Next Article in Journal
Symmetry Breaking Induced Pockels Effect in a Tilted Field Switching BPIII Cell
Next Article in Special Issue
Crystal Structure of Bacterial Cystathionine Γ-Lyase in The Cysteine Biosynthesis Pathway of Staphylococcus aureus
Previous Article in Journal
A Newly Synthesized Heterobimetallic NiII-GdIII Salamo-BDC-Based Coordination Polymer: Structural Characterization, DFT Calculation, Fluorescent and Antibacterial Properties
Previous Article in Special Issue
Crystal Structures of the 43 kDa ATPase Domain of Xanthomonas Oryzae pv. Oryzae Topoisomerase IV ParE Subunit and its Complex with Novobiocin
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Carboxylic Ester Hydrolases in Bacteria: Active Site, Structure, Function and Application

1
Department of Molecular Cell Biology, Sungkyunkwan University School of Medicine, Suwon 16419, Korea
2
Department of Chemistry, College of Natural Science, Sookmyung Women’s University, Seoul 04310, Korea
*
Authors to whom correspondence should be addressed.
Crystals 2019, 9(11), 597; https://doi.org/10.3390/cryst9110597
Submission received: 4 October 2019 / Revised: 30 October 2019 / Accepted: 7 November 2019 / Published: 14 November 2019
(This article belongs to the Special Issue Crystallographic Studies of Enzymes)

Abstract

:
Carboxylic ester hydrolases (CEHs), which catalyze the hydrolysis of carboxylic esters to produce alcohol and acid, are identified in three domains of life. In the Protein Data Bank (PDB), 136 crystal structures of bacterial CEHs (424 PDB codes) from 52 genera and metagenome have been reported. In this review, we categorize these structures based on catalytic machinery, structure and substrate specificity to provide a comprehensive understanding of the bacterial CEHs. CEHs use Ser, Asp or water as a nucleophile to drive diverse catalytic machinery. The α/β/α sandwich architecture is most frequently found in CEHs, but 3-solenoid, β-barrel, up-down bundle, α/β/β/α 4-layer sandwich, 6 or 7 propeller and α/β barrel architectures are also found in these CEHs. Most are substrate-specific to various esters with types of head group and lengths of the acyl chain, but some CEHs exhibit peptidase or lactamase activities. CEHs are widely used in industrial applications, and are the objects of research in structure- or mutation-based protein engineering. Structural studies of CEHs are still necessary for understanding their biological roles, identifying their structure-based functions and structure-based engineering and their potential industrial applications.

1. Introduction

Carboxylic ester hydrolases (CEHs, EC 3.1.1.-) are catalysts that hydrolyze linear and cyclic carboxylic ester bonds to produce carboxyl groups (–COOH) and alcohol groups (–OH) at termini. CEHs are found in all living organisms, including vertebrates, insects, fungi, plants, archaea and bacteria. Hydrolysis by CEHs is important for metabolite regulation [1], signal transduction [2], protein synthesis [3], stem elongation [4] and thermal stress response [5]. Offensive and defensive interactions between insects and plants are also mediated by CEHs [6]. Bacteria use CEHs to demolish cell walls and membrane structures for food uptake [7], and to infect hosts [8,9,10]. The functional diversity of CEHs is mediated by substrate specificities on various biomolecules, such as carbohydrates [11], lipids [12], polypeptides [13,14], nucleic acids [15] and other small molecules [16]. Their catalytic reaction is followed by the cooperation of catalytic residues including a classical Ser-His-Asp triad and substrate-binding residues.
In this review, we classified 136 bacterial CEHs deposited in the Protein Data Bank (PDB) [17] based on the source, their Enzyme Commission (EC) number, substrate, localization, active site structure, physiological function and three-dimensional (3D) structure, in order to provide a general overview of bacterial CEHs. In addition, their potency for industrial application is discussed.

2. Sampling of Bacterial CEH Structures in the Protein Data Bank

The structures of 525 bacterial enzymes belonging to the EC 3.1.1.- group were released in the PDB database until 2019 July. The 518 structures were identified using X-ray crystallography (XRC), and seven structures were identified using nuclear magnetic resonance (NMR). Here we only focused upon crystal structures. According to the PDB description (www.RCSB.org), EC numbers are assigned based on UniProtKB, GenBank, the Kyoto Encyclopedia of Genes and Genomes (KEGG), and authors’ descriptions. Among 518 structures, 91 of them, such as metallo-β-lactamase, have been excluded, because they were reported to have other enzymatic roles that did not describe CEH activity. Theoretical structures (e.g., 1A20) are not analyzed. CEHs not annotated with an EC number, such as carbohydrate esterase SmAcE1, are also beyond the scope of this review, although they do have esterase activity [18]. As a result, 424 structures of 136 CEHs from 52 genera and metagenomic (uncultured and unidentified) samples were selected for the analysis, and 136 CEHs were classified based on EC numbers (Table 1). When several crystal structures were available in one enzyme (several PDB codes), the first reported structure was used as a CEH structure. Their sequences containing tag or mutation were analyzed after being recovered to wild type sequences. Exceptionally, the sequences were not recovered when the mutations in the active sites of CEHs induced a gain of function. The majority of the crystal structures were identified in 1.0–3.0 Å resolution, and even extremely high resolution (below 1 Å) cases were also reported (Figure 1A). The major bacterial sources of CEHs in this analysis are Pseudomonas (9.6%), Bacillus (8.8%) and Geobacillus (7.4%), as shown in Figure 1B. Ten CEHs (7.4%) originate in metagenomic samples.

3. Classification of CEHs Based on Substrates

CEHs can be categorized based on substrates, which can be identified by EC number as shown in Table 1. CEHs using lipids as substrates are phospholipases (only PLA1 [EC 3.1.1.32] and PLA2 [EC 3.1.1.4]), carboxylesterases (EC 3.1.1.1), lysophospholipases (EC 3.1.1.5) and triacylglycerol lipases (also called true lipases [EC 3.1.1.3]). However, bacterial lipolytic CEHs are not sensitive to alcohol-group substrates, but to acyl chain length. Carboxylesterase (EC 3.1.1.1) also cleaves carboxylic esters, but the length of the acyl chain is much shorter than the substrates of lipases. Enzyme kinetics and lid structure can also be used to distinguish between lipase and carboxylesterase [178]. In general, lipase contains a lid that covers the active site. However, these approaches are controversial, as some carboxylesterases, such as a lid-containing carboxylesterase, contain a lid similar to those found in lipases [72,179,180].
CEHs that use carbohydrates as substrates are called carbohydrate esterases [181,182]. Carbohydrate esterases that are active on xylan, cutin, and pectin are known as acetylxylan esterase (EC 3.1.1.72), cutinase (EC 3.1.1.74) and pectinesterase (EC 3.1.1.11), respectively, and the acyl chain, most often a member of the acetyl group, is removable in the monomeric and polymeric forms of carbohydrates.

4. Classification Based on Localization

CEHs are distributed from extracellular to cytosolic regions. Ninety of 136 CEHs are localized in the cytosolic region, 43 CEHs have signal peptides for secretion and the remaining three CEHs are transmembrane proteins (Figure 2). Outer membrane phospholipase As (OMPLAs) from Escherichia coli (representative PDB code: 1FW2) [107,108,109] and Salmonella typhi (PDB code: 5DQX) span membranes. Autotransporter EstA from Pseudomonas aeruginosa (PDB code: 3KVN) is another outer membrane-spanning protein [50]. Representatively, pectin methylesterase from Dickeya dadantii (PDB code: 2NSP) with a signal sequence at the N-termini, is a representative secretary CEH for the bacterial invasion of plant tissues [115]. LipA from Xanthomonas oryzae (representative PDB code: 3H2G) [28,183], lipase from Geobacillus zalihae (PDB code: 2DSN) [80] and phospholipase A2 from Streptomyces violaceoruber (PDB code: 1LWB) [184] have been physiologically verified as secreted proteins, as they can be isolated from culture media.

5. Classification Based on the Active Site Residues

Ser-His-Asp in the catalytic triad, which works as a nucleophile, a base and an acid, respectively, is necessary for hydrolysis [189]. In addition to the conventional catalytic triad (Ser-His-Asp), various types of nonconventional triads and dyads have been reported [190]. Gariev et al. classified hydrolases based on components in active sites to produce hierarchical four-digit layers and a web-based database (http://www.enzyme.chem.msu.ru/hcs) [191]. In the hydrolysis reaction, Oγ in Ser, Sγ in Cys, Oγ1 in Thr, Oδ in Asp and O in the water molecule are the nucleophiles, attacking carbonyl carbon in carboxylic ester bonds. The base, usually His residue, deprotonates the nucleophile, and increases the activity of these nucleophiles. The acid stabilizes the position of the base, and assists the function of base to the nucleophile. Along with catalytic residues, the oxyanion hole plays a key role in stabilizing transition states. CEHs can be classified into several groups based on consensus sequences encompassing their active site residues. Here, we divide CEHs into four groups (groups 1 to 4), based on catalytic residues. Each group is divided into sub-groups according to motifs and conserved residues in the catalytic domain. The information of key residues in each group is provided using the representative structures in Figure 3.

5.1. Ser Hydrolases (Group 1)

A sequence analysis of CEHs revealed that many CEHs contain the catalytic triad Ser-His-Asp. We defined group 1 CEHs as those containing the catalytic triad with Ser serving as a nucleophile. Based on motifs containing catalytic Ser, CEHs can be classified into several groups, such as GXSXG, GDSX, SXXK and YTQ/HXSNG groups (underlined residues are nucleophiles). Among them, the GXSXG group is the most common, containing 88 esterases among 136 CEHs.
  • Group 1-1
Group 1-1 is the most abundant CEH, and contains the GXSXG motif, along with the AXSXG and GXSXXG variants. The GXSXG motif is localized in the loop region, and forms a catalytic triad with Asp and His in other loops in the C-terminal region. The GXSXXG motif is found in glucuronoyl esterase from Solibacter usitatus (PDB code: 6GRY) [42], carbohydrate esterase 15 from a marine metagenome (PDB code: 6EHN) [41], cocaine esterase from Rhodococcus spp. (representative PDB code: 3I2K) [29], and alpha-amino acid ester hydrolases from Acetobacter pasteurianus (representative PDB code: 2B9V) [148] and Xanthomonas citri (PDB code: 1MPX) [147].
Artificial dienelactone hydrolases were obtained through protein engineering, including the mutation of C123S in the GXCXG motif of carboxymethylenebutenolidase from Pseudomonas knackmussii (representative PDB code: 4U2B) [153] and from Pseudomonas putida (representative PDB codes: 1ZI8) [151]. Introducing the GXSXG motif reportedly enables the production of an artificial dienelactone hydrolase [193]. The other two triad components are most frequently identified as His-Asp by order in CEHs with GXSXG motifs. Exceptionally, Glu is positioned instead of Asp in the following six CEHs: naproxen esterase from Bacillus carboxylesterases cleaving naproxen ester (PDB code: 4CCW) [53], carboxylesterase CesB from Bacillus sp (PDB code: 4CCY) [53], Est1 from Hungatella hathewayi (PDB code: 5A2G) [37], pNB esterase from Bacillus subtilis (PDB code: 1C7I) [19], a putative carboxylesterase from B. subtilis (PDB code: 2R11) and metagenomic Est5 (PDB code: 3FAK) [26]. Their catalytic triad therefore is composed of Ser-His-Glu. In this group, the GGGX, GX and Y motifs, which are located mostly in the N-terminal region of a CEH, are involved in forming the oxyanion hole [1,194]. In the GGGX motif, most oxyanion hole components are positioned at the second Gly and third Gly residues. In CEHs containing the GX or Y motifs, with residue X in the GX motif or Tyr in the Y motif, an oxyanion hole forms with the second X in the active site GXSXG motif. In addition, GGAX (representative PDB code: 4V2I [58] and 3DOH [49]) and GAGX (representative PDB code: 1C7I [19], 5A2G [37] and 4C89 [52]) motifs have also been also reported. Y-motif-containing CEHs have been reported in amino acid ester hydrolases (PDB code: 2B9V [148] and 1MPX [147]), and cocaine esterase (representative PDB code: 3I2K [29]). Catalytic His and Asp/Glu are typically positioned with the 20–30 amino acid gap in the order of Asp-His. However, in chemotaxis methylesterase (CheB) from Salmonella typhimurium (PDB code: 1CHD), catalytic His190 and D286 are positioned in the reverse order, with 95 unique amino acid gaps [13]. Important residues in this group are shown in Figure 3A.
  • Group 1-2
Group 1-2 includes the GDSX motif-containing CEHs (called the GDSL family), in which catalytic Ser is localized close to the N-terminus in a hydrolase domain [196]. According to previous analysis, in the GDSL family, sequence consensus blocks (Block I, II, III and V) contain the functionally important residues Ser, Gly, Asn and His, and thus named SGNH hydrolases, as shown in Figure 4. Catalytic Ser is found in Block I, the oxyanion hole components Gly and Asn are located in Blocks II and III, and a general base known as His exists in Block V (Each residue is marked with an asterisk in Figure 4). The general base His and a general acid Asp form a DXXH motif near the C-terminus of the hydrolase domains. As a rare group, xylan esterase from Cellvibrio japonicus (CjCE2A, PDB code: 2WAA) has Asp789 and His791 in a DXH motif (Figure 4) [22].
Moreover, catalytic dyads lacking Asp in the DXXH motif are also identified in lipase from Streptomyces rimosus (PDB code: 5MAL) [39] and esterase from Streptomyces scabies (PDB code: 1ESC) as shown in Figure 4 [20]. In their structures, Asp residues in the catalytic triad are replaced by nonfunctional Asn in the S. rimosus lipase and Trp in the S. scabies esterase, in which both Asn and Trp only stabilize the orientation of the catalytic His instead of playing a role as acids. In phospholipase A1 from S. albidoflavus (PDB code: 4HYQ), its sequence shows a conserved DXXH motif of Block V, but the 3D structure reveals that the Ser-His dyads form because of the position of Asp in the DXXH motif is not proper for its function as a general acid [141]. In oxyanion hole formation, aryl esterase from Mycobacterium smegmatis (representative PDB code: 2Q0Q) has Ala instead of Gly in Block II, but its function is similar [65]. Important residues in this group are shown in Figure 3B.
  • Group 1-3
In group 1-3, the SXXK motif is commonly found in peptidases including β-lactamase [197], and also in family VIII lipases [198,199]. The lipases in family VIII have β-lactamase-like structures, but they usually have carboxylic esterase activity, not β-lactamase activity. Ser71, Lys74 and Tyr160 compose catalytic triads in the EstSRT1, family VIII lipase from the metagenome (PDB code: 5GMX), as in Figure 3C [61]. Another family VIII lipase, EstU1 from the metagenome (PDB code: 4IVK), also has the SXXK motif in which Ser64 and Lys67 form a catalytic triad with Tyr150 [55].
  • Group 1-4
Group 1-4 CEHs contain YTQ and HXSNG motifs. OMPLAs from E. coli (representative PDB code: 1FW2) [108] and from S. typhi (PDB code: 5DQX) belong to this group. The YTQ motif is essential for dimerization of OMPLAs in the membrane, and the HXSNG motif is critical for hydrolase activity [200,201]. In E. coli OMPLA, His142 and Ser144 in the HXSNG motif compose a catalytic triad with Asn156, and consecutive Asn145 and Gly146 of the motif are components of an oxyanion hole, as shown in Figure 3D. S. typhi OMPLA also has a Ser164-His162-Asn176 catalytic triad and an oxyanion hole formed by Asn165 and Gly166 with the YTQ motif (residues 112–114).

5.2. Aspartyl Hydrolases (Group 2)

Group 2 CEHs are aspartyl hydrolases containing an Asp-Asp dyad with a nucleophilic Asp and a basic Asp, as described in Figure 3E. Epoxide hydrolases [202], and glycosyl hydrolases [203] belong to Asp hydrolases. The Asp-Asp catalytic dyad is also identified in pectin methylesterases (orange squares in Figure 2). Pectin methylesterase A proteins from Dickeya chrysanthemi (representative PDB code: 1QJV) [114] and D. dadanti (PDB code: 2NSP) [115] contain the GXSXXG motif, although Ser in this motif does not work as a nucleophile. In these enzymes, commonly, Asp199-Asp178 form a catalytic dyad, Gln177 is the oxyanion hole-forming residue, and Arg267 and Trp269 are involved in pectin binding [114,115]. The hydrolysis reaction proceeds metal-independently, and without a nucleophilic water molecule. In a similar manner, in pectin methylesterase (or carbohydrate esterase family VIII) from Yersinia enterocolitica (PDB code: 3UW0), Asp199 and Asp177 work as a nucleophile and a general acid/base, respectively [116]. In this enzyme, Arg264 and Trp266 are functionally conserved as a pectin-binding site, and Gln176 belongs to an oxyanion hole [116].

5.3. Metal-independent Hydrolase with a Nucleophilic Water (Group 3)

Group 3 CEHs are nonconventional hydrolases, lacking typical nucleophilic residues such as Ser/Thr/Cys, but containing a water molecule that functions as a nucleophile. The water molecule is activated by general base His (Group 3-1, 3-3 and 3-4), Asp (Group 3-5) and Gln (Group 3-2) residues without metal coordination or the assistance of other cofactors.
  • Group 3-1
In the peptidyl-tRNA hydrolase from Vibrio cholerae (representative PDB code: 4Z86) belonging to Group 3-1 CEHs, Asn14, His24, Asn72, Asp97 and Asn118 are functionally important [136]. His24 and Asp97 work as a general base and an acid, respectively. Asn72 and Asn118 form an oxyanion hole, and two Asp residues at 14 and 118 stabilize tRNA-binding in its active site. As described in Figure 5, catalytic His-Asp (denoted by an asterisk), two oxyanion hole-forming Asn residues (O-marked) and Asn involved in substrate binding (S-marked) are also conserved in other peptidyl-tRNA hydrolases, such as those from Mycobacterium tuberculosis (PDB code: 2Z2I) [124], M. smegmatis (PDB code: 3P2J), E. coli (PDB code: 3VJR, 2PTH, and 3OFV) [123], S. typhimurium (PDB code: 4P7B) [132], Acinetobacter baumannii (PDB code: 4LWQ), Pseudomonas aeruginosa (PDB code: 4FYJ) [131], Francisella tularensis (PDB code: 3NEA) [125], Streptococcus pyogenes (PDB code: 4QT4) [133], Staphylococcus aureus (PDB code: 4YLY) [135] and Burkholderia thailandensis (PDB code: 3V2I) [127]. In the Thermus thermophilus peptidyl-tRNA hydrolase (PDB code: 5ZX8), Arg103 plays a role in substrate binding and oxyanion formation similar to Asn118 in V. cholerae (Figure 5) [138]. Important residues in this group are shown in Figure 3F.
  • Group 3-2
Another type of peptidyl-tRNA hydrolases belonging to the RF-1 family is defined as group 3-2 CEHs in this review (Figure 3G). A common feature of RF-1 family peptidyl-tRNA hydrolase is a GGQ motif. Gln in this motif works as a base that stabilizes nucleophilic water, and also plays a key role in interactions with tRNA [204]. The Oε1 of Gln28 in YaeJ, a peptidyl-tRNA hydrolase from E. coli (PDB code: 4V95) and the nucleophilic water molecule, form a hydrogen bond [134]. The water molecule attacks the carbonyl carbon of the carboxylic ester bond between A76 in tRNA and the peptide.
  • Group 3-3
Secreted phospholipase A2, defined as a group 3-3 CEH, is considered a metal-independent hydrolase using His-Asp as a base-and-acid dyad in the catalytic site to stabilize the nucleophilic water, as shown in Figure 3H [205]. The secreted PLA2 from S. violaceoruber (PDB code: 1LWB) uses His64 and Asp85 as residues in the catalytic dyad and Tyr68 to encourage substrate-binding [106]. In the Ca2+-free form of PLA2 (PDB code: 1LWB), W260 is the inferred nucleophilic water, attacking sn-1 carbonyl carbon. In the Ca2+-binding form of PLA2 (PDB code: 1KP4), the calcium ion, coordinated by Oδ1 of Asp43, O of Leu44, Oδ2 of Asp65 and three water molecules (W201, W202, and W203), induces a hydrogen bond network in a substrate binding pocket. In this enzyme, the water molecule (W256) is regarded as a nucleophile and attacks the sn-2 carbonyl carbon of the substrate [105,184].
  • Group 3-4
6-phosphogluconolactonases from M. smegmatis (representative PDB code: 3OC6) [140] belonging to group 3-4 CEHs contains His151-Glu149 as a catalytic dyad with nucleophilic water (Figure 3I). Similarly, 6-phosphogluconolactonases from T. tuberclosis (PDB code:3ICO) [139] contains His152-Glu150 as a catalytic dyad, and Thermotoga maritima 6-phosphogluconolactonase (PDB code: 1VL1) contains the conserved dyad His138-Asp136.
  • Group 3-5
Unlike hydrolases described above, the Asp residue of the enzyme belonging to group 3-5 CEH works as a general base on nucleophilic water (Figure 3J). For example, LigI from Sphingomonas paucimobilis (PDB code: 4D8L, light green with a black outline in Figure 2) belonging to the amidohydrolase superfamily has lactonase activity using Asp248 as a catalytic base without metal–ion coordination [154]. The water molecule forming a hydrogen bond with Asp248 attacks carbonyl carbon in the lactone group as a nucleophile. His31 and His180 form an oxyanion hole, and His33 contributes to the lactonase reaction by stabilizing the tetrahedral intermediate.

5.4. Metal-dependent Metallohydrolases (Group 4)

  • Group 4-1
In the group 4-1 CEH, the HXHXDH motif is involved in Zn2+-binding [206], and His and Asp residues in this motif mainly coordinate the metal ion that stabilizes a nucleophilic water. In N-acyl homoserine lactone hydrolase from Bacillus thuringiensis (PDB code: 2A7M) [21,166], the HXHXDH motif composed of His104-X-His106-X-Asp108-His109 coordinates two zinc ions with additional Asp191, His169 and His235, as in Figure 3K. The first Zn2+ is coordinated by His104, His106 and His169, and the second Zn2+ is coordinated by Asp108, His109, Asp191 and His235. According to a known mechanism, one water molecule coordinated by two zinc ions works as a nucleophile, and Tyr194 works as an acid in hydrolysis [24,164]. In 4-pyridoxolactonase from Mesorhizobium loti (PDB code: 4KEP), His96, His98, Asp100 and His101 are HXHXDH motif components. Asp100, His101, Asp207 and His252 coordinate the first Zn2+, and His96, His98, His185 and Asp207 coordinate the second Zn2+. In contrast to the HXHXDH-containing group 4-1 CEHs, lactonase UlaG from E. coli (representative PDB code: 2WYM) has only one Mn2+ at the second Zn2+ position, although it has the HXHXDH motif (Figure 6) [23]. HXHXDH motif-containing CEHs are highlighted using the pale blueish-green diamond in Figure 2.
  • Group 4-2
This group of CEHs does not contain an HXHXDH motif. For example, the de-O-acetylases family CE4 from Streptomyces lividans (PDB code: 2CC0) contains a single Zn2+ coordinated by Asp13, His62 and His66 (Figure 3L). For hydrolysis, the nucleophilic water molecule that binds to Zn2+ attacks the carbonyl carbon of substrates, and His62 stabilizes the carbonyl group of the substrate by forming an oxyanion hole [157]. Amidohydrolase from Mycoplasma synoviae (PDB code: 3OVG) does not have an HXHXDH motif, but two Zn2+-binding sites. His186 and His214 coordinate the first Zn2+, and His24, His26 and Asp272 coordinate the second Zn2+. It also has an His26-Asp68 catalytic dyad.

6. Classification Based on Tertiary Structure

The CATH database (for class, homology, architecture, homologous superfamily; http://www.cathdb.info) is designed to predict protein function from structures [207]. In the most recent version of CATH (v4.2.0), a total of 150,885 PDB-deposited structures are classified into four classes, 41 architectures, 1391 topologies and 6119 homologous superfamilies. According to CATH-based analysis, 107 structures of CEHs are classified into an α/β/α sandwich architecture (Figure 7A).
Among the remaining 29 CEH structures, 16 belong to various architectures, but 13 CEHs (representative PDB codes: 4V95 [134], 5AH1 [100], 5GMX [61], 5H3B [62], 5H6B [102], 5UGQ [40], 5XPX, 5YAL [160], 6A12 [103], 6AAE [63], 6EHN [41], 6GRY [42] and 6QGB [177]) are not classified under the CATH database classification. Secreted PLA2 from S. violaceoruber (PDB code: 1LWB) forms an up–down bundle structure with only helices (Figure 7B). Gluconolactonase from Xanthomonas campestris (PDB code: 3DR2) is composed of six blades in a propeller structure (Figure 7C) [117], and similarly, the tertiary structure of 6-phosphogluconolactonase from Klebsiella pneumoniae (PDB code: 6NAU) is a 7-blade propeller (Figure 7D). E. coli OMPLA (representative PDB code: 1FW2) [108] and S. typhi OMPLA (PDB code: 5DQX) form β-barrel structures, and contain a catalytic triad at a dimeric interface in the β-barrel transmembrane area (Figure 7E). Lactone hydrolases (PDB code: 4KEP, 2WYM and 2A7M) have an α/β/β/α 4-layered fold (Figure 7F). The overall shapes of amidohydrolase LigI from Sphingobium (PDB code: 4D8L) [154], acetyl xylan esterases from S. lividans (PDB code: 2CC0) [157] and amidohydrolase from M. synoviae (PDB code: 3OVG) are α/β barrels (Figure 7G). Apo and substrate-binding forms of pectin methylesterase A from D. chrysanthemi (representative PDB code: 1QJV and 2NSP) [114,115] and Y. enterocolitica (PDB code: 3UW0) [116] are right-handed parallel β helices, called 3-solenoids (Figure 7H).
The CEH structures can be also classified according to the CASTLE database (https://castle.cbe.iastate.edu) in which bacteria, archaea and eukaryote CEHs are clustered into three clans of α/β hydrolase based on the number and order of β-sheets; eight β-sheets arranged with 1-2-4-3-5-6-7-8 order (Clan A, Figure 8A), five β-sheets arranged with 2-1-3-4-5 order (Clan B, Figure 8B) and seven β-sheets arranged with 1-3-2-4-5-6-7 order (Clan C, Figure 8C) β-sheet sequences. In addition, there are two non-α/β hydrolase clans in the CEHs; a six-bladed β-propeller (Clan D) and three α-helix bundle (Clan E) [208]. Initially, the α/β hydrolase fold was described as α helices surrounding eight central β sheets [209], but it was extended to involve variations including a smaller fold composed of five central β sheets and sandwiched α helices [210]. Among 136 CEHs, 46 structures (33.8%) are turned to have an α/β hydrolase fold in Clans A, B and C (Table 2). Only single structure (PDB code: 3DR2) belongs to Clan D, and no structure is assigned to Clan E.

7. Substrate-Structure Connection of CEHs

The development of sequencing technology enables the identification of new enzymes from various organisms, including bacteria, and even from the metagenome [212,213,214,215,216]. Functional annotation of those enzymes has been followed by direct or indirect approaches, such as computational sequence/structure analysis and comparison with characterized enzymes [217]. 3D-Fun [218], MOLMAP descriptors [219], ECAssigner [220], EC-blast [221] and the recently released DeepEC [222] have been developed to find links between the EC number and enzyme structures.
Accuracy in predicting the functional assignment of enzymes has improved, but as Gerlt’s statistical analysis shows, only 0.63% of proteins by computationally automated annotation have been manually assigned to an EC class [223]. The Ferrer group with the Industrial Applications of Marine Enzymes Consortium (INMARE) attempted to predict enzyme-substrate correlation using the results of a high-throughput assay (145 ester hydrolases sequences and 96 substrates) [224]. Based on this analysis, enzyme promiscuity was proposed. It has been consistently reported that one enzyme can be assigned to multiple classes from primary to quaternary orders of EC class [225]. For example, E. coli lysophospholipase L1 (representative PDB code: 1IVN) is assigned to be a lysophospholipase (EC 3.1.1.5), but shows arylesterase (EC 3.1.1.2), palmitoyl-CoA hydrolase (EC 3.1.2.2), acyl-[acyl-carrier-protein] hydrolase (EC 3.1.2.14) and protease (EC 3.4.21.-) activities (Table 1).

8. Physiological Functions of CEHs

Bacterial CEHs participate in various phycological processes, such as signaling pathways, protein synthesis and offensive-defensive responses. Bacteria recognize ligands and respond, moving toward or repelling from ligands through flagellar movement [226]. The response to ligands in surrounding environment conditions occurs using chemotaxis receptors, which are regulated by methyl transferase CheR and hydrolase CheB [14]. CheB removes methyl groups from methylated glutamate residues in the cytoplasm and inactivates receptors. When recognizing high cell density, the Gram-negative bacteria release autoinducers, such as N-acyl homoserine lactone (AHL) derivatives that induce virulence gene expression by self-receptors [227,228,229,230]. To remove autoinducing signals, they release AHL lactonase, inactivating and degrading AHL by cleavage lactone rings [16,231]. From the function of lactonase, the regulation of AHL through lactonases has been suggested for medicinal applications [232]. PLAs, including secreted, cytosolic and membrane-integrated forms, work in nutrient digestion, inflammation and intra-signaling cascades [233,234]. Mono- and diacylglycerol lipases are important for lipid metabolism in bacteria [235]. Short-chain fatty acids produced by CEHs in gut microbiota regulate host signaling and metabolic systems [236,237]. In bacterial ribosomal machinery, CEHs also resolve non-stop translation problems by cleaving peptide- or amino acid-conjugated tRNA [238,239].

9. Industrial Applications of CEHs

CEHs are widely used in industrial fields because their endogenous characteristics, such as the substrate specificity and stability of structures [240,241,242,243]. Moreover, it has been used for green chemistry, since reactions using biocatalysts employ less steps for the chemical synthesis and produce less harmful wastes during the reaction compared to reactions using chemical catalysts. Furthermore, biocatalysts cover wide substrates with fewer unexpected side products. Carbohydrate esterases have been widely used in animal- and plant-oriented biomass degradation, in the production of biofuels such as ethanol [244,245,246] and in coffee fermentation to improve flavor and taste [247]. Lipases or carboxylesterases are useful for providing the flavor of yogurt and cheese [248,249]. Polyethylene terephthalate hydrolases are suggested as potent biocatalysts for waste management [250]. These characteristics of CEHs can be further improved through structure-based engineering and directed evolution [251]. Additionally, enzyme reusability through immobilization methods using nanostructure [252], cross-linking [253], encapsulation [254] and entrapment [255], are also being considered, and the immobilized enzymes make it possible to reduce the high initial costs associated with enzyme preparation [256].

10. Perspectives on Identifying More CEHs and Their Functions

CEHs are one of the largest group of enzymes, comprising lipases, carboxylesterases, carbohydrate esterases, peptidyl-tRNA hydrolases and lactonases. They target not only carboxylic ester bonds, but also amide (EC 3.5.1.-), thioester (EC 3.1.2.-) and peroxide (EC 1.-.-.-) bonds rarely. The promiscuity of CEHs should be considered when identifying enzyme characteristics.
For these purposes, the factors that induce promiscuity should be identified and the range of specificity chosen. Collecting functional and structural genomics data and linking these large datasets should be done systemically. Amin et al. mentioned the importance of motif in structures when defining enzyme function [257], and Kingsley et al. used various kinetic models to confirm that substrate tunnels in enzymes affect substrate specificity [258]. However, in contrast to an abundance of structural information, fewer structures have been properly matched to biochemical data or functional annotation in the PDB. Moreover, much is unknown about orphan reactions, in which the substrate and the products are already known, but the responsible enzymes are not [259,260,261]. All existing information should be gathered and used to fill in the remaining blanks to generate a full understanding at the molecular level and draw a heuristic map of the biochemical universe.

Funding

This research was funded by grant from the National Research Foundation of Korea, grant number 2017M3A9E4078553.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Lian, J.; Nelson, R.; Lehner, R. Carboxylesterases in lipid metabolism: From mouse to human. Protein Cell 2018, 9, 178–195. [Google Scholar] [CrossRef]
  2. Soreq, H.; Seidman, S. Acetylcholinesterase--new roles for an old actor. Nat. Rev. Neurosci. 2001, 2, 294–302. [Google Scholar] [CrossRef] [PubMed]
  3. Giudice, E.; Gillet, R. The task force that rescues stalled ribosomes in bacteria. Trends Biochem. Sci. 2013, 38, 403–411. [Google Scholar] [CrossRef] [PubMed]
  4. Pilling, J.; Willmitzer, L.; Fisahn, J. Expression of a Petunia inflata pectin methyl esterase in Solanum tuberosum l. Enhances stem elongation and modifies cation distribution. Planta 2000, 210, 391–399. [Google Scholar] [CrossRef] [PubMed]
  5. Wu, H.C.; Bulgakov, V.P.; Jinn, T.L. Pectin methylesterases: Cell wall remodeling proteins are required for plant response to heat stress. Front. Plant Sci. 2018, 9, 1612. [Google Scholar] [CrossRef]
  6. Schafer, M.; Fischer, C.; Meldau, S.; Seebald, E.; Oelmuller, R.; Baldwin, I.T. Lipase activity in insect oral secretions mediates defense responses in arabidopsis. Plant Physiol. 2011, 156, 1520–1534. [Google Scholar] [CrossRef]
  7. Biely, P. Microbial carbohydrate esterases deacetylating plant polysaccharides. Biotechnol. Adv. 2012, 30, 1575–1588. [Google Scholar] [CrossRef]
  8. Gendrin, C.; Contreras-Martel, C.; Bouillot, S.; Elsen, S.; Lemaire, D.; Skoufias, D.A.; Huber, P.; Attree, I.; Dessen, A. Structural basis of cytotoxicity mediated by the type III secretion toxin exou from Pseudomonas aeruginosa. PLoS Pathog. 2012, 8, e1002637. [Google Scholar] [CrossRef]
  9. Sitkiewicz, I.; Stockbauer, K.E.; Musser, J.M. Secreted bacterial phospholipase A2 enzymes: Better living through phospholipolysis. Trends Microbiol. 2007, 15, 63–69. [Google Scholar] [CrossRef]
  10. Phillips, R.M.; Six, D.A.; Dennis, E.A.; Ghosh, P. In vivo phospholipase activity of the Pseudomonas aeruginosa cytotoxin exou and protection of mammalian cells with phospholipase A2 inhibitors. J. Biol. Chem. 2003, 278, 41326–41332. [Google Scholar] [CrossRef]
  11. Armendáriz-Ruiz, M.; Rodríguez-González, J.A.; Camacho-Ruíz, R.M.; Mateos-Díaz, J.C. Carbohydrate esterases: An overview. In Lipases and Phospholipases: Methods and Protocols; Sandoval, G., Ed.; Springer: New York, NY, USA, 2018; pp. 39–68. [Google Scholar]
  12. Casas-Godoy, L.; Duquesne, S.; Bordes, F.; Sandoval, G.; Marty, A. Lipases: An overview. In Lipases and Phospholipases: Methods and Protocols; Sandoval, G., Ed.; Humana Press: Totowa, NJ, USA, 2012; pp. 3–30. [Google Scholar]
  13. West, A.H.; Martinez-Hackert, E.; Stock, A.M. Crystal structure of the catalytic domain of the chemotaxis receptor methylesterase, CheB. J. Mol. Biol. 1995, 250, 276–290. [Google Scholar] [CrossRef] [PubMed]
  14. Djordjevic, S.; Goudreau, P.N.; Xu, Q.; Stock, A.M.; West, A.H. Structural basis for methylesterase CheB regulation by a phosphorylation-activated domain. Proc. Natl. Acad. Sci. USA 1998, 95, 1381–1386. [Google Scholar] [CrossRef] [PubMed]
  15. Das, G.; Varshney, U. Peptidyl-tRNA hydrolase and its critical role in protein biosynthesis. Microbiology 2006, 152, 2191–2195. [Google Scholar] [CrossRef] [PubMed]
  16. Dong, Y.H.; Xu, J.L.; Li, X.Z.; Zhang, L.H. AiiA, an enzyme that inactivates the acylhomoserine lactone quorum-sensing signal and attenuates the virulence of Erwinia carotovora. Proc. Natl. Acad. Sci. USA 2000, 97, 3526–3531. [Google Scholar] [CrossRef]
  17. Berman, H.M.; Westbrook, J.; Feng, Z.; Gilliland, G.; Bhat, T.N.; Weissig, H.; Shindyalov, I.N.; Bourne, P.E. The protein data bank. Nucleic Acids Res. 2000, 28, 235–242. [Google Scholar] [CrossRef]
  18. Oh, C.; Ryu, B.; Yoo, W.; Nguyen, D.; Kim, T.; Ha, S.-C.; Kim, T.; Kim, K. Identification and crystallographic analysis of a new carbohydrate acetylesterase (SmAcE1) from Sinorhizobium meliloti. Crystals 2018, 8, 12. [Google Scholar] [CrossRef]
  19. Spiller, B.; Gershenson, A.; Arnold, F.H.; Stevens, R.C. A structural view of evolutionary divergence. Proc. Natl. Acad. Sci. USA 1999, 96, 12305–12310. [Google Scholar] [CrossRef]
  20. Wei, Y.; Schottel, J.L.; Derewenda, U.; Swenson, L.; Patkar, S.; Derewenda, Z.S. A novel variant of the catalytic triad in the Streptomyces scabies esterase. Nat. Struct. Biol. 1995, 2, 218–223. [Google Scholar] [CrossRef]
  21. Liu, D.; Lepore, B.W.; Petsko, G.A.; Thomas, P.W.; Stone, E.M.; Fast, W.; Ringe, D. Three-dimensional structure of the quorum-quenching N-acyl homoserine lactone hydrolase from Bacillus thuringiensis. Proc. Natl. Acad. Sci. USA 2005, 102, 11882–11887. [Google Scholar] [CrossRef]
  22. Montanier, C.; Money, V.A.; Pires, V.M.; Flint, J.E.; Pinheiro, B.A.; Goyal, A.; Prates, J.A.; Izumi, A.; Stalbrand, H.; Morland, C.; et al. The active site of a carbohydrate esterase displays divergent catalytic and noncatalytic binding functions. PLoS Biol. 2009, 7, e71. [Google Scholar] [CrossRef]
  23. Garces, F.; Fernandez, F.J.; Montella, C.; Penya-Soler, E.; Prohens, R.; Aguilar, J.; Baldoma, L.; Coll, M.; Badia, J.; Vega, M.C. Molecular architecture of the Mn2+-dependent lactonase ulag reveals an RNase-like metallo-beta-lactamase fold and a novel quaternary structure. J. Mol. Biol. 2010, 398, 715–729. [Google Scholar] [CrossRef] [PubMed]
  24. Momb, J.; Wang, C.; Liu, D.; Thomas, P.W.; Petsko, G.A.; Guo, H.; Ringe, D.; Fast, W. Mechanism of the quorum-quenching lactonase (AiiA) from Bacillus thuringiensis. 2. Substrate modeling and active site mutations. Biochemistry 2008, 47, 7715–7725. [Google Scholar] [CrossRef] [PubMed]
  25. Nam, K.H.; Kim, M.Y.; Kim, S.J.; Priyadarshi, A.; Kwon, S.T.; Koo, B.S.; Yoon, S.H.; Hwang, K.Y. Structural and functional analysis of a novel hormone-sensitive lipase from a metagenome library. Proteins 2009, 74, 1036–1040. [Google Scholar] [CrossRef] [PubMed]
  26. Nam, K.H.; Kim, M.Y.; Kim, S.J.; Priyadarshi, A.; Lee, W.H.; Hwang, K.Y. Structural and functional analysis of a novel EstE5 belonging to the subfamily of hormone-sensitive lipase. Biochem. Biophys. Res. Commun. 2009, 379, 553–556. [Google Scholar] [CrossRef] [PubMed]
  27. Nam, K.H.; Kim, S.J.; Priyadarshi, A.; Kim, H.S.; Hwang, K.Y. The crystal structure of an HSL-homolog EstE5 complex with PMSF reveals a unique configuration that inhibits the nucleophile Ser144 in catalytic triads. Biochem. Biophys. Res. Commun. 2009, 389, 247–250. [Google Scholar] [CrossRef]
  28. Aparna, G.; Chatterjee, A.; Sonti, R.V.; Sankaranarayanan, R. A cell wall-degrading esterase of Xanthomonas oryzae requires a unique substrate recognition module for pathogenesis on rice. Plant Cell 2009, 21, 1860–1873. [Google Scholar] [CrossRef]
  29. Narasimhan, D.; Nance, M.R.; Gao, D.; Ko, M.C.; Macdonald, J.; Tamburi, P.; Yoon, D.; Landry, D.M.; Woods, J.H.; Zhan, C.G.; et al. Structural analysis of thermostabilizing mutations of cocaine esterase. Protein Eng. Des. Sel. 2010, 23, 537–547. [Google Scholar] [CrossRef]
  30. Brim, R.L.; Nance, M.R.; Youngstrom, D.W.; Narasimhan, D.; Zhan, C.G.; Tesmer, J.J.; Sunahara, R.K.; Woods, J.H. A thermally stable form of bacterial cocaine esterase: A potential therapeutic agent for treatment of cocaine abuse. Mol. Pharmacol. 2010, 77, 593–600. [Google Scholar] [CrossRef]
  31. Alterio, V.; Aurilia, V.; Romanelli, A.; Parracino, A.; Saviano, M.; D’Auria, S.; De Simone, G. Crystal structure of an S-formylglutathione hydrolase from Pseudoalteromonas haloplanktis tac125. Biopolymers 2010, 93, 669–677. [Google Scholar]
  32. Lai, K.K.; Stogios, P.J.; Vu, C.; Xu, X.; Cui, H.; Molloy, S.; Savchenko, A.; Yakunin, A.; Gonzalez, C.F. An inserted alpha/beta subdomain shapes the catalytic pocket of Lactobacillus johnsonii cinnamoyl esterase. PLoS ONE 2011, 6, e23269. [Google Scholar] [CrossRef]
  33. Narasimhan, D.; Collins, G.T.; Nance, M.R.; Nichols, J.; Edwald, E.; Chan, J.; Ko, M.C.; Woods, J.H.; Tesmer, J.J.; Sunahara, R.K. Subunit stabilization and polyethylene glycolation of cocaine esterase improves in vivo residence time. Mol. Pharmacol. 2011, 80, 1056–1065. [Google Scholar] [CrossRef] [PubMed]
  34. Filippova, E.V.; Weston, L.A.; Kuhn, M.L.; Geissler, B.; Gehring, A.M.; Armoush, N.; Adkins, C.T.; Minasov, G.; Dubrovska, I.; Shuvalova, L.; et al. Large scale structural rearrangement of a serine hydrolase from Francisella tularensis facilitates catalysis. J. Biol. Chem. 2013, 288, 10522–10535. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Schiefner, A.; Gerber, K.; Brosig, A.; Boos, W. Structural and mutational analyses of Aes, an inhibitor of MalT in Escherichia coli. Proteins 2014, 82, 268–277. [Google Scholar] [CrossRef] [PubMed]
  36. Dou, S.; Kong, X.D.; Ma, B.D.; Chen, Q.; Zhang, J.; Zhou, J.; Xu, J.H. Crystal structures of Pseudomonas putida esterase reveal the functional role of residues 187 and 287 in substrate binding and chiral recognition. Biochem. Biophys. Res. Commun. 2014, 446, 1145–1150. [Google Scholar] [CrossRef]
  37. Perz, V.; Hromic, A.; Baumschlager, A.; Steinkellner, G.; Pavkov-Keller, T.; Gruber, K.; Bleymaier, K.; Zitzenbacher, S.; Zankel, A.; Mayrhofer, C.; et al. An esterase from anaerobic Clostridium hathewayi can hydrolyze aliphatic-aromatic polyesters. Environ. Sci. Technol. 2016, 50, 2899–2907. [Google Scholar] [CrossRef]
  38. Huang, J.; Huo, Y.Y.; Ji, R.; Kuang, S.; Ji, C.; Xu, X.W.; Li, J. Structural insights of a hormone sensitive lipase homologue Est22. Sci. Rep. 2016, 6, 28550. [Google Scholar] [CrossRef]
  39. Leščić Ašler, I.; Štefanić, Z.; Maršavelski, A.; Vianello, R.; Kojić-Prodić, B. Catalytic dyad in the sgnh hydrolase superfamily: In-depth insight into structural parameters tuning the catalytic process of extracellular lipase from Streptomyces rimosus. ACS Chem. Biol. 2017, 12, 1928–1936. [Google Scholar] [CrossRef]
  40. Naffin-Olivos, J.L.; Daab, A.; White, A.; Goldfarb, N.E.; Milne, A.C.; Liu, D.; Baikovitz, J.; Dunn, B.M.; Rengarajan, J.; Petsko, G.A.; et al. Structure determination of Mycobacterium tuberculosis serine protease Hip1 (Rv2224c). Biochemistry 2017, 56, 2304–2314. [Google Scholar] [CrossRef]
  41. De Santi, C.; Gani, O.A.; Helland, R.; Williamson, A. Structural insight into a CE15 esterase from the marine bacterial metagenome. Sci. Rep. 2017, 7, 17278. [Google Scholar] [CrossRef] [Green Version]
  42. Arnling Baath, J.; Mazurkewich, S.; Knudsen, R.M.; Poulsen, J.N.; Olsson, L.; Lo Leggio, L.; Larsbrink, J. Biochemical and structural features of diverse bacterial glucuronoyl esterases facilitating recalcitrant biomass conversion. Biotechnol. Biofuels 2018, 11, 213. [Google Scholar] [CrossRef]
  43. Kim, K.K.; Song, H.K.; Shin, D.H.; Hwang, K.Y.; Choe, S.; Yoo, O.J.; Suh, S.W. Crystal structure of carboxylesterase from Pseudomonas fluorescens, an alpha/beta hydrolase with broad substrate specificity. Structure 1997, 5, 1571–1584. [Google Scholar] [CrossRef] [Green Version]
  44. De Simone, G.; Galdiero, S.; Manco, G.; Lang, D.; Rossi, M.; Pedone, C. A snapshot of a transition state analogue of a novel thermophilic esterase belonging to the subfamily of mammalian hormone-sensitive lipase. J. Mol. Biol. 2000, 303, 761–771. [Google Scholar] [CrossRef] [PubMed]
  45. Turner, J.M.; Larsen, N.A.; Basran, A.; Barbas, C.F., 3rd; Bruce, N.C.; Wilson, I.A.; Lerner, R.A. Biochemical characterization and structural analysis of a highly proficient cocaine esterase. Biochemistry 2002, 41, 12297–12307. [Google Scholar] [CrossRef] [PubMed]
  46. Liu, P.; Wang, Y.F.; Ewis, H.E.; Abdelal, A.T.; Lu, C.D.; Harrison, R.W.; Weber, I.T. Covalent reaction intermediate revealed in crystal structure of the Geobacillus stearothermophilus carboxylesterase Est30. J. Mol. Biol. 2004, 342, 551–561. [Google Scholar] [CrossRef]
  47. Mandrich, L.; Menchise, V.; Alterio, V.; De Simone, G.; Pedone, C.; Rossi, M.; Manco, G. Functional and structural features of the oxyanion hole in a thermophilic esterase from Alicyclobacillus acidocaldarius. Proteins 2008, 71, 1721–1731. [Google Scholar] [CrossRef] [PubMed]
  48. Pesaresi, A.; Lamba, D. Insights into the fatty acid chain length specificity of the carboxylesterase PA3859 from Pseudomonas aeruginosa: A combined structural, biochemical and computational study. Biochimie 2010, 92, 1787–1792. [Google Scholar] [CrossRef]
  49. Levisson, M.; Sun, L.; Hendriks, S.; Swinkels, P.; Akveld, T.; Bultema, J.B.; Barendregt, A.; van den Heuvel, R.H.; Dijkstra, B.W.; van der Oost, J.; et al. Crystal structure and biochemical properties of a novel thermostable esterase containing an immunoglobulin-like domain. J. Mol. Biol. 2009, 385, 949–962. [Google Scholar] [CrossRef] [Green Version]
  50. Van den Berg, B. Crystal structure of a full-length autotransporter. J. Mol. Biol. 2010, 396, 627–633. [Google Scholar] [CrossRef]
  51. Benavente, R.; Esteban-Torres, M.; Acebron, I.; de Las Rivas, B.; Munoz, R.; Alvarez, Y.; Mancheno, J.M. Structure, biochemical characterization and analysis of the pleomorphism of carboxylesterase Cest-2923 from Lactobacillus plantarum wcfs1. FEBS J. 2013, 280, 6658–6671. [Google Scholar] [CrossRef] [Green Version]
  52. Alvarez, Y.; Esteban-Torres, M.; Cortes-Cabrera, A.; Gago, F.; Acebron, I.; Benavente, R.; Mardo, K.; de Las Rivas, B.; Munoz, R.; Mancheno, J.M. Esterase LpEst1 from Lactobacillus plantarum: A novel and atypical member of the alphabeta hydrolase superfamily of enzymes. PLoS ONE 2014, 9, e92257. [Google Scholar] [CrossRef]
  53. Rozeboom, H.J.; Godinho, L.F.; Nardini, M.; Quax, W.J.; Dijkstra, B.W. Crystal structures of two Bacillus carboxylesterases with different enantioselectivities. Biochim. Biophys. Acta 2014, 1844, 567–575. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Ma, J.; Wu, L.; Guo, F.; Gu, J.; Tang, X.; Jiang, L.; Liu, J.; Zhou, J.; Yu, H. Enhanced enantioselectivity of a carboxyl esterase from Rhodobacter sphaeroides by directed evolution. Appl. Microbiol. Biotechnol. 2013, 97, 4897–4906. [Google Scholar] [CrossRef] [PubMed]
  55. Cha, S.S.; An, Y.J.; Jeong, C.S.; Kim, M.K.; Jeon, J.H.; Lee, C.M.; Lee, H.S.; Kang, S.G.; Lee, J.H. Structural basis for the beta-lactamase activity of EstU1, a family VIII carboxylesterase. Proteins 2013, 81, 2045–2051. [Google Scholar] [CrossRef] [PubMed]
  56. Kovacic, F.; Granzin, J.; Wilhelm, S.; Kojic-Prodic, B.; Batra-Safferling, R.; Jaeger, K.E. Structural and functional characterisation of TesA—A novel lysophospholipase A from Pseudomonas aeruginosa. PLoS ONE 2013, 8, e69125. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Sayer, C.; Isupov, M.N.; Bonch-Osmolovskaya, E.; Littlechild, J.A. Structural studies of a thermophilic esterase from a new planctomycetes species, Thermogutta terrifontis. FEBS J. 2015, 282, 2846–2857. [Google Scholar] [CrossRef]
  58. De Santi, C.; Leiros, H.K.; Di Scala, A.; de Pascale, D.; Altermark, B.; Willassen, N.P. Biochemical characterization and structural analysis of a new cold-active and salt-tolerant esterase from the marine Bacterium thalassospira sp. Extremophiles 2016, 20, 323–336. [Google Scholar] [CrossRef]
  59. Pereira, M.R.; Maester, T.C.; Mercaldi, G.F.; de Macedo Lemos, E.G.; Hyvonen, M.; Balan, A. From a metagenomic source to a high-resolution structure of a novel alkaline esterase. Appl. Microbiol. Biotechnol. 2017, 101, 4935–4949. [Google Scholar] [CrossRef]
  60. Sayer, C.; Szabo, Z.; Isupov, M.N.; Ingham, C.; Littlechild, J.A. The structure of a novel thermophilic esterase from the Planctomycetes species, Thermogutta terrifontis reveals an open active site due to a minimal ‘cap’ domain. Front Microbiol. 2015, 6, 1294. [Google Scholar] [CrossRef]
  61. Cha, S.S.; An, Y.J. Crystal structure of EstSRT1, a family VIII carboxylesterase displaying hydrolytic activity toward oxyimino cephalosporins. Biochem. Biophys. Res. Commun. 2016, 478, 818–824. [Google Scholar] [CrossRef]
  62. Shi, J.; Cao, X.; Chen, Y.; Cronan, J.E.; Guo, Z. An atypical alpha/beta-hydrolase fold revealed in the crystal structure of pimeloyl-acyl carrier protein methyl esterase biog from Haemophilus influenzae. Biochemistry 2016, 55, 6705–6717. [Google Scholar] [CrossRef]
  63. Kim, S.H.; Kang, P.A.; Han, K.; Lee, S.W.; Rhee, S. Crystal structure of chloramphenicol-metabolizing enzyme EstDL136 from a metagenome. PLoS ONE 2019, 14, e0210298. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Cheeseman, J.D.; Tocilj, A.; Park, S.; Schrag, J.D.; Kazlauskas, R.J. Structure of an aryl esterase from Pseudomonas fluorescens. Acta Crystallogr. D Biol. Crystallogr. 2004, 60, 1237–1243. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Mathews, I.; Soltis, M.; Saldajeno, M.; Ganshaw, G.; Sala, R.; Weyler, W.; Cervin, M.A.; Whited, G.; Bott, R. Structure of a novel enzyme that catalyzes acyl transfer to alcohols in aqueous conditions. Biochemistry 2007, 46, 8969–8979. [Google Scholar] [CrossRef] [PubMed]
  66. Yin, D.L.; Bernhardt, P.; Morley, K.L.; Jiang, Y.; Cheeseman, J.D.; Purpero, V.; Schrag, J.D.; Kazlauskas, R.J. Switching catalysis from hydrolysis to perhydrolysis in Pseudomonas fluorescens esterase. Biochemistry 2010, 49, 1931–1942. [Google Scholar] [CrossRef] [Green Version]
  67. Jiang, Y.; Morley, K.L.; Schrag, J.D.; Kazlauskas, R.J. Different active-site loop orientation in serine hydrolases versus acyltransferases. Chembiochem 2011, 12, 768–776. [Google Scholar] [CrossRef]
  68. Kim, K.; Ryu, B.H.; Kim, S.S.; An, D.R.; Ngo, T.D.; Pandian, R.; Kim, K.K.; Kim, T.D. Structural and biochemical characterization of a carbohydrate acetylesterase from Sinorhizobium meliloti 1021. FEBS Lett. 2015, 589, 117–122. [Google Scholar] [CrossRef] [Green Version]
  69. Lang, D.; Hofmann, B.; Haalck, L.; Hecht, H.J.; Spener, F.; Schmid, R.D.; Schomburg, D. Crystal structure of a bacterial lipase from Chromobacterium viscosum ATCC 6918 refined at 1.6 Å resolution. J. Mol. Biol. 1996, 259, 704–717. [Google Scholar] [CrossRef]
  70. Nardini, M.; Lang, D.A.; Liebeton, K.; Jaeger, K.E.; Dijkstra, B.W. Crystal structure of Pseudomonas aeruginosa lipase in the open conformation. The prototype for family I.1 of bacterial lipases. J. Biol. Chem. 2000, 275, 31219–31225. [Google Scholar] [CrossRef] [Green Version]
  71. Luic, M.; Tomic, S.; Lescic, I.; Ljubovic, E.; Sepac, D.; Sunjic, V.; Vitale, L.; Saenger, W.; Kojic-Prodic, B. Complex of Burkholderia cepacia lipase with transition state analogue of 1-phenoxy-2-acetoxybutane: Biocatalytic, structural and modelling study. Eur. J. Biochem. 2001, 268, 3964–3973. [Google Scholar] [CrossRef]
  72. Van Pouderoyen, G.; Eggert, T.; Jaeger, K.E.; Dijkstra, B.W. The crystal structure of Bacillus subtilis lipase: A minimal alpha/beta hydrolase fold enzyme. J. Mol. Biol. 2001, 309, 215–226. [Google Scholar] [CrossRef] [Green Version]
  73. Kawasaki, K.; Kondo, H.; Suzuki, M.; Ohgiya, S.; Tsuda, S. Alternate conformations observed in catalytic serine of Bacillus subtilis lipase determined at 1.3 Å resolution. Acta. Crystallogr. D Biol. Crystallogr. 2002, 58, 1168–1174. [Google Scholar] [CrossRef] [PubMed]
  74. Tyndall, J.D.; Sinchaikul, S.; Fothergill-Gilmore, L.A.; Taylor, P.; Walkinshaw, M.D. Crystal structure of a thermostable lipase from Bacillus stearothermophilus P1. J. Mol. Biol. 2002, 323, 859–869. [Google Scholar] [CrossRef]
  75. Jeong, S.T.; Kim, H.K.; Kim, S.J.; Chi, S.W.; Pan, J.G.; Oh, T.K.; Ryu, S.E. Novel zinc-binding center and a temperature switch in the Bacillus stearothermophilus L1 lipase. J. Biol. Chem. 2002, 277, 17041–17047. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Kim, K.K.; Song, H.K.; Shin, D.H.; Hwang, K.Y.; Suh, S.W. The crystal structure of a triacylglycerol lipase from Pseudomonas cepacia reveals a highly open conformation in the absence of a bound inhibitor. Structure 1997, 5, 173–185. [Google Scholar] [CrossRef] [Green Version]
  77. Dröge, M.J.; Boersma, Y.L.; van Pouderoyen, G.; Vrenken, T.E.; Rüggeberg, C.J.; Reetz, M.T.; Dijkstra, B.W.; Quax, W.J. Directed evolution of Bacillus subtilis lipase a by use of enantiomeric phosphonate inhibitors: Crystal structures and phage display selection. ChemBioChem 2006, 7, 149–157. [Google Scholar] [CrossRef]
  78. Acharya, P.; Rajakumara, E.; Sankaranarayanan, R.; Rao, N.M. Structural basis of selection and thermostability of laboratory evolved Bacillus subtilis lipase. J. Mol. Biol. 2004, 341, 1271–1281. [Google Scholar] [CrossRef]
  79. Mezzetti, A.; Schrag, J.D.; Cheong, C.S.; Kazlauskas, R.J. Mirror-image packing in enantiomer discrimination: molecular basis for the enantioselectivity of B. cepacia lipase toward 2-methyl-3-phenyl-1-propanol. Chem. Biol. 2005, 12, 427–437. [Google Scholar] [CrossRef] [Green Version]
  80. Matsumura, H.; Yamamoto, T.; Leow, T.C.; Mori, T.; Salleh, A.B.; Basri, M.; Inoue, T.; Kai, Y.; Rahman, R.N. Novel cation-pi interaction revealed by crystal structure of thermoalkalophilic lipase. Proteins 2008, 70, 592–598. [Google Scholar] [CrossRef]
  81. Pauwels, K.; Lustig, A.; Wyns, L.; Tommassen, J.; Savvides, S.N.; Van Gelder, P. Structure of a membrane-based steric chaperone in complex with its lipase substrate. Nat. Struct. Mol. Biol. 2006, 13, 374–375. [Google Scholar] [CrossRef]
  82. Tiesinga, J.J.; van Pouderoyen, G.; Nardini, M.; Ransac, S.; Dijkstra, B.W. Structural basis of phospholipase activity of Staphylococcus hyicus lipase. J. Mol. Biol. 2007, 371, 447–456. [Google Scholar] [CrossRef] [Green Version]
  83. Schrag, J.D.; Li, Y.; Cygler, M.; Lang, D.; Burgdorf, T.; Hecht, H.J.; Schmid, R.; Schomburg, D.; Rydel, T.J.; Oliver, J.D.; et al. The open conformation of a Pseudomonas lipase. Structure 1997, 5, 187–202. [Google Scholar] [CrossRef] [Green Version]
  84. Luic, M.; Stefanic, Z.; Ceilinger, I.; Hodoscek, M.; Janezic, D.; Lenac, T.; Asler, I.L.; Sepac, D.; Tomic, S. Combined X-ray diffraction and QM/MM study of the Burkholderia cepacia lipase-catalyzed secondary alcohol esterification. J. Phys. Chem. B 2008, 112, 4876–4883. [Google Scholar] [CrossRef] [PubMed]
  85. Jung, S.K.; Jeong, D.G.; Lee, M.S.; Lee, J.K.; Kim, H.K.; Ryu, S.E.; Park, B.C.; Kim, J.H.; Kim, S.J. Structural basis for the cold adaptation of psychrophilic M37 lipase from Photobacterium lipolyticum. Proteins 2008, 71, 476–484. [Google Scholar] [CrossRef] [PubMed]
  86. Meier, R.; Drepper, T.; Svensson, V.; Jaeger, K.E.; Baumann, U. A calcium-gated lid and a large beta-roll sandwich are revealed by the crystal structure of extracellular lipase from Serratia marcescens. J. Biol. Chem. 2007, 282, 31477–31483. [Google Scholar] [CrossRef] [Green Version]
  87. Rajakumara, E.; Acharya, P.; Ahmad, S.; Sankaranaryanan, R.; Rao, N.M. Structural basis for the remarkable stability of Bacillus subtilis lipase (Lip A) at low pH. Biochim. Biophys. Acta 2008, 1784, 302–311. [Google Scholar] [CrossRef]
  88. Carrasco-Lopez, C.; Godoy, C.; de Las Rivas, B.; Fernandez-Lorente, G.; Palomo, J.M.; Guisan, J.M.; Fernandez-Lafuente, R.; Martinez-Ripoll, M.; Hermoso, J.A. Activation of bacterial thermoalkalophilic lipases is spurred by dramatic structural rearrangements. J. Biol. Chem. 2009, 284, 4365–4372. [Google Scholar] [CrossRef] [Green Version]
  89. Angkawidjaja, C.; You, D.J.; Matsumura, H.; Kuwahara, K.; Koga, Y.; Takano, K.; Kanaya, S. Crystal structure of a family I.3 lipase from Pseudomonas sp. MIS38 in a closed conformation. FEBS Lett. 2007, 581, 5060–5064. [Google Scholar] [CrossRef] [Green Version]
  90. Kuwahara, K.; Angkawidjaja, C.; Matsumura, H.; Koga, Y.; Takano, K.; Kanaya, S. Importance of the Ca2+-binding sites in the N-catalytic domain of a family I.3 lipase for activity and stability. Protein Eng. Des. Sel. 2008, 21, 737–744. [Google Scholar] [CrossRef]
  91. Angkawidjaja, C.; Matsumura, H.; Koga, Y.; Takano, K.; Kanaya, S. X-ray crystallographic and MD simulation studies on the mechanism of interfacial activation of a family I.3 lipase with two lids. J. Mol. Biol. 2010, 400, 82–95. [Google Scholar] [CrossRef]
  92. Ahmad, S.; Kamal, M.Z.; Sankaranarayanan, R.; Rao, N.M. Thermostable Bacillus subtilis lipases: In vitro evolution and structural insight. J. Mol. Biol. 2008, 381, 324–340. [Google Scholar] [CrossRef]
  93. Kamal, M.Z.; Ahmad, S.; Molugu, T.R.; Vijayalakshmi, A.; Deshmukh, M.V.; Sankaranarayanan, R.; Rao, N.M. In vitro evolved non-aggregating and thermostable lipase: Structural and thermodynamic investigation. J. Mol. Biol. 2011, 413, 726–741. [Google Scholar] [CrossRef] [PubMed]
  94. Augustyniak, W.; Brzezinska, A.A.; Pijning, T.; Wienk, H.; Boelens, R.; Dijkstra, B.W.; Reetz, M.T. Biophysical characterization of mutants of Bacillus subtilis lipase evolved for thermostability: Factors contributing to increased activity retention. Protein Sci. 2012, 21, 487–497. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Ruslan, R.; Abd Rahman, R.N.; Leow, T.C.; Ali, M.S.; Basri, M.; Salleh, A.B. Improvement of thermal stability via outer-loop ion pair interaction of mutated T1 lipase from Geobacillus zalihae strain T1. Int. J. Mol. Sci. 2012, 13, 943–960. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Abd Rahman, R.N.; Shariff, F.M.; Basri, M.; Salleh, A.B. 3D structure elucidation of thermostable L2 lipase from thermophilic Bacillus sp. L2. Int. J. Mol. Sci. 2012, 13, 9207–9217. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Korman, T.P.; Bowie, J.U. Crystal structure of Proteus mirabilis lipase, a novel lipase from the Proteus/psychrophilic subfamily of lipase family I.1. PLoS ONE 2012, 7, e52890. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  98. Lang, D.A.; Mannesse, M.L.; de Haas, G.H.; Verheij, H.M.; Dijkstra, B.W. Structural basis of the chiral selectivity of Pseudomonas cepacia lipase. Eur. J. Biochem. 1998, 254, 333–340. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  99. Dror, A.; Kanteev, M.; Kagan, I.; Gihaz, S.; Shahar, A.; Fishman, A. Structural insights into methanol-stable variants of lipase T6 from Geobacillus stearothermophilus. Appl. Microbiol. Biotechnol. 2015, 99, 9449–9461. [Google Scholar] [CrossRef]
  100. Perz, V.; Baumschlager, A.; Bleymaier, K.; Zitzenbacher, S.; Hromic, A.; Steinkellner, G.; Pairitsch, A.; Lyskowski, A.; Gruber, K.; Sinkel, C.; et al. Hydrolysis of synthetic polyesters by Clostridium botulinum esterases. Biotechnol. Bioeng. 2016, 113, 1024–1034. [Google Scholar] [CrossRef]
  101. Nordwald, E.M.; Plaks, J.G.; Snell, J.R.; Sousa, M.C.; Kaar, J.L. Crystallographic investigation of imidazolium ionic liquid effects on enzyme structure. Chembiochem 2015, 16, 2456–2459. [Google Scholar] [CrossRef] [Green Version]
  102. Zhao, Z.; Hou, S.; Lan, D.; Wang, X.; Liu, J.; Khan, F.I.; Wang, Y. Crystal structure of a lipase from Streptomyces sp. Strain W007—Implications for thermostability and regiospecificity. FEBS J. 2017, 284, 3506–3519. [Google Scholar] [CrossRef] [Green Version]
  103. Moharana, T.R.; Pal, B.; Rao, N.M. X-ray structure and characterization of a thermostable lipase from Geobacillus thermoleovorans. Biochem. Biophys. Res. Commun. 2019, 508, 145–151. [Google Scholar] [CrossRef] [PubMed]
  104. Gihaz, S.; Kanteev, M.; Pazy, Y.; Fishman, A. Filling the void: Introducing aromatic interactions into solvent tunnels to enhance lipase stability in methanol. Appl. Microbiol. Biotechnol. 2018, 84, e02143-18. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Matoba, Y.; Katsube, Y.; Sugiyama, M. The crystal structure of prokaryotic phospholipase A2. J. Biol. Chem. 2002, 277, 20059–20069. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Matoba, Y.; Sugiyama, M. Atomic resolution structure of prokaryotic phospholipase A2: Analysis of internal motion and implication for a catalytic mechanism. Proteins 2003, 51, 453–469. [Google Scholar] [CrossRef]
  107. Snijder, H.J.; Ubarretxena-Belandia, I.; Blaauw, M.; Kalk, K.H.; Verheij, H.M.; Egmond, M.R.; Dekker, N.; Dijkstra, B.W. Structural evidence for dimerization-regulated activation of an integral membrane phospholipase. Nature 1999, 401, 717–721. [Google Scholar] [CrossRef] [PubMed]
  108. Snijder, H.J.; Kingma, R.L.; Kalk, K.H.; Dekker, N.; Egmond, M.R.; Dijkstra, B.W. Structural investigations of calcium binding and its role in activity and activation of outer membrane phospholipase A from Escherichia coli. J. Mol. Biol. 2001, 309, 477–489. [Google Scholar] [CrossRef] [Green Version]
  109. Snijder, H.J.; Van Eerde, J.H.; Kingma, R.L.; Kalk, K.H.; Dekker, N.; Egmond, M.R.; Dijkstra, B.W. Structural investigations of the active-site mutant Asn156Ala of outer membrane phospholipase A: Function of the Asn-his interaction in the catalytic triad. Protein Sci. 2001, 10, 1962–1969. [Google Scholar] [CrossRef]
  110. Lo, Y.C.; Lin, S.C.; Shaw, J.F.; Liaw, Y.C. Crystal structure of Escherichia coli thioesterase I/protease I/lysophospholipase L1: Consensus sequence blocks constitute the catalytic center of SGNH-hydrolases through a conserved hydrogen bond network. J. Mol. Biol. 2003, 330, 539–551. [Google Scholar] [CrossRef]
  111. Lo, Y.C.; Lin, S.C.; Shaw, J.F.; Liaw, Y.C. Substrate specificities of Escherichia coli thioesterase I/protease I/lysophospholipase L1 are governed by its switch loop movement. Biochemistry 2005, 44, 1971–1979. [Google Scholar] [CrossRef]
  112. Grisewood, M.J.; Hernandez Lozada, N.J.; Thoden, J.B.; Gifford, N.P.; Mendez-Perez, D.; Schoenberger, H.A.; Allan, M.F.; Floy, M.E.; Lai, R.Y.; Holden, H.M.; et al. Computational redesign of acyl-ACP thioesterase with improved selectivity toward medium-chain-length fatty acids. ACS Catal. 2017, 7, 3837–3849. [Google Scholar] [CrossRef] [Green Version]
  113. Montoro-García, S.; Gil-Ortiz, F.; García-Carmona, F.; Polo, L.M.; Rubio, V.; Sánchez-Ferrer, Á. The crystal structure of the cephalosporin deacetylating enzyme acetyl xylan esterase bound to paraoxon explains the low sensitivity of this serine hydrolase to organophosphate inactivation. Biochem. J. 2011, 436, 321–330. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Jenkins, J.; Mayans, O.; Smith, D.; Worboys, K.; Pickersgill, R.W. Three-dimensional structure of Erwinia chrysanthemi pectin methylesterase reveals a novel esterase active site. J. Mol. Biol. 2001, 305, 951–960. [Google Scholar] [CrossRef] [PubMed]
  115. Fries, M.; Ihrig, J.; Brocklehurst, K.; Shevchik, V.E.; Pickersgill, R.W. Molecular basis of the activity of the phytopathogen pectin methylesterase. EMBO J. 2007, 26, 3879–3887. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Boraston, A.B.; Abbott, D.W. Structure of a pectin methylesterase from Yersinia enterocolitica. Acta Crystallogr. F Struct Biol. Commun. 2012, 68, 129–133. [Google Scholar] [CrossRef]
  117. Chen, C.N.; Chin, K.H.; Wang, A.H.; Chou, S.H. The first crystal structure of gluconolactonase important in the glucose secondary metabolic pathways. J. Mol. Biol. 2008, 384, 604–614. [Google Scholar] [CrossRef]
  118. Matoba, Y.; Tanaka, N.; Noda, M.; Higashikawa, F.; Kumagai, T.; Sugiyama, M. Crystallographic and mutational analyses of tannase from Lactobacillus plantarum. Proteins 2013, 81, 2052–2058. [Google Scholar] [CrossRef]
  119. Rengachari, S.; Bezerra, G.A.; Riegler-Berket, L.; Gruber, C.C.; Sturm, C.; Taschler, U.; Boeszoermenyi, A.; Dreveny, I.; Zimmermann, R.; Gruber, K.; et al. The structure of monoacylglycerol lipase from Bacillus sp. H257 reveals unexpected conservation of the cap architecture between bacterial and human enzymes. Biochim. Biophys. Acta 2012, 1821, 1012–1021. [Google Scholar] [CrossRef]
  120. Rengachari, S.; Aschauer, P.; Schittmayer, M.; Mayer, N.; Gruber, K.; Breinbauer, R.; Birner-Gruenberger, R.; Dreveny, I.; Oberer, M. Conformational plasticity and ligand binding of bacterial monoacylglycerol lipase. J. Biol. Chem. 2013, 288, 31093–31104. [Google Scholar] [CrossRef] [Green Version]
  121. Tsurumura, T.; Tsuge, H. Substrate selectivity of bacterial monoacylglycerol lipase based on crystal structure. J. Struct. Funct. Genom. 2014, 15, 83–89. [Google Scholar] [CrossRef]
  122. Bains, J.; Kaufman, L.; Farnell, B.; Boulanger, M.J. A product analog bound form of 3-oxoadipate-enol-lactonase (PcaD) reveals a multifunctional role for the divergent cap domain. J. Mol. Biol. 2011, 406, 649–658. [Google Scholar] [CrossRef]
  123. Schmitt, E.; Mechulam, Y.; Fromant, M.; Plateau, P.; Blanquet, S. Crystal structure at 1.2 Å resolution and active site mapping of Escherichia coli peptidyl-tRNA hydrolase. EMBO J. 1997, 16, 4760–4769. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Selvaraj, M.; Roy, S.; Singh, N.S.; Sangeetha, R.; Varshney, U.; Vijayan, M. Structural plasticity and enzyme action: Crystal structures of Mycobacterium tuberculosis peptidyl-tRNA hydrolase. J. Mol. Biol. 2007, 372, 186–193. [Google Scholar] [CrossRef] [PubMed]
  125. Clarke, T.E.; Romanov, V.; Lam, R.; Gothe, S.A.; Peddi, S.R.; Razumova, E.B.; Lipman, R.S.; Branstrom, A.A.; Chirgadze, N.Y. Structure of Francisella tularensis peptidyl-tRNA hydrolase. Acta Crystallogr. F Struct. Biol. Commun. 2011, 67, 446–449. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. Selvaraj, M.; Ahmad, R.; Varshney, U.; Vijayan, M. Structures of new crystal forms of Mycobacterium tuberculosis peptidyl-tRNA hydrolase and functionally important plasticity of the molecule. Acta Crystallogr. F Struct. Biol. Commun. 2012, 68, 124–128. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  127. Baugh, L.; Gallagher, L.A.; Patrapuvich, R.; Clifton, M.C.; Gardberg, A.S.; Edwards, T.E.; Armour, B.; Begley, D.W.; Dieterich, S.H.; Dranow, D.M.; et al. Combining functional and structural genomics to sample the essential Burkholderia structome. PLoS ONE 2013, 8, e53851. [Google Scholar] [CrossRef]
  128. Ito, K.; Murakami, R.; Mochizuki, M.; Qi, H.; Shimizu, Y.; Miura, K.; Ueda, T.; Uchiumi, T. Structural basis for the substrate recognition and catalysis of peptidyl-tRNA hydrolase. Nucleic Acids Res. 2012, 40, 10521–10531. [Google Scholar] [CrossRef] [PubMed]
  129. Singh, A.; Kumar, A.; Gautam, L.; Sharma, P.; Sinha, M.; Bhushan, A.; Kaur, P.; Sharma, S.; Arora, A.; Singh, T.P. Structural and binding studies of peptidyl-tRNA hydrolase from Pseudomonas aeruginosa provide a platform for the structure-based inhibitor design against peptidyl-tRNA hydrolase. Biochem. J. 2014, 463, 329–337. [Google Scholar] [CrossRef] [PubMed]
  130. Kaushik, S.; Singh, N.; Yamini, S.; Singh, A.; Sinha, M.; Arora, A.; Kaur, P.; Sharma, S.; Singh, T.P. The mode of inhibitor binding to peptidyl-tRNA hydrolase: Binding studies and structure determination of unbound and bound peptidyl-tRNA hydrolase from Acinetobacter baumannii. PLoS ONE 2013, 8, e67547. [Google Scholar] [CrossRef]
  131. Hughes, R.C.; McFeeters, H.; Coates, L.; McFeeters, R.L. Recombinant production, crystallization and X-ray crystallographic structure determination of the peptidyl-tRNA hydrolase of Pseudomonas aeruginosa. Acta Crystallogr. F Struct. Biol. Commun. 2012, 68, 1472–1476. [Google Scholar] [CrossRef] [Green Version]
  132. Vandavasi, V.; Taylor-Creel, K.; McFeeters, R.L.; Coates, L.; McFeeters, H. Recombinant production, crystallization and X-ray crystallographic structure determination of peptidyl-tRNA hydrolase from Salmonella typhimurium. Acta Crystallogr. F Struct. Biol. Commun. 2014, 70, 872–877. [Google Scholar] [CrossRef] [Green Version]
  133. Singh, A.; Gautam, L.; Sinha, M.; Bhushan, A.; Kaur, P.; Sharma, S.; Singh, T.P. Crystal structure of peptidyl-tRNA hydrolase from a gram-positive bacterium, Streptococcus pyogenes at 2.19 Å resolution shows the closed structure of the substrate-binding cleft. FEBS Open Bio. 2014, 4, 915–922. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Gagnon, M.G.; Seetharaman, S.V.; Bulkley, D.; Steitz, T.A. Structural basis for the rescue of stalled ribosomes: Structure of YaeJ bound to the ribosome. Science 2012, 335, 1370–1372. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Zhang, F.; Song, Y.; Niu, L.; Teng, M.; Li, X. Crystal structure of Staphylococcus aureus peptidyl-tRNA hydrolase at a 2.25 Å resolution. Acta Biochim. Biophys. Sin. (Shanghai) 2015, 47, 1005–1010. [Google Scholar] [PubMed] [Green Version]
  136. Kabra, A.; Shahid, S.; Pal, R.K.; Yadav, R.; Pulavarti, S.V.; Jain, A.; Tripathi, S.; Arora, A. Unraveling the stereochemical and dynamic aspects of the catalytic site of bacterial peptidyl-tRNA hydrolase. RNA 2017, 23, 202–216. [Google Scholar] [CrossRef] [Green Version]
  137. Kaushik, S.; Iqbal, N.; Singh, N.; Sikarwar, J.S.; Singh, P.K.; Sharma, P.; Kaur, P.; Sharma, S.; Owais, M.; Singh, T.P. Search of multiple hot spots on the surface of peptidyl-tRNA hydrolase: Structural, binding and antibacterial studies. Biochem. J. 2018, 475, 547–560. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  138. Matsumoto, A.; Uehara, Y.; Shimizu, Y.; Ueda, T.; Uchiumi, T.; Ito, K. High-resolution crystal structure of peptidyl-tRNA hydrolase from Thermus thermophilus. Proteins 2019, 87, 226–235. [Google Scholar] [CrossRef]
  139. Baugh, L.; Phan, I.; Begley, D.W.; Clifton, M.C.; Armour, B.; Dranow, D.M.; Taylor, B.M.; Muruthi, M.M.; Abendroth, J.; Fairman, J.W.; et al. Increasing the structural coverage of tuberculosis drug targets. Tuberculosis (Edinb) 2015, 95, 142–148. [Google Scholar] [CrossRef] [Green Version]
  140. Fujieda, N.; Schätti, J.; Stuttfeld, E.; Ohkubo, K.; Maier, T.; Fukuzumi, S.; Ward, T.R. Enzyme repurposing of a hydrolase as an emergent peroxidase upon metal binding. Chem. Sci. 2015, 6, 4060–4065. [Google Scholar] [CrossRef] [Green Version]
  141. Murayama, K.; Kano, K.; Matsumoto, Y.; Sugimori, D. Crystal structure of phospholipase A1 from Streptomyces albidoflavus NA297. J. Struct. Biol. 2013, 182, 192–196. [Google Scholar] [CrossRef]
  142. Vincent, F.; Charnock, S.J.; Verschueren, K.H.; Turkenburg, J.P.; Scott, D.J.; Offen, W.A.; Roberts, S.; Pell, G.; Gilbert, H.J.; Davies, G.J.; et al. Multifunctional xylooligosaccharide/cephalosporin c deacetylase revealed by the hexameric structure of the Bacillus subtilis enzyme at 1.9 Å resolution. J. Mol. Biol. 2003, 330, 593–606. [Google Scholar] [CrossRef]
  143. Levisson, M.; Han, G.W.; Deller, M.C.; Xu, Q.; Biely, P.; Hendriks, S.; Ten Eyck, L.F.; Flensburg, C.; Roversi, P.; Miller, M.D.; et al. Functional and structural characterization of a thermostable acetyl esterase from Thermotoga maritima. Proteins 2012, 80, 1545–1559. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Singh, M.K.; Manoj, N. An extended loop in CE7 carbohydrate esterase family is dispensable for oligomerization but required for activity and thermostability. J. Struct. Biol. 2016, 194, 434–445. [Google Scholar] [CrossRef] [PubMed]
  145. Singh, M.K.; Manoj, N. Structural role of a conserved active site cis proline in the Thermotoga maritima acetyl esterase from the carbohydrate esterase family 7. Proteins 2017, 85, 694–708. [Google Scholar] [CrossRef] [PubMed]
  146. Singh, M.K.; Manoj, N. Crystal structure of Thermotoga maritima acetyl esterase complex with a substrate analog: Insights into the distinctive substrate specificity in the CE7 carbohydrate esterase family. Biochem. Biophys. Res. Commun. 2016, 476, 63–68. [Google Scholar] [CrossRef] [PubMed]
  147. Barends, T.R.; Polderman-Tijmes, J.J.; Jekel, P.A.; Hensgens, C.M.; de Vries, E.J.; Janssen, D.B.; Dijkstra, B.W. The sequence and crystal structure of the alpha-amino acid ester hydrolase from Xanthomonas citri define a new family of beta-lactam antibiotic acylases. J. Biol. Chem. 2003, 278, 23076–23084. [Google Scholar] [CrossRef] [Green Version]
  148. Barends, T.R.; Polderman-Tijmes, J.J.; Jekel, P.A.; Williams, C.; Wybenga, G.; Janssen, D.B.; Dijkstra, B.W. Acetobacter turbidans alpha-amino acid ester hydrolase: How a single mutation improves an antibiotic-producing enzyme. J. Biol. Chem. 2006, 281, 5804–5810. [Google Scholar] [CrossRef] [Green Version]
  149. Pathak, D.; Ollis, D. Refined structure of dienelactone hydrolase at 1.8 Å. J. Mol. Biol. 1990, 214, 497–525. [Google Scholar] [CrossRef]
  150. Robinson, A.; Edwards, K.J.; Carr, P.D.; Barton, J.D.; Ewart, G.D.; Ollis, D.L. Structure of the C123S mutant of dienelactone hydrolase (DLH) bound with the PMS moiety of the protease inhibitor phenylmethylsulfonyl fluoride (PMSF). Acta Crystallogr. D Biol. Crystallogr. 2000, 56, 1376–1384. [Google Scholar] [CrossRef]
  151. Kim, H.K.; Liu, J.W.; Carr, P.D.; Ollis, D.L. Following directed evolution with crystallography: Structural changes observed in changing the substrate specificity of dienelactone hydrolase. Acta Crystallogr. D Biol. Crystallogr. 2005, 61, 920–931. [Google Scholar] [CrossRef] [Green Version]
  152. Porter, J.L.; Carr, P.D.; Collyer, C.A.; Ollis, D.L. Crystallization of dienelactone hydrolase in two space groups: Structural changes caused by crystal packing. Acta Crystallogr. F Struct. Biol. Commun. 2014, 70, 884–889. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  153. Porter, J.L.; Boon, P.L.; Murray, T.P.; Huber, T.B.; Collyer, C.A.; Ollis, D.L. Directed evolution of new and improved enzyme functions using an evolutionary intermediate and multidirectional search. ACS Chem. Biol. 2015, 10, 611–621. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  154. Hobbs, M.E.; Malashkevich, V.; Williams, H.J.; Xu, C.; Sauder, J.M.; Burley, S.K.; Almo, S.C.; Raushel, F.M. Structure and catalytic mechanism of LigI: Insight into the amidohydrolase enzymes of cog3618 and lignin degradation. Biochemistry 2012, 51, 3497–3507. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  155. Cho, K.H.; Crane, B.R.; Park, S.Y. An insight into the interaction mode between CheB and chemoreceptor from two crystal structures of CheB methylesterase catalytic domain. Biochem. Biophys. Res. Commun. 2011, 411, 69–75. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  156. Park, S.Y.; Crane, B.R. Structural insight into the low affinity between Thermotoga maritima CheA and CheB compared to their Escherichia coli/Salmonella typhimurium counterparts. Int. J. Biol. Macromol. 2011, 49, 794–800. [Google Scholar] [CrossRef] [Green Version]
  157. Taylor, E.J.; Gloster, T.M.; Turkenburg, J.P.; Vincent, F.; Brzozowski, A.M.; Dupont, C.; Shareck, F.; Centeno, M.S.; Prates, J.A.; Puchart, V.; et al. Structure and activity of two metal ion-dependent acetylxylan esterases involved in plant cell wall degradation reveals a close similarity to peptidoglycan deacetylases. J. Biol. Chem. 2006, 281, 10968–10975. [Google Scholar] [CrossRef] [Green Version]
  158. Correia, M.A.; Prates, J.A.; Bras, J.; Fontes, C.M.; Newman, J.A.; Lewis, R.J.; Gilbert, H.J.; Flint, J.E. Crystal structure of a cellulosomal family 3 carbohydrate esterase from Clostridium thermocellum provides insights into the mechanism of substrate recognition. J. Mol. Biol. 2008, 379, 64–72. [Google Scholar] [CrossRef]
  159. Lansky, S.; Alalouf, O.; Solomon, H.V.; Alhassid, A.; Govada, L.; Chayen, N.E.; Belrhali, H.; Shoham, Y.; Shoham, G. A unique octameric structure of Axe2, an intracellular acetyl-xylooligosaccharide esterase from Geobacillus stearothermophilus. Acta Crystallogr. D Biol. Crystallogr. 2014, 70, 261–278. [Google Scholar] [CrossRef]
  160. Uraji, M.; Tamura, H.; Mizohata, E.; Arima, J.; Wan, K.; Ogawa, K.; Inoue, T.; Hatanaka, T. Loop of streptomyces feruloyl esterase plays an important role in the enzyme’s catalyzing the release of ferulic acid from biomass. Appl. Environ. Microbiol. 2018, 84, e02300–e02317. [Google Scholar] [CrossRef] [Green Version]
  161. Roth, C.; Wei, R.; Oeser, T.; Then, J.; Follner, C.; Zimmermann, W.; Strater, N. Structural and functional studies on a thermostable polyethylene terephthalate degrading hydrolase from Thermobifida fusca. Appl. Microbiol. Biotechnol. 2014, 98, 7815–7823. [Google Scholar] [CrossRef]
  162. Numoto, N.; Kamiya, N.; Bekker, G.J.; Yamagami, Y.; Inaba, S.; Ishii, K.; Uchiyama, S.; Kawai, F.; Ito, N.; Oda, M. Structural dynamics of the PET-degrading cutinase-like enzyme from Saccharomonospora viridis AHK190 in substrate-bound states elucidates the Ca(2+)-driven catalytic cycle. Biochemistry 2018, 57, 5289–5300. [Google Scholar] [CrossRef]
  163. Jendrossek, D.; Hermawan, S.; Subedi, B.; Papageorgiou, A.C. Biochemical analysis and structure determination of Paucimonas lemoignei poly(3-hydroxybutyrate) (PHB) depolymerase PhaZ7 muteins reveal the PHB binding site and details of substrate-enzyme interactions. Mol. Microbiol. 2013, 90, 649–664. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  164. Kim, M.H.; Choi, W.C.; Kang, H.O.; Lee, J.S.; Kang, B.S.; Kim, K.J.; Derewenda, Z.S.; Oh, T.K.; Lee, C.H.; Lee, J.K. The molecular structure and catalytic mechanism of a quorum-quenching N-acyl-L-homoserine lactone hydrolase. Proc. Natl. Acad. Sci. USA 2005, 102, 17606–17611. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  165. Liu, D.; Momb, J.; Thomas, P.W.; Moulin, A.; Petsko, G.A.; Fast, W.; Ringe, D. Mechanism of the quorum-quenching lactonase (AiiA) from Bacillus thuringiensis. 1. Product-bound structures. Biochemistry 2008, 47, 7706–7714. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  166. Liu, C.F.; Liu, D.; Momb, J.; Thomas, P.W.; Lajoie, A.; Petsko, G.A.; Fast, W.; Ringe, D. A phenylalanine clamp controls substrate specificity in the quorum-quenching metallo-gamma-lactonase from Bacillus thuringiensis. Biochemistry 2013, 52, 1603–1610. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  167. Yang, G.; Hong, N.; Baier, F.; Jackson, C.J.; Tokuriki, N. Conformational tinkering drives evolution of a promiscuous activity through indirect mutational effects. Biochemistry 2016, 55, 4583–4593. [Google Scholar] [CrossRef] [PubMed]
  168. Larsen, N.A.; Turner, J.M.; Stevens, J.; Rosser, S.J.; Basran, A.; Lerner, R.A.; Bruce, N.C.; Wilson, I.A. Crystal structure of a bacterial cocaine esterase. Nat. Struct. Biol. 2002, 9, 17–21. [Google Scholar] [CrossRef]
  169. Fang, L.; Chow, K.M.; Hou, S.; Xue, L.; Chen, X.; Rodgers, D.W.; Zheng, F.; Zhan, C.G. Rational design, preparation, and characterization of a therapeutic enzyme mutant with improved stability and function for cocaine detoxification. ACS Chem. Biol. 2014, 9, 1764–1772. [Google Scholar] [CrossRef]
  170. Agarwal, V.; Lin, S.; Lukk, T.; Nair, S.K.; Cronan, J.E. Structure of the enzyme-acyl carrier protein (ACP) substrate gatekeeper complex required for biotin synthesis. Proc. Natl. Acad. Sci. USA 2012, 109, 17406–17411. [Google Scholar] [CrossRef] [Green Version]
  171. Jansson, A.; Niemi, J.; Mantsala, P.; Schneider, G. Crystal structure of aclacinomycin methylesterase with bound product analogues: Implications for anthracycline recognition and mechanism. J. Biol. Chem. 2003, 278, 39006–39013. [Google Scholar] [CrossRef] [Green Version]
  172. Han, X.; Liu, W.; Huang, J.W.; Ma, J.; Zheng, Y.; Ko, T.P.; Xu, L.; Cheng, Y.S.; Chen, C.C.; Guo, R.T. Structural insight into catalytic mechanism of pet hydrolase. Nat. Commun. 2017, 8, 2106. [Google Scholar] [CrossRef] [Green Version]
  173. Joo, S.; Cho, I.J.; Seo, H.K.; Son, H.F.; Sagong, H.Y.; Shin, T.J.; Choi, S.Y.; Lee, S.Y.; Kim, K.J. Structural insight into molecular mechanism of poly(ethylene terephthalate) degradation. Nat. Commun. 2018, 9, 382. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  174. Liu, B.; He, L.; Wang, L.; Li, T.; Li, C.; Liu, H.; Luo, Y.; Bao, R. Protein crystallography and site-direct mutagenesis analysis of the poly(ethylene terephthalate) hydrolase petase from Ideonella sakaiensis. Chembiochem 2018, 19, 1471–1475. [Google Scholar] [CrossRef] [PubMed]
  175. Fecker, T.; Galaz-Davison, P.; Engelberger, F.; Narui, Y.; Sotomayor, M.; Parra, L.P.; Ramírez-Sarmiento, C.A. Active site flexibility as a hallmark for efficient PET degradation by I. sakaiensis PETase. Biophys. J. 2018, 114, 1302–1312. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  176. Liu, C.; Shi, C.; Zhu, S.; Wei, R.; Yin, C.C. Structural and functional characterization of polyethylene terephthalate hydrolase from Ideonella sakaiensis. Biochem. Biophys. Res. Commun. 2019, 508, 289–294. [Google Scholar] [CrossRef] [PubMed]
  177. Palm, G.J.; Reisky, L.; Bottcher, D.; Muller, H.; Michels, E.A.P.; Walczak, M.C.; Berndt, L.; Weiss, M.S.; Bornscheuer, U.T.; Weber, G. Structure of the plastic-degrading Ideonella sakaiensis mhetase bound to a substrate. Nat. Commun. 2019, 10, 1717. [Google Scholar] [CrossRef]
  178. Ali, Y.B.; Verger, R.; Abousalham, A. Lipases or esterases: Does it really matter? Toward a new bio-physico-chemical classification. In Lipases and Phospholipases: Methods and Protocols; Sandoval, G., Ed.; Humana Press: Totowa, NJ, USA, 2012; Volume 861, pp. 31–51. [Google Scholar]
  179. Jaeger, K.E.; Ransac, S.; Koch, H.B.; Ferrato, F.; Dijkstra, B.W. Topological characterization and modeling of the 3D structure of lipase from Pseudomonas aeruginosa. FEBS Lett. 1993, 332, 143–149. [Google Scholar] [CrossRef] [Green Version]
  180. Lesuisse, E.; Schanck, K.; Colson, C. Purification and preliminary characterization of the extracellular lipase of Bacillus subtilis 168, an extremely basic pH-tolerant enzyme. Eur. J. Biochem. 1993, 216, 155–160. [Google Scholar] [CrossRef]
  181. Cantarel, B.L.; Coutinho, P.M.; Rancurel, C.; Bernard, T.; Lombard, V.; Henrissat, B. The carbohydrate-active enzymes database (CAZy): An expert resource for glycogenomics. Nucleic Acids Res. 2009, 37, D233–D238. [Google Scholar] [CrossRef]
  182. Lombard, V.; Golaconda Ramulu, H.; Drula, E.; Coutinho, P.M.; Henrissat, B. The carbohydrate-active enzymes database (CAZy) in 2013. Nucleic Acids Res. 2014, 42, D490–D495. [Google Scholar] [CrossRef] [Green Version]
  183. Ray, S.K.; Rajeshwari, R.; Sonti, R.V. Mutants of Xanthomonas oryzae pv. Oryzae deficient in general secretory pathway are virulence deficient and unable to secrete xylanase. Mol. Plant Microbe Interact. 2000, 13, 394–401. [Google Scholar] [CrossRef] [Green Version]
  184. Sugiyama, M.; Ohtani, K.; Izuhara, M.; Koike, T.; Suzuki, K.; Imamura, S.; Misaki, H. A novel prokaryotic phospholipase A2. Characterization, gene cloning, and solution structure. J. Biol. Chem. 2002, 277, 20051–20058. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  185. Pei, J.; Grishin, N.V. PROMALS3D: Multiple protein sequence alignment enhanced with evolutionary and three-dimensional structural information. Methods Mol. Biol. 2014, 1079, 263–271. [Google Scholar] [PubMed] [Green Version]
  186. Pei, J.; Kim, B.H.; Grishin, N.V. PROMALS3D: A tool for multiple protein sequence and structure alignments. Nucleic Acids Res. 2008, 36, 2295–2300. [Google Scholar] [CrossRef] [PubMed]
  187. Pei, J.; Tang, M.; Grishin, N.V. PROMALS3D web server for accurate multiple protein sequence and structure alignments. Nucleic Acids Res. 2008, 36, W30–W34. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  188. Kumar, S.; Stecher, G.; Li, M.; Knyaz, C.; Tamura, K. MEGA X: Molecular evolutionary genetics analysis across computing platforms. Mol. Biol. Evol. 2018, 35, 1547–1549. [Google Scholar] [CrossRef]
  189. Rauwerdink, A.; Kazlauskas, R.J. How the same core catalytic machinery catalyzes 17 different reactions: The serine-histidine-aspartate catalytic triad of α/β-hydrolase fold enzymes. ACS Catal. 2015, 5, 6153–6176. [Google Scholar] [CrossRef] [Green Version]
  190. Dodson, G.; Wlodawer, A. Catalytic triads and their relatives. Trends Biochem. Sci. 1998, 23, 347–352. [Google Scholar] [CrossRef]
  191. Gariev, I.A.; Varfolomeev, S.D. Hierarchical classification of hydrolases catalytic sites. Bioinformatics 2006, 22, 2574–2576. [Google Scholar] [CrossRef]
  192. DeLano, W.L. The Pymol Molecular Graphics System; Delano Scientific: San Carlos, CA, USA, 2002. [Google Scholar]
  193. Walker, I.; Easton, C.J.; Ollis, D.L. Site-directed mutagenesis of dienelactone hydrolase produces dienelactone isomerase. ChemComm 2000, 671–672. [Google Scholar] [CrossRef]
  194. Borrelli, G.M.; Trono, D. Recombinant lipases and phospholipases and their use as biocatalysts for industrial applications. Int. J. Mol. Sci. 2015, 16, 20774–20840. [Google Scholar] [CrossRef] [Green Version]
  195. Robert, X.; Gouet, P. Deciphering key features in protein structures with the new endscript server. Nucleic Acids Res. 2014, 42, W320–W324. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  196. Akoh, C.C.; Lee, G.C.; Liaw, Y.C.; Huang, T.H.; Shaw, J.F. GDSL family of serine esterases/lipases. Prog. Lipid Res. 2004, 43, 534–552. [Google Scholar] [CrossRef] [PubMed]
  197. Goffin, C.; Ghuysen, J.M. Biochemistry and comparative genomics of SXXK superfamily acyltransferases offer a clue to the mycobacterial paradox: Presence of penicillin-susceptible target proteins versus lack of efficiency of penicillin as therapeutic agent. Microbiol. Mol. Biol. Rev. 2002, 66, 702–738. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  198. Arpigny, J.L.; Jaeger, K.E. Bacterial lipolytic enzymes: Classification and properties. Biochem. J. 1999, 343 Pt 1, 177–183. [Google Scholar] [CrossRef]
  199. Hausmann, S.; Jaeger, K.-E. Lipolytic enzymes from bacteria. In Handbook of Hydrocarbon and Lipid Microbiology; Timmis, K.N., Ed.; Springer: Berlin/Heidelberg, Germany, 2010; pp. 1099–1126. [Google Scholar]
  200. Wang, X.; Jiang, F.; Zheng, J.; Chen, L.; Dong, J.; Sun, L.; Zhu, Y.; Liu, B.; Yang, J.; Yang, G.; et al. The outer membrane phospholipase A is essential for membrane integrity and type III secretion in Shigella flexneri. Open Biol. 2016, 6, 160073. [Google Scholar] [CrossRef] [Green Version]
  201. Istivan, T.S.; Coloe, P.J. Phospholipase A in gram-negative bacteria and its role in pathogenesis. Microbiology 2006, 152, 1263–1274. [Google Scholar] [CrossRef] [Green Version]
  202. Van Loo, B.; Kingma, J.; Arand, M.; Wubbolts, M.G.; Janssen, D.B. Diversity and biocatalytic potential of epoxide hydrolases identified by genome analysis. Appl. Environ. Microbiol. 2006, 72, 2905–2917. [Google Scholar] [CrossRef] [Green Version]
  203. Davies, G.; Henrissat, B. Structures and mechanisms of glycosyl hydrolases. Structure 1995, 3, 853–859. [Google Scholar] [CrossRef] [Green Version]
  204. Burroughs, A.M.; Aravind, L. The origin and evolution of release factors: Implications for translation termination, ribosome rescue, and quality control pathways. Int. J. Mol. Sci. 2019, 20, 1981. [Google Scholar] [CrossRef] [Green Version]
  205. Burke, J.E.; Dennis, E.A. Phospholipase A2 structure/function, mechanism, and signaling. J. Lipid Res. 2009, 50, S237–S242. [Google Scholar] [CrossRef] [Green Version]
  206. Crichton, R.R. Chapter 12—Zinc—lewis acid and gene regulator. In Biological Inorganic Chemistry, 2nd ed.; Crichton, R.R., Ed.; Elsevier: Oxford, UK, 2012; pp. 229–246. [Google Scholar]
  207. Dawson, N.L.; Lewis, T.E.; Das, S.; Lees, J.G.; Lee, D.; Ashford, P.; Orengo, C.A.; Sillitoe, I. CATH: An expanded resource to predict protein function through structure and sequence. Nucleic Acids Res. 2016, 45, D289–D295. [Google Scholar] [CrossRef] [PubMed]
  208. Chen, Y.; Black, D.S.; Reilly, P.J. Carboxylic ester hydrolases: Classification and database derived from their primary, secondary, and tertiary structures. Protein Sci. 2016, 25, 1942–1953. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  209. Kourist, R.; Jochens, H.; Bartsch, S.; Kuipers, R.; Padhi, S.K.; Gall, M.; Bottcher, D.; Joosten, H.J.; Bornscheuer, U.T. The alpha/beta-hydrolase fold 3DM database (ABHDB) as a tool for protein engineering. Chembiochem 2010, 11, 1635–1643. [Google Scholar] [CrossRef] [PubMed]
  210. Heikinheimo, P.; Goldman, A.; Jeffries, C.; Ollis, D.L. Of barn owls and bankers: A lush variety of alpha/beta hydrolases. Structure 1999, 7, R141–R146. [Google Scholar] [CrossRef] [Green Version]
  211. Korman, T.P.; Sahachartsiri, B.; Charbonneau, D.M.; Huang, G.L.; Beauregard, M.; Bowie, J.U. Dieselzymes: Development of a stable and methanol tolerant lipase for biodiesel production by directed evolution. Biotechnol. Biofuels 2013, 6, 70. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  212. Solomon, K.V.; Haitjema, C.H.; Thompson, D.A.; O’Malley, M.A. Extracting data from the muck: Deriving biological insight from complex microbial communities and non-model organisms with next generation sequencing. Curr. Opin. Biotechnol. 2014, 28, 103–110. [Google Scholar] [CrossRef]
  213. Mukherjee, S.; Seshadri, R.; Varghese, N.J.; Eloe-Fadrosh, E.A.; Meier-Kolthoff, J.P.; Goker, M.; Coates, R.C.; Hadjithomas, M.; Pavlopoulos, G.A.; Paez-Espino, D.; et al. 1003 reference genomes of bacterial and archaeal isolates expand coverage of the tree of life. Nat. Biotechnol. 2017, 35, 676–683. [Google Scholar] [CrossRef]
  214. Lewin, H.A.; Robinson, G.E.; Kress, W.J.; Baker, W.J.; Coddington, J.; Crandall, K.A.; Durbin, R.; Edwards, S.V.; Forest, F.; Gilbert, M.T.P.; et al. Earth biogenome project: Sequencing life for the future of life. Proc. Natl. Acad. Sci. USA 2018, 115, 4325–4333. [Google Scholar] [CrossRef] [Green Version]
  215. Mukherjee, S.; Stamatis, D.; Bertsch, J.; Ovchinnikova, G.; Katta, H.Y.; Mojica, A.; Chen, I.A.; Kyrpides, N.C.; Reddy, T. Genomes Online database (GOLD) v.7: Updates and new features. Nucleic Acids Res. 2019, 47, D649–D659. [Google Scholar] [CrossRef]
  216. Koutsandreas, T.; Ladoukakis, E.; Pilalis, E.; Zarafeta, D.; Kolisis, F.N.; Skretas, G.; Chatziioannou, A.A. Anastasia: An automated metagenomic analysis pipeline for novel enzyme discovery exploiting next generation sequencing data. Front. Genet. 2019, 10, 469. [Google Scholar] [CrossRef] [Green Version]
  217. Prosser, G.A.; Larrouy-Maumus, G.; de Carvalho, L.P. Metabolomic strategies for the identification of new enzyme functions and metabolic pathways. EMBO Rep. 2014, 15, 657–669. [Google Scholar] [CrossRef] [PubMed]
  218. Von Grotthuss, M.; Plewczynski, D.; Vriend, G.; Rychlewski, L. 3D-Fun: Predicting enzyme function from structure. Nucleic Acids Res. 2008, 36, W303–W307. [Google Scholar] [CrossRef] [PubMed]
  219. Latino, D.A.; Aires-de-Sousa, J. Assignment of EC numbers to enzymatic reactions with MOLMAP reaction descriptors and random forests. J. Chem. Inf. Model. 2009, 49, 1839–1846. [Google Scholar] [CrossRef] [PubMed]
  220. Hu, Q.N.; Zhu, H.; Li, X.; Zhang, M.; Deng, Z.; Yang, X.; Deng, Z. Assignment of EC numbers to enzymatic reactions with reaction difference fingerprints. PLoS ONE 2012, 7, e52901. [Google Scholar] [CrossRef] [PubMed]
  221. Rahman, S.A.; Cuesta, S.M.; Furnham, N.; Holliday, G.L.; Thornton, J.M. EC-BLAST: A tool to automatically search and compare enzyme reactions. Nat. Methods 2014, 11, 171–174. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  222. Ryu, J.Y.; Kim, H.U.; Lee, S.Y. Deep learning enables high-quality and high-throughput prediction of enzyme commission numbers. Proc. Natl. Acad. Sci. USA 2019, 116, 13996–14001. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  223. Gerlt, J.A. Genomic enzymology: Web tools for leveraging protein family sequence-function space and genome context to discover novel functions. Biochemistry 2017, 56, 4293–4308. [Google Scholar] [CrossRef] [Green Version]
  224. Martinez-Martinez, M.; Coscolin, C.; Santiago, G.; Chow, J.; Stogios, P.J.; Bargiela, R.; Gertler, C.; Navarro-Fernandez, J.; Bollinger, A.; Thies, S.; et al. Determinants and prediction of esterase substrate promiscuity patterns. ACS Chem. Biol. 2018, 13, 225–234. [Google Scholar] [CrossRef] [Green Version]
  225. Martinez Cuesta, S.; Rahman, S.A.; Furnham, N.; Thornton, J.M. The classification and evolution of enzyme function. Biophys. J. 2015, 109, 1082–1086. [Google Scholar] [CrossRef] [Green Version]
  226. Wadhams, G.H.; Armitage, J.P. Making sense of it all: Bacterial chemotaxis. Nat. Rev. Mol. Cell Biol. 2004, 5, 1024–1037. [Google Scholar] [CrossRef]
  227. Miller, M.B.; Bassler, B.L. Quorum sensing in bacteria. Annu. Rev. Microbiol. 2001, 55, 165–199. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  228. Rutherford, S.T.; Bassler, B.L. Bacterial quorum sensing: Its role in virulence and possibilities for its control. Cold Spring Harb. Perspect. Med. 2012, 2, a012427. [Google Scholar] [CrossRef] [PubMed]
  229. Wang, L.H.; Weng, L.X.; Dong, Y.H.; Zhang, L.H. Specificity and enzyme kinetics of the quorum-quenching N-acyl homoserine lactone lactonase (AHL-lactonase). J. Biol. Chem. 2004, 279, 13645–13651. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  230. Papenfort, K.; Bassler, B.L. Quorum sensing signal-response systems in gram-negative bacteria. Nat. Rev. Microbiol. 2016, 14, 576–588. [Google Scholar] [CrossRef]
  231. Dong, Y.H.; Wang, L.H.; Xu, J.L.; Zhang, H.B.; Zhang, X.F.; Zhang, L.H. Quenching quorum-sensing-dependent bacterial infection by an N-acyl homoserine lactonase. Nature 2001, 411, 813–817. [Google Scholar] [CrossRef]
  232. Remy, B.; Mion, S.; Plener, L.; Elias, M.; Chabriere, E.; Daude, D. Interference in bacterial quorum sensing: A biopharmaceutical perspective. Front. Pharmacol. 2018, 9, 203. [Google Scholar] [CrossRef]
  233. Chiurchiù, V.; Leuti, A.; Maccarrone, M. Bioactive lipids and chronic inflammation: Managing the fire within. Front. Immunol. 2018, 9, 38. [Google Scholar] [CrossRef] [Green Version]
  234. Bonventre, J.V. Phospholipase A2 and signal transduction. J. Am. Soc. Nephrol. 1992, 3, 128–150. [Google Scholar]
  235. Rameshwaram, N.R.; Singh, P.; Ghosh, S.; Mukhopadhyay, S. Lipid metabolism and intracellular bacterial virulence: Key to next-generation therapeutics. Future Microbiol. 2018, 13, 1301–1328. [Google Scholar] [CrossRef]
  236. Tilg, H.; Zmora, N.; Adolph, T.E.; Elinav, E. The intestinal microbiota fuelling metabolic inflammation. Nat. Rev. Immunol. 2019, 1–15. [Google Scholar] [CrossRef]
  237. Makki, K.; Deehan, E.C.; Walter, J.; Backhed, F. The impact of dietary fiber on gut microbiota in host health and disease. Cell Host Microbe 2018, 23, 705–715. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  238. Keiler, K.C.; Feaga, H.A. Resolving nonstop translation complexes is a matter of life or death. J. Bacteriol. 2014, 196, 2123–2130. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  239. Keiler, K.C. Mechanisms of ribosome rescue in bacteria. Nat. Rev. Microbiol. 2015, 13, 285–297. [Google Scholar] [CrossRef] [PubMed]
  240. Littlechild, J.A. Enzymes from extreme environments and their industrial applications. Front. Bioeng. Biotechnol. 2015, 3, 161. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  241. Chapman, J.; Ismail, A.E.; Dinu, C.Z. Industrial applications of enzymes: Recent advances, techniques, and outlooks. Catalysts 2018, 8, 238. [Google Scholar] [CrossRef] [Green Version]
  242. Sheldon, R.A.; Woodley, J.M. Role of biocatalysis in sustainable chemistry. Chem. Rev. 2018, 118, 801–838. [Google Scholar] [CrossRef]
  243. Sheldon, R.A. Biocatalysis and green chemistry. In Green Biocatalysis, 1st ed.; Patel, R.N., Ed.; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2016; pp. 1–15. [Google Scholar]
  244. Sweeney, M.D.; Xu, F. Biomass converting enzymes as industrial biocatalysts for fuels and chemicals: Recent developments. Catalysts 2012, 2, 244–263. [Google Scholar] [CrossRef] [Green Version]
  245. Yeoman, C.J.; Han, Y.; Dodd, D.; Schroeder, C.M.; Mackie, R.I.; Cann, I.K.O. Chapter 1—Thermostable enzymes as biocatalysts in the biofuel industry. In Advances in Applied Microbiology; Academic Press: San Diego, CA, USA, 2010; Volume 70, pp. 1–55. [Google Scholar]
  246. Mukherjee, J.; Gupta, M.N. Biocatalysis for biomass valorization. Sustain. Chem. Process. 2015, 3, 7. [Google Scholar] [CrossRef]
  247. Haile, M.; Kang, W.H. The role of microbes in coffee fermentation and their impact on coffee quality. J. Food Qual. 2019, 2019, 6. [Google Scholar] [CrossRef]
  248. Chen, C.; Zhao, S.; Hao, G.; Yu, H.; Tian, H.; Zhao, G. Role of lactic acid bacteria on the yogurt flavour: A review. Int. J. Food Prop. 2017, 20, S316–S330. [Google Scholar] [CrossRef] [Green Version]
  249. Holland, R.; Liu, S.Q.; Crow, V.L.; Delabre, M.L.; Lubbers, M.; Bennett, M.; Norris, G. Esterases of lactic acid bacteria and cheese flavour: Milk fat hydrolysis, alcoholysis and esterification. Int. Dairy J. 2005, 15, 711–718. [Google Scholar] [CrossRef]
  250. Kawai, F.; Kawabata, T.; Oda, M. Current knowledge on enzymatic PET degradation and its possible application to waste stream management and other fields. Appl. Microbiol. Biotechnol. 2019, 103, 4253–4268. [Google Scholar] [CrossRef] [Green Version]
  251. Bornscheuer, U.T.; Huisman, G.W.; Kazlauskas, R.J.; Lutz, S.; Moore, J.C.; Robins, K. Engineering the third wave of biocatalysis. Nature 2012, 485, 185–194. [Google Scholar] [CrossRef] [PubMed]
  252. Verma, M.L.; Naebe, M.; Barrow, C.J.; Puri, M. Enzyme immobilisation on amino-functionalised multi-walled carbon nanotubes: Structural and biocatalytic characterisation. PLoS ONE 2013, 8, e73642. [Google Scholar] [CrossRef] [PubMed]
  253. Cao, L.; van Rantwijk, F.; Sheldon, R.A. Cross-linked enzyme aggregates: A simple and effective method for the immobilization of penicillin acylase. Org. Lett. 2000, 2, 1361–1364. [Google Scholar] [CrossRef] [PubMed]
  254. Su, F.; Li, G.; Fan, Y.; Yan, Y. Enhanced performance of lipase via microcapsulation and its application in biodiesel preparation. Sci. Rep. 2016, 6, 29670. [Google Scholar] [CrossRef] [Green Version]
  255. Mingarro, I.; Gonzalez-Navarro, H.; Braco, L. Trapping of different lipase conformers in water-restricted environments. Biochemistry 1996, 35, 9935–9944. [Google Scholar] [CrossRef]
  256. Singh, R.K.; Tiwari, M.K.; Singh, R.; Lee, J.K. From protein engineering to immobilization: Promising strategies for the upgrade of industrial enzymes. Int. J. Mol. Sci. 2013, 14, 1232–1277. [Google Scholar] [CrossRef]
  257. Amin, S.R.; Erdin, S.; Ward, R.M.; Lua, R.C.; Lichtarge, O. Prediction and experimental validation of enzyme substrate specificity in protein structures. Proc. Natl. Acad. Sci. USA 2013, 110, E4195–E4202. [Google Scholar] [CrossRef] [Green Version]
  258. Kingsley, L.J.; Lill, M.A. Substrate tunnels in enzymes: Structure–function relationships and computational methodology. Proteins 2015, 83, 599–611. [Google Scholar] [CrossRef] [Green Version]
  259. Chen, L.; Vitkup, D. Distribution of orphan metabolic activities. Trends Biotechnol. 2007, 25, 343–348. [Google Scholar] [CrossRef] [PubMed]
  260. Yamada, T.; Waller, A.S.; Raes, J.; Zelezniak, A.; Perchat, N.; Perret, A.; Salanoubat, M.; Patil, K.R.; Weissenbach, J.; Bork, P. Prediction and identification of sequences coding for orphan enzymes using genomic and metagenomic neighbours. Mol. Syst. Biol. 2012, 8, 581. [Google Scholar] [CrossRef] [PubMed]
  261. Hadadi, N.; MohammadiPeyhani, H.; Miskovic, L.; Seijo, M.; Hatzimanikatis, V. Enzyme annotation for orphan and novel reactions using knowledge of substrate reactive sites. Proc. Natl. Acad. Sci. USA 2019, 116, 7298–7307. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. Distribution of the resolution from 424 crystals and the genera of the 136 functionally-annotated CEHs from the PDB are described. (A) The distribution of crystal resolution is shown as a histogram, and the number on top of each bar means the number of crystals in the defined range of resolution. (B) The distribution of genera of CEHs is shown using circle graph. The numbers beside the genera indicate the numbers of CEHs in each genus. When groups have fewer than three members, all are assigned to others.
Figure 1. Distribution of the resolution from 424 crystals and the genera of the 136 functionally-annotated CEHs from the PDB are described. (A) The distribution of crystal resolution is shown as a histogram, and the number on top of each bar means the number of crystals in the defined range of resolution. (B) The distribution of genera of CEHs is shown using circle graph. The numbers beside the genera indicate the numbers of CEHs in each genus. When groups have fewer than three members, all are assigned to others.
Crystals 09 00597 g001
Figure 2. Phylogenetic tree of CEH sequences of all 136 CEHs are aligned and analyzed using a neighbor-joining method and the Jones–Taylor–Thornton (JTT) substitution model. Active site residue-based classification of CEHs is described using a combination of shape and color. Pink-to-purple closed circle: serine hydrolases (Group 1); orangish closed square: aspartidyl hydrolase (Group 2); greenish closed diamond: metal-independent non-serine hydrolases with water molecule as a nucleophile (Group 3); blueish closed triangle: metal-dependent metallohydrolases (Group 4). Different colors are used to distinguish reaction mechanisms. No symbol label means no information is available on their catalytic reaction. Localization of CEHs is described using the following markers at the outside of PDB codes in the phylogenetic tree: Asterisks (*) denote CEHs with signal peptide for secretion and dollar signs ($) denote membrane protein CEHs. CEHs without a marker are cytosolic CEHs. The figure is prepared using PROMALS3D [185,186,187] for sequence alignment and MEGA X [188] for phylogenetic tree.
Figure 2. Phylogenetic tree of CEH sequences of all 136 CEHs are aligned and analyzed using a neighbor-joining method and the Jones–Taylor–Thornton (JTT) substitution model. Active site residue-based classification of CEHs is described using a combination of shape and color. Pink-to-purple closed circle: serine hydrolases (Group 1); orangish closed square: aspartidyl hydrolase (Group 2); greenish closed diamond: metal-independent non-serine hydrolases with water molecule as a nucleophile (Group 3); blueish closed triangle: metal-dependent metallohydrolases (Group 4). Different colors are used to distinguish reaction mechanisms. No symbol label means no information is available on their catalytic reaction. Localization of CEHs is described using the following markers at the outside of PDB codes in the phylogenetic tree: Asterisks (*) denote CEHs with signal peptide for secretion and dollar signs ($) denote membrane protein CEHs. CEHs without a marker are cytosolic CEHs. The figure is prepared using PROMALS3D [185,186,187] for sequence alignment and MEGA X [188] for phylogenetic tree.
Crystals 09 00597 g002
Figure 3. Representative structures of CEHs in groups based on active site are described in 3D structures. The colors in Figure 2 are applied in the cartoon structures of (A) Group 1-1, (B) Group 1-2, (C) Group 1-3, (D) Group 1-4, (E) Group 2, (F) Group 3-1, (G) Group 3-2, (H) Group 3-3, (I) Group 3-4, (J) Group 3-5, (K) Group 4-1 and (L) Group 4-2. Catalytic residues, oxyanion hole residues and important residues are depicted using a sticks model with light gray. The catalytic important water molecule is shown with a red sphere, and metal ions are shown with blue-gray spheres. The names of functional core motifs are depicted near their composing residues. The four-digit PDB codes of the models are noted in bottom-right corner of each panel. Structures are visualized using PyMOL software [192].
Figure 3. Representative structures of CEHs in groups based on active site are described in 3D structures. The colors in Figure 2 are applied in the cartoon structures of (A) Group 1-1, (B) Group 1-2, (C) Group 1-3, (D) Group 1-4, (E) Group 2, (F) Group 3-1, (G) Group 3-2, (H) Group 3-3, (I) Group 3-4, (J) Group 3-5, (K) Group 4-1 and (L) Group 4-2. Catalytic residues, oxyanion hole residues and important residues are depicted using a sticks model with light gray. The catalytic important water molecule is shown with a red sphere, and metal ions are shown with blue-gray spheres. The names of functional core motifs are depicted near their composing residues. The four-digit PDB codes of the models are noted in bottom-right corner of each panel. Structures are visualized using PyMOL software [192].
Crystals 09 00597 g003
Figure 4. Consensus sequence blocks in the GDSX superfamily. Conserved regions in sequences of the GDSX motif-containing CEHs are described in black-outlined boxes. Assigned numbers above each box are the position of residues in CjCE2A (PDB code: 2WAA) as representative GDSX motif CEHs, and oxyanion hole-forming residues are highlighted using asterisks below the boxes. Fully conserved positions are marked by red shading, and highly conserved positions (>70%) are highlighted in yellow. The conserved residues with high similarity are in bold. PROMALS3D [185,186,187] for sequence alignment and ESPript 3 [195] for visualization were used.
Figure 4. Consensus sequence blocks in the GDSX superfamily. Conserved regions in sequences of the GDSX motif-containing CEHs are described in black-outlined boxes. Assigned numbers above each box are the position of residues in CjCE2A (PDB code: 2WAA) as representative GDSX motif CEHs, and oxyanion hole-forming residues are highlighted using asterisks below the boxes. Fully conserved positions are marked by red shading, and highly conserved positions (>70%) are highlighted in yellow. The conserved residues with high similarity are in bold. PROMALS3D [185,186,187] for sequence alignment and ESPript 3 [195] for visualization were used.
Crystals 09 00597 g004
Figure 5. Sequences of peptidyl-tRNA hydrolases are aligned, and functionally important areas are described. The numbers above the sequences are the residual numbers of peptidyl-tRNA hydrolase from V. cholerae (PDB 4Z86). Catalytic His and Asp are marked using asterisks, substrate-binding residues are marked using S and oxyanion hole-forming Asn residues are marked using O beneath the sequences. Fully conserved positions are marked with red shading, and highly conserved positions (>70%) are highlighted by yellow shading. The conserved residues with high similarity are in bold. PROMALS3D [185,186,187] for sequence alignment and ESPript 3 [195] for visualization were used.
Figure 5. Sequences of peptidyl-tRNA hydrolases are aligned, and functionally important areas are described. The numbers above the sequences are the residual numbers of peptidyl-tRNA hydrolase from V. cholerae (PDB 4Z86). Catalytic His and Asp are marked using asterisks, substrate-binding residues are marked using S and oxyanion hole-forming Asn residues are marked using O beneath the sequences. Fully conserved positions are marked with red shading, and highly conserved positions (>70%) are highlighted by yellow shading. The conserved residues with high similarity are in bold. PROMALS3D [185,186,187] for sequence alignment and ESPript 3 [195] for visualization were used.
Crystals 09 00597 g005
Figure 6. The sequences of lactonases containing the HXHXDH motif are aligned. The numbers above the aligned sequences are the residue positions in N-acyl lactonase from Bacillus thuringiensis (PDB code: 2A7M). The position, which is close to the first and the second metal ions, is marked bottom of the alignment using 1 (the first) and 2 (the second). Fully conserved positions are marked by red shading, and highly conserved positions (>70%) are in yellow. The conserved residues with high similarity are in bold. PROMALS3D [185,186,187] for sequence alignment and ESPript 3 [195] for visualization are used.
Figure 6. The sequences of lactonases containing the HXHXDH motif are aligned. The numbers above the aligned sequences are the residue positions in N-acyl lactonase from Bacillus thuringiensis (PDB code: 2A7M). The position, which is close to the first and the second metal ions, is marked bottom of the alignment using 1 (the first) and 2 (the second). Fully conserved positions are marked by red shading, and highly conserved positions (>70%) are in yellow. The conserved residues with high similarity are in bold. PROMALS3D [185,186,187] for sequence alignment and ESPript 3 [195] for visualization are used.
Crystals 09 00597 g006
Figure 7. Various tertiary architectures of CEHs in PDB are described using rainbow colors in the direction of the N to C terminus. The 3D models are obtained from representative structures, which contain each architecture: (A) α/β/α sandwich (PDB code: 3FAK), (B) up-down bundle (PDB code: 1LWB), (C) 6-blade propeller (PDB code: 3DR2), (D) 7-blade propeller (PDB code: 6NAU), (E) β-barrel (PDB code: 5DQX), (F) α/β/β/α 4-layer (PDB code: 4KEP), (G) α/β barrel (PDB code: 4D8L), (H) 3-solenoid (PDB code: 2NSP). Structures are visualized using PyMOL software [192].
Figure 7. Various tertiary architectures of CEHs in PDB are described using rainbow colors in the direction of the N to C terminus. The 3D models are obtained from representative structures, which contain each architecture: (A) α/β/α sandwich (PDB code: 3FAK), (B) up-down bundle (PDB code: 1LWB), (C) 6-blade propeller (PDB code: 3DR2), (D) 7-blade propeller (PDB code: 6NAU), (E) β-barrel (PDB code: 5DQX), (F) α/β/β/α 4-layer (PDB code: 4KEP), (G) α/β barrel (PDB code: 4D8L), (H) 3-solenoid (PDB code: 2NSP). Structures are visualized using PyMOL software [192].
Crystals 09 00597 g007
Figure 8. Various topologies of α/β hydrolases in CEHs are described. (A) Classical α/β hydrolase fold with 1-2-4-3-5-6-7-8 order of β sheets. (B) A variant of α/β hydrolase fold with 1-3-2-4-5-6-7 order of β sheets. (C) Minimal α/β hydrolase fold with 2-1-3-4-5 order of β sheets. β sheets are described with rainbow colors (red to navy blue) by the order of β sheets, and all α helices are described with gray. Structures are visualized using PyMOL software [192].
Figure 8. Various topologies of α/β hydrolases in CEHs are described. (A) Classical α/β hydrolase fold with 1-2-4-3-5-6-7-8 order of β sheets. (B) A variant of α/β hydrolase fold with 1-3-2-4-5-6-7 order of β sheets. (C) Minimal α/β hydrolase fold with 2-1-3-4-5 order of β sheets. β sheets are described with rainbow colors (red to navy blue) by the order of β sheets, and all α helices are described with gray. Structures are visualized using PyMOL software [192].
Crystals 09 00597 g008
Table 1. List of 424 Protein Data Bank (PDB) codes in carboxylic ester hydrolases (CEHs).
Table 1. List of 424 Protein Data Bank (PDB) codes in carboxylic ester hydrolases (CEHs).
EC NumberEnzyme FunctionPDB Codes
EC 3.1.1.-carboxylic ester hydrolases1C7I[19], 1C7J[19], 1ESC[20], 1ESD[20], 1ESE[20], 1QE3[19], 2A7M[21], 2WAA[22], 2WYL[23], 2WYM[23], 3DHA[24], 3DHB[24], 3DHC[24], 3DNM[25], 3F67, 3FAK[26], 3G9T, 3G9U, 3G9Z, 3GA7, 3H17[27], 3H18[27], 3H19, 3H1A, 3H1B, 3H2G[28], 3H2H[28], 3H2I[28], 3H2J[28], 3H2K[28], 3I2F[29], 3I2G[29], 3I2H[29], 3I2I[29], 3I2J[29], 3I2K[29], 3IDA[30], 3K6K, 3L1H, 3L1I, 3L1J, 3LS21[31], 3PF8[32], 3PF9[32], 3PFB[32], 3PFC[32], 3PUH[33], 3PUI[33], 3QM1[32], 3S2Z[32], 4F21[34], 4KRX[35], 4KRY[35], 4OB6[36], 4OB7[36], 4OB8[36], 5A2G[37], 5HC0[38], 5HC2[38], 5HC3[38], 5HC4[38], 5HC5[38], 5MAL2[39], 5UGQ[40], 5UNO[40], 5UOH[40], 6EHN[41], 6GRY[42]
EC 3.1.1.1carboxylesterase1AUO[43], 1AUR[43], 1EVQ[44], 1L7Q[45], 1L7R[45], 1R1D, 1TQH[46], 2H1I, 2HM7[47], 2R11, 3CN7[48], 3CN9[48], 3DOH[49], 3DOI[49], 3KVN[50], 4BZW[51], 4BZZ[51], 4C01[51], 4C87[52], 4C88[52], 4C89[52], 4CCW[53], 4CCY[53], 4FHZ[54], 4FTW[54], 4IVI[55], 4IVK[55], 4JGG[56], 4OU4[36], 4OU5[36], 4ROT, 4UHC[57], 4UHD[57], 4UHE[57], 4UHF[57], 4UHH[57], 4V2I[58], 4YPV[59], 5AO9[60], 5AOA[60], 5AOB[60], 5AOC[60], 5DWD, 5EGN, 5GMX[61], 5H3B[62], 6AAE[63], 6IEY[63]
EC 3.1.1.2arylesterase1VA43[64], 2Q0Q[65], 2Q0S[65], 3HEA3[64,66], 3HI43[66], 3IA23[67], 3T4U3, 3T523, 4ROT, 4TX1[68]
EC 3.1.1.3triacylglycerol lipase1CVL[69], 1EX9[70], 1HQD[71], 1I6W[72], 1ISP4[73], 1JI3[74], 1KU0[75], 1OIL[76], 1QGE, 1R4Z[77], 1R50[77], 1T2N[78], 1T4M[78], 1TAH[78], 1YS1[79], 1YS2[79], 2DSN[80], 2ES4[81], 2FX5, 2HIH[82], 2LIP[83], 2NW6[84], 2ORY[85], 2QUA[86], 2QUB[86], 2QXT[87], 2QXU[87], 2W22[88], 2Z5G[80], 2Z8X[89], 2Z8Z[89], 2ZJ6[90], 2ZJ7[90], 2ZVD[91], 3A6Z[91], 3A70[91], 3AUK, 3D2A[92], 3D2B[92], 3D2C[92], 3LIP[83], 3QMM[93], 3QZU[94], 3UMJ[95], 3W9U, 4FDM[96], 4FKB, 4FMP, 4GW3[97], 4GXN[97], 4HS9[97], 4LIP[98], 4OPM, 4X6U[99], 4X71[99], 4X7B[99], 4X85[99], 5AH1[100], 5CE5[100], 5CRI[101], 5CT4[101], 5CT6[101], 5CT9[101], 5CTA[101], 5CUR[101], 5H6B[102], 5LIP[98], 5MAL2[39], 5XPX, 6A12[103], 6CL4[103], 6FZ1[104], 6FZ7[104], 6FZ8[104], 6FZ9[104], 6FZA[104], 6FZC[104], 6FZD[104]
EC 3.1.1.4phospholipase A21FAZ[105], 1KP4[105], 1LWB[106], 1QD5[107], 1QD6[107], 1FW2[108], 1FW3[108], 1ILD[109], 1ILZ[109], 1IM0[109], 5DQX
EC 3.1.1.5lysophospholipase1IVN5[110], 1J005[110], 1JRL5[110], 1U8U5[111], 1V2G5[111], 5TIC5[112], 5TID5[112], 5TIE5[112], 5TIF5[112]
EC 3.1.1.6acetylesterase2XLB[113], 2XLC[113], 3FVR, 3FVT, 3FYT, 3FYU, 4NS4
EC 3.1.1.11pectinesterase1QJV[114], 2NSP[115], 2NST[115], 2NT6[115], 2NT9[115], 2NTB[115], 2NTP[115], 2NTQ[115], 3UW0[116]
EC 3.1.1.17gluconolactonase3DR2[117]
EC 3.1.1.20tannase3WA6[118], 3WA7[118]
EC 3.1.1.23acylglycerol lipase3RLI[119], 3RM3[119], 4KE6[120], 4KE7[120], 4KE8[120], 4KE9[120], 4KEA[120], 4LHE[121], 5XKS
EC 3.1.1.243-oxoadipate enol-lactonase2XUA[122]
EC 3.1.1.251,4-lactonase3MSR, 3OVG
EC 3.1.1.274-pyridoxolactonase3AJ3, 4KEP, 4KEQ
EC 3.1.1.29aminoacyl-tRNA hydrolase2PTH[123], 2Z2I[124], 2Z2J[124], 2Z2K[124], 3KJZ, 3KK0, 3NEA[125], 3OFV, 3P2J, 3TCK[126], 3TCN[126], 3TD2[126], 3TD6[126], 3V2I[127], 3VJR[128], 4DHW, 4DJJ, 4ERX, 4FNO[129], 4FOP[130], 4FOT[130], 4FYJ[131], 4HOY[130], 4IKO[130], 4JC4[129], 4JWK[130], 4JX9[130], 4JY7[130], 4LIP[98], 4LWQ, 4OLJ, 4P7B[132], 4QAJ[129], 4QBK[129], 4QD3[129], 4QT4[133], 4V95[134], 4YLY[135], 4Z86[136], 4ZXP[136], 5B6J[136], 5EKT, 5GVZ, 5IKE[136], 5IMB[136], 5IVP, 5Y98[137], 5Y9A[137], 5YL8, 5YLA, 5YN4, 5ZK0, 5ZX8[138], 5ZZV, 6A31, 6IVV, 6IX6, 6IYE, 6J93, 6JGU, 6JJ1, 6JJQ, 6JKX, 6JQT
EC 3.1.1.316-phosphoglucono-lactonase1PBT, 1VL1, 3ICO[139], 3LWD, 3OC6[139], 4TM7[140], 4TM8[140], 6NAU
EC 3.1.1.32phospholipase A11QD5[107], 1QD6[107], 1FW2[108], 1FW3[108], 1ILD[109], 1ILZ[109], 1IM0[109], 4HYQ[141], 5DQX
EC 3.1.1.41cephalosporin-C deacetylase1L7A, 1ODS[142], 1ODT[142], 1VLQ[143], 3FCY, 3M81[143], 3M82[143], 3M83[143], 5FDF[144], 5GMA[145], 5HFN[144], 5JIB[146]
EC 3.1.1.43alpha-amino-acid esterase1MPX[147], 1NX9[148], 1RYY[148], 2B4K[148], 2B9V[148]
EC 3.1.1.45carboxymethylene-butenolidase1DIN[149], 1GGV[150], 1ZI6[151], 1ZI8[151], 1ZI9[151], 1ZIC[151], 1ZIX[151], 1ZIY[151], 1ZJ4[151], 1ZJ5[151], 4P92[152], 4P93[152], 4U2B[153], 4U2C[153], 4U2D[153], 4U2E[153], 4U2F[153], 4U2G[153]
EC 3.1.1.572-pyrone-4,6-dicarboxylate lactonase4D8L[154]
EC 3.1.1.61protein-glutamate methylesterase1CHD[13], 3SFT[155], 3T8Y[156]
EC 3.1.1.72acetylxylan esterase2CC0[157], 2VPT6[158], 2WAA[22], 2XLB[113], 2XLC[113], 3FCY, 3W7V[159], 4JHL[159], 4JJ4, 4JJ6, 4JKO[159], 4OAO, 4OAP, 5BN1, 5FDF[144], 5GMA[145], 5HFN[144], 5JIB[146]
EC 3.1.1.73feruloyl esterase5YAE[160], 5YAL[160]
EC 3.1.1.74cutinase4CG1[161], 4CG2[161], 4CG3[161], 4EB0, 5ZOA, 5ZNO[162]
EC 3.1.1.75poly(3-hydroxybutyrate) depolymerase4BRS[163], 4BTV[163], 4BVJ[163], 4BVK[163], 4BVL[163], 4BYM[163]
EC 3.1.1.81quorum-quenching N-acyl-homoserine lactonase2A7M[21], 2BR6[164], 2BTN[164], 3DHA[165], 3DHB[165], 3DHC[165], 4J5F[166], 4J5H[166], 5EH9[167], 5EHT[167]
EC 3.1.1.84cocaine esterase1JU3[168], 1JU4[168], 1L7Q[45], 1L7R[45], 3IDA[30], 4P08[169]
EC 3.1.1.85pimeloyl-[acyl-carrier protein] methyl ester esterase4ETW[170], 4NMW
EC 3.1.1.95aclacinomycin methylesterase1Q0R[171], 1Q0Z[171]
EC 3.1.1.101poly(ethylene terephthalate) hydrolase5XFY[172], 5XFZ[172], 5XG0[172], 5XH2[172], 5XH3[172], 5XJH[173], 5YFE[174], 5YNS[173], 6ANE[175], 6EQD[174], 6EQE[174], 6EQF[174], 6EQG[174], 6EQH[174], 6ILW[176], 6ILX[176], 6QGC[177]
EC 3.1.1.102mono(ethylene terephthalate) hydrolase6QG9[177], 6QGA[177], 6QGB[177]
* PDB codes with superscript and underline have extra-assigned EC numbers corresponding to any additional activities of the CEHs. (1 = [EC 3.1.2.12]; 2 = [EC 3.1.2.2]; 3 [EC 1.-.-.-]; 4 = [EC 3.4.21.-]; 5 = [EC 3.1.2.2, EC 3.1.2.14, EC 3.4.21.- and EC 3.1.2.-]; and 6 = [EC 3.2.1.4]). ** Representative PDB codes are highlighted using bold characters. *** PDB codes in [EC 3.1.1.84] are the same proteins with 3I2K PDB code as a cocaine esterase. However, 3I2F, 3I2G, 3I2H, 3I2I, 3I2J and 3I2K were not assigned to [EC 3.1.1.84].
Table 2. Representative PDB Code List of CEH Clans.
Table 2. Representative PDB Code List of CEH Clans.
Clan TypesPDB Codes
Clan A1C7I[19], 1EVQ[44], 1L7A, 1Q0R[171], 2R11, 2XLB[113], 2XUA[122], 3FAK[26], 3FCY, 3GA7, 3K6K, 4CCW[53], 4CCY[53], 4CG1[161], 4EB0, 4KRX[35], 4OB7[36]
Clan B1CVL[69], 1ESC[20], 1EX9[70], 1JI3[74], 1KU0[75], 1QGE, 1YS1[79], 2DSN[80], 2HIH[82], 2W22[88], 3AUK, 4FDM[96], 4FKB, 4HS9[211], 4HYQ[141], 4X6U[99]
Clan C1AUO[43], 1TQH[46], 1ZI8 [151], 3CN9[48], 3F67, 3IA2[67], 3RLI[119], 4ETW[170], 4F21[34], 4FHZ[54], 4NMW, 4U2D[153], 4UHC[57]
Clan D3DR2[117]
Clan EN/A

Share and Cite

MDPI and ACS Style

Oh, C.; Kim, T.D.; Kim, K.K. Carboxylic Ester Hydrolases in Bacteria: Active Site, Structure, Function and Application. Crystals 2019, 9, 597. https://doi.org/10.3390/cryst9110597

AMA Style

Oh C, Kim TD, Kim KK. Carboxylic Ester Hydrolases in Bacteria: Active Site, Structure, Function and Application. Crystals. 2019; 9(11):597. https://doi.org/10.3390/cryst9110597

Chicago/Turabian Style

Oh, Changsuk, T. Doohun Kim, and Kyeong Kyu Kim. 2019. "Carboxylic Ester Hydrolases in Bacteria: Active Site, Structure, Function and Application" Crystals 9, no. 11: 597. https://doi.org/10.3390/cryst9110597

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