or absence of signal peptides, SingalP 6.0 also allowed for identification of the possible signal peptide type, along with the length of the signal.

3.5. Multiple Sequence Alignment

Multiple sequence alignment for all 17 ASST amino acid sequences was performed with the help of UniProt alignment tool, which utilizes the Clustal Omega program 2.1. This analysis provided information about sequence conservation in predicted ASST proteins. Additionally, percent identity matrix for these 17 ASST sequences was obtained from the alignments. Multiple alignments were performed with and without the signal peptides. Alignments with signal peptides were generated to understand differences in these regions. Alignments without signal peptides were generated to understand the variability in ASST sequences.

3.6. ASST Structure Predictions with AlphaFold2

Amino acid sequences of ASSTs without signal peptides were uploaded to Colaboratory AlphaFold2 (ColabFold v1.5.3: AlphaFold2 using MMseqs2, DeepMind Technologies Limited) and default settings were used for all runs to predict protein structures [91,92]. The three conserved regions found from the multiple sequence alignment of SwASSTs with EcASST were highlighted in red in all ASST structures.

3.7. ASST Structural Alignments with PyMOL

The PyMOL (version 2.5.4, Schrödinger) alignment tool with the default parameters and align command with five iteration cycles and a cutoff of 2 Å was utilized to align all the structures except for the alignment ‘c’ described below. These alignments are as follows: (a) For the alignment among all 17 ASSTs, all structures were aligned to SwASST1. (b) The alignment of E. coli ASST (EcASST) with SwASST3 and SwASST6, where two separate PDB structures of EcASST were utilized. Structure with PDB ID 3ETT has para-nitrophenol (pNP) bound to the active site, while 3ELQ has no bound ligands in the active site. (c) The alignment of SwASST1 with Bacteroides thetaiotaomicron VPI-5482 sulfotransferase BT_0416 (structure generated via AlphaFold2), human sulfotransferase 1A1 (hSULT1A1, PDB ID 1LS6), and human sulfotransferase 2B1b (hSULT2B1b, PDB ID 1Q1Z). hSULT2B1b, hSULT1A1, BT_0416, and SwASST1 were aligned using the cealign command in PyMOL’s alignment tool due to low sequence similarity.

3.8. Search for PAPS-Dependent and PAPS-Independent Sulfotransferases in the Members of the Human Gut Microbiome

Using NCBI protein blast (National Library of Medicine, Bethesda, Maryland, USA), two separate blasts, one with a sulfotransferase from Bacteroides thetaiotaomicron sulfotransferase, BT_0416 and another one with an aryl sulfate sulfotransferase from Sutterella wadsworthensis 3_1_45B, SwASST1 were performed against members of the human gut microbiome. BT_0416 is PAPS-dependent sulfotransferase and SwASST1 is a PAPS-independent sulfotransferase. Bacteroides and Sutterella were each excluded from their respective blast searches. Each genus of gut microbes was entered into the filter bar to determine if that genus contained PAPS-dependent and/or PAPS-independent sulfotransferases. From this search a binary dataset was created where presence was marked by 1 and absence was marked by 0. This dataset was utilized to create a heatmap. Predicted sulfotransferase proteins from different species and strains of each genus were added together to calculate the total number of predicted sulfotransferases (PAPS-dependent or independent) per genus.

4. Conclusions

Sulfotransferases from Sutterella wadsworthensis 3_1_45B exhibit a mixture of shared and unique structural characteristics. While these enzymes generally align with a common structural fold, certain regions display variability in both sequence conservation and structural configuration. The enzymes evaluated in this study are recognized for their activity on phenolic compounds, as established by prior research involving E. coli ASSTs and ASSTs from other gut microbes. These enzymes facilitate the transfer of sulfo groups, which modulates the levels of phenolic molecules. Uniquely, within the Sutterella genus, certain members possess a diversity of ASSTs within a single organism, with significant sequence variation—no two ASSTs examined here exhibit more than 75% sequence similarity. This is notable since proteins from the same genus and species, but different strains, that catalyze identical reactions usually display around 90% or more sequence homology.
In contrast, the ASSTs in this study, despite catalyzing the same reaction, exhibit lower homology, suggesting potential structural and functional variation. This is paralleled in human sulfotransferases (hSULTs). Humans have several hSULTs, which, although catalyzing identical chemical reactions, have divergent substrate specificities and functional diversifications. This diversity is partly due to variations in the sequences and conformations of loops near the active sites of these enzymes, which influence substrate specificity. Our predictions of functional divergence are based on sequence and structural alignments of SwASSTs and the understanding of the differences between ASSTs and human SULTs. Despite having structural distinctions and limited sequence similarity between these two classes of enzymes, the type of chemical reactions catalyzed by both are the same but with different donor and acceptor specificities. Therefore, it appears there has been an evolutionary branching in these two categories of sulfotransferases.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijms25052983/s1.

Author Contributions

Conceptualization: D.D.S.; methodology, L.L. and D.D.S.; formal analysis, L.L. and D.D.S.; investigation, L.L. and D.D.S.; resources, D.D.S.; data curation, L.L.; writing—original draft preparation, L.L. and D.D.S.; writing—review and editing, D.D.S.; visualization, L.L.; supervision, D.D.S.; project administration, D.D.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The original contributions presented in the study are included in the article/Supplementary Material, further inquiries can be directed to the corresponding author.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Coughtrie, M.W.; Bamforth, K.J.; Sharp, S.; Jones, A.L.; Borthwick, E.B.; Barker, E.V.; Roberts, R.C.; Hume, R.; Burchell, A. Sulfation of endogenous compounds and xenobiotics—Interactions and function in health and disease. Chem. Biol. Interact. 1994, 92, 247–256. [Google Scholar] [CrossRef]
  2. Idris, M.; Mitchell, D.J.; Gordon, R.; Sidharthan, N.P.; Butcher, N.J.; Minchin, R.F. Interaction of the Brain-Selective Sulfotransferase SULT4A1 with Other Cytosolic Sulfotransferases: Effects on Protein Expression and Function. Drug Metab. Dispos. 2020, 48, 337–344. [Google Scholar] [CrossRef]
  3. Bairam, A.F.; Rasool, M.I.; Alherz, F.A.; Abunnaja, M.S.; El Daibani, A.A.; Gohal, S.A.; Alatwi, E.S.; Kurogi, K.; Liu, M. Impact of SULT1A3/SULT1A4 genetic polymorphisms on the sulfation of phenylephrine and salbutamol by human SULT1A3 allozymes. Pharmacogenet. Genom. 2019, 29, 99–105. [Google Scholar] [CrossRef]
  4. Bairam, A.F.; Rasool, M.I.; Alherz, F.A.; Abunnaja, M.S.; El Daibani, A.A.; Kurogi, K.; Liu, M. Effects of human SULT1A3/SULT1A4 genetic polymorphisms on the sulfation of acetaminophen and opioid drugs by the cytosolic sulfotransferase SULT1A3. Arch. Biochem. Biophys. 2018, 648, 44–52. [Google Scholar] [CrossRef] [PubMed]
  5. Kurogi, K.; Shimohira, T.; Kouriki-Nagatomo, H.; Zhang, G.; Miller, E.R.; Sakakibara, Y.; Suiko, M.; Liu, M.-C. Human Cytosolic Sulphotransferase SULT1C3: Genomic analysis and functional characterization of splice variant SULT1C3a and SULT1C3d. J. Biochem. 2017, 162, 403–414. [Google Scholar] [CrossRef]
  6. Guidry, A.L.; Tibbs, Z.E.; Runge-Morris, M.; Falany, C.N. Expression, purification and characterization of human cytosolic sulfotransferase (SULT) 1C4. Horm. Mol. Biol. Clin. Investig. 2017, 29, 27–36. [Google Scholar] [CrossRef]
  7. Lu, J.; Li, H.; Zhang, J.; Li, M.; An, X.; Liu, M.-C.; Chang, W. Crystal structures of SULT1A2 and SULT1A1 *3: Insights into the substrate inhibition and the role of Tyr149 in SULT1A2. Biochem. Biophys. Res. Commun. 2010, 396, 429–434. [Google Scholar] [CrossRef]
  8. Cook, I.T.; Duniec-Dmuchowski, Z.; Kocarek, T.A.; Runge-Morris, M.; Falany, C.N. 24-hydroxycholesterol sulfation by human cytosolic sulfotransferases: Formation of monosulfates and disulfates, molecular modeling, sulfatase sensitivity, and inhibition of liver x receptor activation. Drug Metab. Dispos. 2009, 37, 2069–2078. [Google Scholar] [CrossRef]
  9. Lu, L.-Y.; Hsieh, Y.-C.; Liu, M.-Y.; Lin, Y.-H.; Chen, C.-J.; Yang, Y.-S. Identification and characterization of two amino acids critical for the substrate inhibition of human dehydroepiandrosterone sulfotransferase (SULT2A1). Mol. Pharmacol. 2008, 73, 660–668. [Google Scholar] [CrossRef]
  10. Allali-Hassani, A.; Pan, P.W.; Dombrovski, L.; Najmanovich, R.; Tempel, W.; Dong, A.; Loppnau, P.; Martin, F.; Thonton, J.; Edwards, A.M.; et al. Structural and chemical profiling of the human cytosolic sulfotransferases. PLoS Biol. 2007, 5, e97. [Google Scholar]
  11. Gamage, N.U.; Tsvetanov, S.; Duggleby, R.G.; McManus, M.E.; Martin, J.L. The structure of human SULT1A1 crystallized with estradiol. An insight into active site plasticity and substrate inhibition with multi-ring substrates. J. Biol. Chem. 2005, 280, 41482–41486. [Google Scholar] [CrossRef]
  12. Pedersen, L.C.; Petrotchenko, E.; Shevtsov, S.; Negishi, M. Crystal structure of the human estrogen sulfotransferase-PAPS complex: Evidence for catalytic role of Ser137 in the sulfuryl transfer reaction. J. Biol. Chem. 2002, 277, 17928–17932. [Google Scholar] [CrossRef]
  13. Fuda, H.; Lee, Y.C.; Shimizu, C.; Javitt, N.B.; Strott, C.A. Mutational analysis of human hydroxysteroid sulfotransferase SULT2B1 isoforms reveals that exon 1B of the SULT2B1 gene produces cholesterol sulfotransferase, whereas exon 1A yields pregnenolone sulfotransferase. J. Biol. Chem. 2002, 277, 36161–36166. [Google Scholar] [CrossRef]
  14. Freimuth, R.R.; Raftogianis, R.B.; Wood, T.C.; Moon, E.; Kim, U.-J.; Xu, J.; Siciliano, M.J.; Weinshilboum, R.M. Human sulfotransferases SULT1C1 and SULT1C2: cDNA characterization, gene cloning, and chromosomal localization. Genomics 2000, 65, 157–165. [Google Scholar] [CrossRef]
  15. Wang, J.; Falany, J.L.; Falany, C.N. Expression and characterization of a novel thyroid hormone-sulfating form of cytosolic sulfotransferase from human liver. Mol. Pharmacol. 1998, 53, 274–282. [Google Scholar] [CrossRef]
  16. Sakakibara, Y.; Yanagisawa, K.; Katafuchi, J.; Ringer, D.P.; Takami, Y.; Nakayama, T.; Suiko, M.; Liu, M. Molecular cloning, expression, and characterization of novel human SULT1C sulfotransferases that catalyze the sulfonation of N-hydroxy-2-acetylaminofluorene. J. Biol. Chem. 1998, 273, 33929–33935. [Google Scholar] [CrossRef]
  17. Fujita, K.-I.; Nagata, K.; Ozawa, S.; Sasano, H.; Yamazoe, Y. Molecular cloning and characterization of rat ST1B1 and human ST1B2 cDNAs, encoding thyroid hormone sulfotransferases. J. Biochem. 1997, 122, 1052–1061. [Google Scholar] [CrossRef]
  18. Falany, C.N.; Krasnykh, V.; Falany, J.L. Bacterial expression and characterization of a cDNA for human liver estrogen sulfotransferase. J. Steroid Biochem. Mol. Biol. 1995, 52, 529–539. [Google Scholar] [CrossRef]
  19. Radominska, A.; A Comer, K.; Zimniak, P.; Falany, J.; Iscan, M.; Falany, C.N. Human liver steroid sulphotransferase sulphates bile acids. Biochem. J. 1990, 272, 597–604. [Google Scholar] [CrossRef]
  20. Whittemore, R.M.; Pearce, L.B.; Roth, J.A. Purification and kinetic characterization of a phenol-sulfating form of phenol sulfotransferase from human brain. Arch. Biochem. Biophys. 1986, 249, 464–471. [Google Scholar] [CrossRef]
  21. Le, H.H.; Lee, M.-T.; Besler, K.R.; Comrie, J.M.C.; Johnson, E.L. Characterization of interactions of dietary cholesterol with the murine and human gut microbiome. Nat. Microbiol. 2022, 7, 1390–1403. [Google Scholar] [CrossRef]
  22. Yao, L.; D’agostino, G.D.; Park, J.; Hang, S.; Adhikari, A.A.; Zhang, Y.; Li, W.; Avila-Pacheco, J.; Bae, S.; Clish, C.B.; et al. A biosynthetic pathway for the selective sulfonation of steroidal metabolites by human gut bacteria. Nat. Microbiol. 2022, 7, 1404–1418. [Google Scholar] [CrossRef]
  23. Kim, D.-H.; Kobashi, K. The role of intestinal flora in metabolism of phenolic sulfate esters. Biochem. Pharmacol. 1986, 35, 3507–3510. [Google Scholar] [CrossRef]
  24. Kim, D.-H.; Konishi, L.; Kobashi, K. Purification, characterization and reaction mechanism of novel arylsulfotransferase obtained from an anaerobic bacterium of human intestine. Biochim. Biophys. Acta 1986, 872, 33–41. [Google Scholar] [CrossRef]
  25. Kim, D.-H.; Yoon, H.-K.; Koizumi, M.; Kobashi, K. Sulfation of phenolic antibiotics by sulfotransferase obtained from a human intestinal bacterium. Chem. Pharm. Bull. 1992, 40, 1056–1057. [Google Scholar] [CrossRef]
  26. Kobashi, K.; Fukaya, Y.; Kim, D.-H.; Akao, T.; Takebe, S. A novel type of aryl sulfotransferase obtained from an anaerobic bacterium of human intestine. Arch. Biochem. Biophys. 1986, 245, 537–539. [Google Scholar] [CrossRef]
  27. Suiko, M.; Kurogi, K.; Hashiguchi, T.; Sakakibara, Y.; Liu, M.-C. Updated perspectives on the cytosolic sulfotransferases (SULTs) and SULT-mediated sulfation. Biosci. Biotechnol. Biochem. 2017, 81, 63–72. [Google Scholar] [CrossRef]
  28. Tibbs, Z.E.; Rohn-Glowacki, K.J.; Crittenden, F.; Guidry, A.L.; Falany, C.N. Structural plasticity in the human cytosolic sulfotransferase dimer and its role in substrate selectivity and catalysis. Drug Metab. Pharmacokinet. 2015, 30, 3–20. [Google Scholar] [CrossRef]
  29. Malojčić, G.; Owen, R.L.; Grimshaw, J.P.A.; Brozzo, M.S.; Dreher-Teo, H.; Glockshuber, R. A structural and biochemical basis for PAPS-independent sulfuryl transfer by aryl sulfotransferase from uropathogenic Escherichia coli. Proc. Natl. Acad. Sci. USA 2008, 105, 19217–19222. [Google Scholar] [CrossRef]
  30. Gamage, N.; Barnett, A.; Hempel, N.; Duggleby, R.G.; Windmill, K.F.; Martin, J.L.; McManus, M.E. Human sulfotransferases and their role in chemical metabolism. Toxicol. Sci. 2006, 90, 5–22. [Google Scholar] [CrossRef]
  31. Zhang, H.; Varlamova, O.; Vargas, F.M.; Falany, C.N.; Leyh, T.S. Sulfuryl transfer: The catalytic mechanism of human estrogen sulfotransferase. J. Biol. Chem. 1998, 273, 10888–10892. [Google Scholar] [CrossRef]
  32. Konishi-Imamura, L.; Dohi, K.; Sato, M.; Kobashi, K. Improved purification of arylsulfate sulfotransferase from human intestinal bacterium by using polyclonal antibody. J. Biochem. 1994, 115, 1097–1100. [Google Scholar] [CrossRef]
  33. Baek, M.-C.; Kwon, A.-R.; Chung, Y.-J.; Kim, B.-K.; Choi, E.-C. Distribution of bacteria with the arylsulfate sulfotransferase activity. Arch. Pharm. Res. 1998, 21, 475–477. [Google Scholar] [CrossRef]
  34. Koppel, N.; Rekdal, V.M.; Balskus, E.P. Chemical transformation of xenobiotics by the human gut microbiota. Science 2017, 356, 6344. [Google Scholar] [CrossRef]
  35. Wexler, H.M.; Reeves, D.; Summanen, P.H.; Molitoris, E.; McTEAGUE, M.; Duncan, J.; Wilson, K.H.; Finegold, S.M. Sutterella wadsworthensis gen. nov., sp. nov., bile-resistant microaerophilic Campylobacter gracilis-like clinical isolates. Int. J. Syst. Bacteriol. 1996, 46, 252–258. [Google Scholar] [CrossRef]
  36. Engberg, J.; On, S.L.W.; Harrington, C.S.; Gerner-Smidt, P. Prevalence of Campylobacter, Arcobacter, Helicobacter, and Sutterella spp. in human fecal samples as estimated by a reevaluation of isolation methods for Campylobacters. J. Clin. Microbiol. 2000, 38, 286–291. [Google Scholar] [CrossRef]
  37. Sakon, H.; Nagai, F.; Morotomi, M.; Tanaka, R. Sutterella parvirubra sp. nov. and Megamonas funiformis sp. nov., isolated from human faeces. Int. J. Syst. Evol. Microbiol. 2008, 58 Pt 4, 970–975. [Google Scholar] [CrossRef]
  38. Ndongo, S.; Cadoret, F.; Dubourg, G.; Delerce, J.; Fournier, P.-E.; Raoult, D.; Lagier, J.-C. ‘Collinsella phocaeensis’ sp. nov., ‘Clostridium merdae’ sp. nov., ‘Sutterella massiliensis’ sp. nov., ‘Sutturella timonensis’ sp. nov., ‘Enorma phocaeensis’ sp. nov., ‘Mailhella massiliensis’ gen. nov., sp. nov., ‘Mordavella massiliensis’ gen. nov., sp. nov. and ‘Massiliprevotella massiliensis’ gen. nov., sp. nov., 9 new species isolated from fresh stool samples of healthy French patients. New Microbes New Infect. 2017, 17, 89–95. [Google Scholar]
  39. Dione, N.; Ngom, I.I.; Valles, C.; Cadoret, F.; Fournier, P.E.; Raoult, D.; Lagier, J.C. ‘Collinsella provencensis’ sp. nov., ‘Parabacteroides bouchesdurhonensis’ sp. nov. and ‘Sutterella seckii’ sp. nov., three new bacterial species identified from human gut microbiota. New Microbes New Infect. 2018, 23, 44–47. [Google Scholar] [CrossRef]
  40. Sakamoto, M.; Ikeyama, N.; Kunihiro, T.; Iino, T.; Yuki, M.; Ohkuma, M. Mesosutterella multiformis gen. nov., sp. nov., a member of the family Sutterellaceae and Sutterella megalosphaeroides sp. nov., isolated from human faeces. Int. J. Syst. Evol. Microbiol. 2018, 68, 3942–3950. [Google Scholar] [CrossRef] [PubMed]
  41. Oh, B.S.; Kim, J.S.; Yu, S.Y.; Ryu, S.W.; Park, S.H.; Kang, S.W.; Park, J.-E.; Choi, S.-H.; Han, K.-I.; Lee, K.C.; et al. Sutterella faecalis sp. nov., isolated from human faeces. J. Microbiol. 2020, 58, 99–104. [Google Scholar] [CrossRef]
  42. Kaakoush, N.O. Sutterella Species, IgA-degrading Bacteria in Ulcerative Colitis. Trends Microbiol. 2020, 28, 519–522. [Google Scholar] [CrossRef]
  43. Williams, B.L.; Hornig, M.; Parekh, T.; Lipkin, W.I. Application of novel PCR-based methods for detection, quantitation, and phylogenetic characterization of Sutterella species in intestinal biopsy samples from children with autism and gastrointestinal disturbances. mBio 2012, 3, e00261-11. [Google Scholar] [CrossRef]
  44. Wang, L.; Christophersen, C.T.; Sorich, M.J.; Gerber, J.P.; Angley, M.T.; Conlon, M.A. Increased abundance of Sutterella spp. and Ruminococcus torques in feces of children with autism spectrum disorder. Mol. Autism. 2013, 4, 42. [Google Scholar] [CrossRef]
  45. Kirk, K.F.; Andersen, K.L.; Tarpgaard, I.H.; Nielsen, H.L. Three cases of Sutterella wadsworthensis bacteremia secondary to abdominal infections. Anaerobe 2021, 72, 102460. [Google Scholar] [CrossRef]
  46. Giri, S.; Mangalam, A. Chapter 34—The Gut Microbiome and Metabolome in Multiple Sclerosis. In Microbiome and Metabolome in Diagnosis, Therapy, and other Strategic Applications; Faintuch, J., Faintuch, S., Eds.; Academic Press: Cambridge, MA, USA, 2019; pp. 333–340. [Google Scholar]
  47. Cornejo-Pareja, I.; Ruiz-Limón, P.; Gómez-Pérez, A.M.; Molina-Vega, M.; Moreno-Indias, I.; Tinahones, F.J. Differential Microbial Pattern Description in Subjects with Autoimmune-Based Thyroid Diseases: A Pilot Study. J. Pers. Med. 2020, 10, 192. [Google Scholar] [CrossRef]
  48. Hiippala, K.; Kainulainen, V.; Kalliomäki, M.; Arkkila, P.; Satokari, R. Mucosal Prevalence and Interactions with the Epithelium Indicate Commensalism of Sutterella spp. Front. Microbiol. 2016, 7, 1706. [Google Scholar] [CrossRef]
  49. Raghavan, R.; Kelkar, Y.D.; Ochman, H. A selective force favoring increased G+C content in bacterial genes. Proc. Natl. Acad. Sci. USA 2012, 109, 14504–14507. [Google Scholar] [CrossRef]
  50. Schell, M.A. Molecular biology of the LysR family of transcriptional regulators. Annu. Rev. Microbiol. 1993, 47, 597–626. [Google Scholar] [CrossRef]
  51. van der Ploeg, J.R.; Eichhorn, E.; Leisinger, T. Sulfonate-sulfur metabolism and its regulation in Escherichia coli. Arch. Microbiol. 2001, 176, 1–8. [Google Scholar] [CrossRef]
  52. Kouzuma, A.; Endoh, T.; Omori, T.; Nojiri, H.; Yamane, H.; Habe, H. Transcription factors CysB and SfnR constitute the hierarchical regulatory system for the sulfate starvation response in Pseudomonas putida. J. Bacteriol. 2008, 190, 4521–4531. [Google Scholar] [CrossRef]
  53. Pokhrel, A.; Dinh, H.; Li, L.; Hassan, K.A.; Cain, A.K.; Paulsen, I.T. Identification of a novel LysR family transcriptional regulator controlling acquisition of sulfur sources in Acinetobacter baumannii. Microb. Physiol. 2023, 33, 27–35. [Google Scholar] [CrossRef]
  54. Wang, J.P.; Zhang, W.M.; Chao, H.J.; Zhou, N.Y. PnpM, a LysR-Type Transcriptional Regulator Activates the Hydroquinone Pathway in para-Nitrophenol Degradation in Pseudomonas sp. Strain WBC-3. Front. Microbiol. 2017, 8, 1714. [Google Scholar] [CrossRef]
  55. Lu, Z.; Imlay, J.A. The Fumarate Reductase of Bacteroides thetaiotaomicron, unlike That of Escherichia coli, Is Configured so that It Does Not Generate Reactive Oxygen Species. mBio 2017, 8, e01873-16. [Google Scholar] [CrossRef]
  56. Laue, H.; Denger, K.; Cook, A.M. Taurine reduction in anaerobic respiration of Bilophila wadsworthia RZATAU. Appl. Environ. Microbiol. 1997, 63, 2016–2021. [Google Scholar] [CrossRef]
  57. Häse, C.C.; A Finkelstein, R. Bacterial extracellular zinc-containing metalloproteases. Microbiol. Rev. 1993, 57, 823–837. [Google Scholar] [CrossRef]
  58. Garmory, H.S.; Titball, R.W. ATP-binding cassette transporters are targets for the development of antibacterial vaccines and therapies. Infect. Immun. 2004, 72, 6757–6763. [Google Scholar] [CrossRef] [PubMed]
  59. Kertesz, M.A. Bacterial transporters for sulfate and organosulfur compounds. Res. Microbiol. 2001, 152, 279–290. [Google Scholar] [CrossRef]
  60. Chen, J.; Xie, J. Role and regulation of bacterial LuxR-like regulators. J. Cell Biochem. 2011, 112, 2694–2702. [Google Scholar] [CrossRef] [PubMed]
  61. Dong, X.; Guthrie, B.G.H.; Alexander, M.; Noecker, C.; Ramirez, L.; Glasser, N.R.; Turnbaugh, P.J.; Balskus, E.P. Genetic manipulation of the human gut bacterium Eggerthella lenta reveals a widespread family of transcriptional regulators. Nat. Commun. 2022, 13, 7624. [Google Scholar] [CrossRef]
  62. Ryan, J.; McKillen, M.; Mason, J. Sulphate/molybdate interactions: In vivo and in vitro studies on the group VI oxyanion transport system in ovine renal tubule epithelial cells. Ann. Rech. Vet. 1987, 18, 47–55. [Google Scholar]
  63. Self, W.T.; Grunden, A.M.; Hasona, A.; Shanmugam, K.T. Molybdate transport. Res. Microbiol. 2001, 152, 311–321. [Google Scholar] [CrossRef]
  64. Mason, J.; Cardin, C. The competition of molybdate and sulphate ions for a transport system in the ovine small intestine. Res. Vet. Sci. 1977, 22, 313–315. [Google Scholar] [CrossRef]
  65. Toth, E.A.; Yeates, T.O. The structure of adenylosuccinate lyase, an enzyme with dual activity in the de novo purine biosynthetic pathway. Structure 2000, 8, 163–174. [Google Scholar] [CrossRef]
  66. Pruss, K.M.; Enam, F.; Battaglioli, E.; DeFeo, M.; Diaz, O.R.; Higginbottom, S.K.; Fischer, C.R.; Hryckowian, A.J.; Van Treuren, W.; Dodd, D.; et al. Oxidative ornithine metabolism supports non-inflammatory C. difficile colonization. Nat. Metab. 2022, 4, 19–28. [Google Scholar] [CrossRef]
  67. Rawlings, N.D.; Barrett, A.J.; Thomas, P.D.; Huang, X.; Bateman, A.; Finn, R.D. The MEROPS database of proteolytic enzymes, their substrates and inhibitors in 2017 and a comparison with peptidases in the PANTHER database. Nucleic Acids Res. 2018, 46, D624–D632. [Google Scholar] [CrossRef]
  68. Amedei, A.; Capasso, C.; Nannini, G.; Supuran, C.T. Microbiota, Bacterial Carbonic Anhydrases, and Modulators of Their Activity: Links to Human Diseases? Mediators Inflamm. 2021, 2021, 6926082. [Google Scholar] [CrossRef]
  69. Grölund, M.-M.; Lehtonen, O.-P.; Eerola, E.; Kero, P. Fecal microflora in healthy infants born by different methods of delivery: Permanent changes in intestinal flora after cesarean delivery. J. Pediatr. Gastroenterol. Nutr. 1999, 28, 19–25. [Google Scholar]
  70. Hall, B.M.; Breidenstein, E.B.M.; de la Fuente-Núñez, C.; Reffuveille, F.; Mawla, G.D.; Hancock, R.E.W.; Baker, T.A. Two Isoforms of Clp Peptidase in Pseudomonas aeruginosa Control Distinct Aspects of Cellular Physiology. J. Bacteriol. 2017, 199, e00568-16. [Google Scholar] [CrossRef]
  71. Bolam, D.N.; van den Berg, B. TonB-dependent transport by the gut microbiota: Novel aspects of an old problem. Curr. Opin. Struct. Biol. 2018, 51, 35–43. [Google Scholar] [CrossRef]
  72. Pollet, R.M.; Martin, L.M.; Koropatkin, N.M. TonB-dependent transporters in the Bacteroidetes: Unique domain structures and potential functions. Mol. Microbiol. 2021, 115, 490–501. [Google Scholar] [CrossRef]
  73. Doerfel, L.K.; Wohlgemuth, I.; Kothe, C.; Peske, F.; Urlaub, H.; Rodnina, M.V. EF-P is essential for rapid synthesis of proteins containing consecutive proline residues. Science 2013, 339, 85–88. [Google Scholar] [CrossRef]
  74. Kang, T.M.; Yuan, J.; Zhou, A.; Beppler, C.; Miller, J.H. Deoxycytidine deaminase-deficient Escherichia coli strains display acute sensitivity to cytidine, adenosine, and guanosine and increased sensitivity to a range of antibiotics, including vancomycin. J. Bacteriol. 2014, 196, 1950–1957. [Google Scholar] [CrossRef]
  75. Sugrue, E.; Fraser, N.J.; Hopkins, D.H.; Carr, P.D.; Khurana, J.L.; Oakeshott, J.G.; Scott, C.; Jackson, C.J. Evolutionary expansion of the amidohydrolase superfamily in bacteria in response to the synthetic compounds molinate and diuron. Appl. Environ. Microbiol. 2015, 81, 2612–2624. [Google Scholar] [CrossRef]
  76. Erb, T.J. Carboxylases in natural and synthetic microbial pathways. Appl. Environ. Microbiol. 2011, 77, 8466–8477. [Google Scholar] [CrossRef]
  77. Monti, S.M.; De Simone, G.; D’Ambrosio, K. L-Histidinol Dehydrogenase as a New Target for Old Diseases. Curr. Top. Med. Chem. 2016, 16, 2369–2378. [Google Scholar] [CrossRef] [PubMed]
  78. Chioccioli, S.; Del Duca, S.; Vassallo, A.; Castronovo, L.M.; Fani, R. Exploring the role of the histidine biosynthetic hisF gene in cellular metabolism and in the evolution of (ancestral) genes: From LUCA to the extant (micro)organisms. Microbiol. Res. 2020, 240, 126555. [Google Scholar] [CrossRef] [PubMed]
  79. Lewinson, O.; Lee, A.T.; Rees, D.C. A P-type ATPase importer that discriminates between essential and toxic transition metals. Proc. Natl. Acad. Sci. USA 2009, 106, 4677–4682. [Google Scholar] [CrossRef]
  80. Pealing, S.L.; Black, A.C.; Manson, F.D.C.; Ward, F.B.; Chapman, S.K.; Reid, G.A. Sequence of the gene encoding flavocytochrome c from Shewanella putrefaciens: A tetraheme flavoenzyme that is a soluble fumarate reductase related to the membrane-bound enzymes from other bacteria. Biochemistry 1992, 31, 12132–12140. [Google Scholar] [CrossRef]
  81. He, S.H.; DerVartanian, D.V.; LeGall, J. Isolation of fumarate reductase from Desulfovibrio multispirans, a sulfate reducing bacterium. Biochem. Biophys. Res. Commun. 1986, 135, 1000–1007. [Google Scholar] [CrossRef]
  82. Natale, P.; Brüser, T.; Driessen, A.J. Sec- and Tat-mediated protein secretion across the bacterial cytoplasmic membrane—Distinct translocases and mechanisms. Biochim. Biophys. Acta 2008, 1778, 1735–1756. [Google Scholar] [CrossRef]
  83. Hinsley, A.P.; Stanley, N.R.; Palmer, T.; Berks, B.C. A naturally occurring bacterial Tat signal peptide lacking one of the ‘invariant’ arginine residues of the consensus targeting motif. FEBS Lett. 2001, 497, 45–49. [Google Scholar] [CrossRef]
  84. Molik, S.; Karnauchov, I.; Weidlich, C.; Herrmann, R.G.; Klosgen, R.B. The Rieske Fe/S protein of the cytochrome b6/f complex in chloroplasts: Missing link in the evolution of protein transport pathways in chloroplasts? J. Biol. Chem. 2001, 276, 42761–42766. [Google Scholar] [CrossRef]
  85. Frain, K.M.; Robinson, C.; van Dijl, J.M. Transport of Folded Proteins by the Tat System. Protein J. 2019, 38, 377–388. [Google Scholar] [CrossRef]
  86. Green, E.R.; Mecsas, J. Bacterial Secretion Systems: An Overview. Microbiol. Spectr. 2016, 4, 1. [Google Scholar] [CrossRef]
  87. Mariani, V.; Biasini, M.; Barbato, A.; Schwede, T. lDDT: A local superposition-free score for comparing protein structures and models using distance difference tests. Bioinformatics 2013, 29, 2722–2728. [Google Scholar] [CrossRef]
  88. Carugo, O.; Pongor, S. A normalized root-mean-square distance for comparing protein three-dimensional structures. Protein Sci. 2001, 10, 1470–1473. [Google Scholar] [CrossRef]
  89. King, C.H.; Desai, H.; Sylvetsky, A.C.; LoTempio, J.; Ayanyan, S.; Carrie, J.; Crandall, K.A.; Fochtman, B.C.; Gasparyan, L.; Gulzar, N.; et al. Baseline human gut microbiota profile in healthy people and standard reporting template. PLoS ONE 2019, 14, e0206484. [Google Scholar] [CrossRef]
  90. Saito, Y.; Sato, T.; Nomoto, K.; Tsuji, H. Identification of phenol- and p-cresol-producing intestinal bacteria by using media supplemented with tyrosine and its metabolites. FEMS Microbiol. Ecol. 2018, 94, fiy125. [Google Scholar] [CrossRef] [PubMed]
  91. Jumper, J.; Evans, R.; Pritzel, A.; Green, T.; Figurnov, M.; Ronneberger, O.; Tunyasuvunakool, K.; Bates, R.; Žídek, A.; Potapenko, A.; et al. Highly accurate protein structure prediction with AlphaFold. Nature 2021, 596, 583–589. [Google Scholar] [CrossRef]
  92. Mirdita, M.; Schütze, K.; Moriwaki, Y.; Heo, L.; Ovchinnikov, S.; Steinegger, M. ColabFold: Making protein folding accessible to all. Nat. Methods 2022, 19, 679–682. [Google Scholar] [CrossRef]
Figure 1. Depiction of a sulfotransferase-catalyzed reaction. This scheme illustrates the general chemical reaction catalyzed by sulfotransferases where a sulfo group is transferred from a donor molecule to an acceptor molecule.
Figure 1. Depiction of a sulfotransferase-catalyzed reaction. This scheme illustrates the general chemical reaction catalyzed by sulfotransferases where a sulfo group is transferred from a donor molecule to an acceptor molecule.
Ijms 25 02983 g001
Figure 2. Distribution of annotated asst genes in the human gut microbes. The pie chart represents the distribution of annotated aryl-sulfate sulfotransferase (asst) genes across various known gut microbial genera. Each segment’s proportion reflects the relative count of asst genes within that particular genus.
Figure 2. Distribution of annotated asst genes in the human gut microbes. The pie chart represents the distribution of annotated aryl-sulfate sulfotransferase (asst) genes across various known gut microbial genera. Each segment’s proportion reflects the relative count of asst genes within that particular genus.
Ijms 25 02983 g002
Figure 3. Annotated asst genes within the genus Sutterella. The bar graph represents the quantity of annotated asst genes across various species and strains of Sutterella, highlighting the substantial variability within the genus.
Figure 3. Annotated asst genes within the genus Sutterella. The bar graph represents the quantity of annotated asst genes across various species and strains of Sutterella, highlighting the substantial variability within the genus.
Ijms 25 02983 g003
Figure 4. Variation in the %GC content of asst genes. This figure shows a bar graph representation of %GC content for all annotated asst genes from S. wadsworthensis 3_1_45B.
Figure 4. Variation in the %GC content of asst genes. This figure shows a bar graph representation of %GC content for all annotated asst genes from S. wadsworthensis 3_1_45B.
Ijms 25 02983 g004
Figure 5. Gene neighborhood maps of annotated asst genes. Each gene is depicted as an arrow and is represented by a unique color. All asst genes (from 1 to 17) are represented in purple color. Size of each arrow is approximately proportional to the gene size.
Figure 5. Gene neighborhood maps of annotated asst genes. Each gene is depicted as an arrow and is represented by a unique color. All asst genes (from 1 to 17) are represented in purple color. Size of each arrow is approximately proportional to the gene size.
Ijms 25 02983 g005
Figure 6. Variation in the length of ASST enzymes of S. wadsworthensis 3_1_45B. The plot illustrates the range of amino acid counts as comparison of protein lengths across the 17 predicted ASST enzymes.
Figure 6. Variation in the length of ASST enzymes of S. wadsworthensis 3_1_45B. The plot illustrates the range of amino acid counts as comparison of protein lengths across the 17 predicted ASST enzymes.
Ijms 25 02983 g006
Figure 7. Signal peptide characteristics in predicted SwASSTs. (A) Depicts the length of signal peptides in amino acid (AA) residues. (B) Categorizes the signal peptides by type (Sec or Tat) within the SwASSTs, where Sec type is represented as green and Tat type as blue.
Figure 7. Signal peptide characteristics in predicted SwASSTs. (A) Depicts the length of signal peptides in amino acid (AA) residues. (B) Categorizes the signal peptides by type (Sec or Tat) within the SwASSTs, where Sec type is represented as green and Tat type as blue.
Ijms 25 02983 g007
Figure 8. Sequence alignments for signal peptides of SwASSTs. (A) Sequence alignment for Sec type signal peptides of SwASSTs. (B) Sequence alignment for Tat type signal peptides of SwASSTs. Color bar above the sequences shows the amino acid conservation at that position, where dark red is highly conserved and dark blue indicates the least conserved position.
Figure 8. Sequence alignments for signal peptides of SwASSTs. (A) Sequence alignment for Sec type signal peptides of SwASSTs. (B) Sequence alignment for Tat type signal peptides of SwASSTs. Color bar above the sequences shows the amino acid conservation at that position, where dark red is highly conserved and dark blue indicates the least conserved position.
Ijms 25 02983 g008
Figure 9. Multiple sequence alignment of SwASSTs. Alignment of SwASSTs with EcASST highlighting catalytically crucial residues in red. Color bar above the consensus line shows the amino acid conservation at that position. Dark red is highly conserved, while dark blue has little to no conservation.
Figure 9. Multiple sequence alignment of SwASSTs. Alignment of SwASSTs with EcASST highlighting catalytically crucial residues in red. Color bar above the consensus line shows the amino acid conservation at that position. Dark red is highly conserved, while dark blue has little to no conservation.
Ijms 25 02983 g009
Figure 10. Sequence similarity analysis of SwASSTs from S. wadsworthensis 3_1_45B. (A) Percent identity matrix (B) Phylogenetic tree of the 17 predicted SwASSTs and E. coli ASST (EcASST) from E. coli strain CFT073.
Figure 10. Sequence similarity analysis of SwASSTs from S. wadsworthensis 3_1_45B. (A) Percent identity matrix (B) Phylogenetic tree of the 17 predicted SwASSTs and E. coli ASST (EcASST) from E. coli strain CFT073.
Ijms 25 02983 g010
Figure 11. Structures of SwASST proteins generated by Alphafold2. From (AQ) are predicted structures of SwASST1 to SwASST17. Conserved regions harboring active site residues are highlighted in red.
Figure 11. Structures of SwASST proteins generated by Alphafold2. From (AQ) are predicted structures of SwASST1 to SwASST17. Conserved regions harboring active site residues are highlighted in red.
Ijms 25 02983 g011
Figure 12. Structural alignment of SwASST proteins. The figure depicts aligned Alphafold2-derived structures of SwASSTs with conserved regions highlighted in red.
Figure 12. Structural alignment of SwASST proteins. The figure depicts aligned Alphafold2-derived structures of SwASSTs with conserved regions highlighted in red.
Ijms 25 02983 g012
Figure 13. Unique clusters in SwASSTs alignments. Dark blue region represents the N-terminus cluster, and the C-terminus cluster is depicted in green. These regions show a lower degree of structural alignments across the SwASSTs.
Figure 13. Unique clusters in SwASSTs alignments. Dark blue region represents the N-terminus cluster, and the C-terminus cluster is depicted in green. These regions show a lower degree of structural alignments across the SwASSTs.
Ijms 25 02983 g013
Figure 14. Alignment SwASST3 and SwASST6 with EcASST. The figure shows aligned structures of SwASST3 (pink) and SwASST6 (yellow) with free (cyan, PDB ID: 3ELQ) and ligand bound (green, PDB ID: 3ETT) structures of EcASST. There is a molecule of para-nitrophenol (shown as sticks in blue color) bound in the active site of EcASST (green, PDB ID: 3ETT). Regions in red show catalytic residues from EcASST (His-252, His-356, Asn-358, Arg-374, and His-436).
Figure 14. Alignment SwASST3 and SwASST6 with EcASST. The figure shows aligned structures of SwASST3 (pink) and SwASST6 (yellow) with free (cyan, PDB ID: 3ELQ) and ligand bound (green, PDB ID: 3ETT) structures of EcASST. There is a molecule of para-nitrophenol (shown as sticks in blue color) bound in the active site of EcASST (green, PDB ID: 3ETT). Regions in red show catalytic residues from EcASST (His-252, His-356, Asn-358, Arg-374, and His-436).
Ijms 25 02983 g014
Figure 15. Active site of SwASSTs (3 and 6) and EcASST highlighting the catalytic histidine residues. The figure displays a zoomed-in active site of SwASST3 (pink), SwASST6 (yellow), EcASST (cyan, free), and EcASST (green, pNP bound). Catalytic histidine residues are presented as sticks and highlighted in the same colors for each of the corresponding structures. pNP is shown as sticks (in cyan).
Figure 15. Active site of SwASSTs (3 and 6) and EcASST highlighting the catalytic histidine residues. The figure displays a zoomed-in active site of SwASST3 (pink), SwASST6 (yellow), EcASST (cyan, free), and EcASST (green, pNP bound). Catalytic histidine residues are presented as sticks and highlighted in the same colors for each of the corresponding structures. pNP is shown as sticks (in cyan).
Ijms 25 02983 g015
Table 1. List of genes found in the neighborhood of annotated asst genes that encode for SwASST proteins.
Table 1. List of genes found in the neighborhood of annotated asst genes that encode for SwASST proteins.
Annotated Genes in the Neighborhood of asst GenesNumber of Occurrences Around asst GenesFound in the Neighborhood of this asst Gene
LysR transcriptional regulator protein71, 2, 6, 9, 10, 16, 17
Fumarate reductase flavoprotein subunit31, 5, 13
Putative peptide zinc metalloprotease11
Putative ATP-binding cassette transporter12
LuxR transcriptional regulator protein13
modC molybdate ABC transporter13
Adenylosuccinate lyase13
Ornithine carbamoyltransferase14
Oligopeptidase metallo peptidase MEROPS family14
Carbonic anhydrase16
ATP-dependent Clp protease proteolytic subunit18
Outer membrane transport energization protein TonB19
Translation elongation factor P110
Deoxycytidylate deaminase110
Amidohydrolase110
Pyruvate carboxylase111
ATP phosphoribosyltransferase114
Information from Figure 5 shown as a list.
Table 2. Signal peptide sequences and types of SwASSTs.
Table 2. Signal peptide sequences and types of SwASSTs.
Predicted ProteinType of SPSignal Peptide Sequence
SwASST1Tat typeMKFQRRQLVAAMAGLMLGLSAGAASA
SwASST2Sec typeMRLFAKSLIAASIAAASMGALA
SwASST3Tat typeMSKMPTSPISRRTLCGLAGSALFLSQFPRAAFA
SwASST4Sec typeMSKQKAHQKLNQLTAVLMMAAAAALAAPAVQA
SwASST5Tat typeMKNITIGRRSFFLGAVAAAVSSALPATAAHA
SwASST6Tat typeMIQRRTLVKAVIATLAPLSLTSAFA
SwASST7Sec typeMRISKTAIAVSVLSLFAATMVPATSMA
SwASST8Sec typeMKMKWIALAVAAAIPAAAMA
SwASST9Sec typeMKLNFHPTRAAVLVGAALLSATFAFPAMA
SwASST10Sec typeMQFQKTAFALCAAGLLSLPGLVAA
SwASST12Sec typeMNIKAKITPVALAIALAAGFGLPAAVQA
SwASST13Sec typeMKSMKLSKIAAFAAFLGFSAAVA
SwASST14Tat typeMQLTKRAFLLSGAALAVSAASGAFA
SwASST15Sec typeMKKIVPICAAVAAAAMMLSISTPAMA
SwASST16Tat typeMLLTKRQFLTSVLALSVSAAARA
SwASST17Sec typeMKFKSTVIAASVLAGIMSLSAGAYA
Table 3. Potential cellular destinations of SwASSTs.
Table 3. Potential cellular destinations of SwASSTs.
Predicted ProteinSignal Peptide
Type
Sec Guided (Unfolded State) Periplasmic/Extracellular ProteinsTat Guided (Folded State) Periplasmic/Extracellular Proteins
SwASST1Tat
SwASST2Sec+
SwASST3Tat +
SwASST4Sec
SwASST5Tat +
SwASST6Tat +
SwASST7Sec
SwASST8Sec
SwASST9Sec
SwASST10Sec+
SwASST11-
SwASST12Sec
SwASST13Sec+
SwASST14Tat
SwASST15Sec
SwASST16Tat +
SwASST17Sec
The table indicates the type of signal peptides from SwASSTs and possible cellular destinations.
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

Langford, L.; Shah, D.D. Bioinformatic Analysis of Sulfotransferases from an Unexplored Gut Microbe, Sutterella wadsworthensis 3_1_45B: Possible Roles towards Detoxification via Sulfonation by Members of the Human Gut Microbiome. Int. J. Mol. Sci. 2024, 25, 2983. https://doi.org/10.3390/ijms25052983

AMA Style

Langford L, Shah DD. Bioinformatic Analysis of Sulfotransferases from an Unexplored Gut Microbe, Sutterella wadsworthensis 3_1_45B: Possible Roles towards Detoxification via Sulfonation by Members of the Human Gut Microbiome. International Journal of Molecular Sciences. 2024; 25(5):2983. https://doi.org/10.3390/ijms25052983

Chicago/Turabian Style

Langford, Lauryn, and Dhara D. Shah. 2024. "Bioinformatic Analysis of Sulfotransferases from an Unexplored Gut Microbe, Sutterella wadsworthensis 3_1_45B: Possible Roles towards Detoxification via Sulfonation by Members of the Human Gut Microbiome" International Journal of Molecular Sciences 25, no. 5: 2983. https://doi.org/10.3390/ijms25052983

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