CAZypedia needs your help! We have many unassigned GH, PL, CE, AA, GT, and CBM pages in need of Authors and Responsible Curators.
Scientists at all career stages, including students, are welcome to contribute to CAZypedia. Read more here, and in the 10th anniversary article in Glycobiology.
New to the CAZy classification? Read this first.
*
Consider attending the 15th Carbohydrate Bioengineering Meeting in Ghent, 5-8 May 2024.

Difference between revisions of "Glycoside Hydrolase Family 3"

From CAZypedia
Jump to navigation Jump to search
Line 45: Line 45:
  
 
== Catalytic Residues ==
 
== Catalytic Residues ==
The catalytic amino acid residues of GH3 members was first informed by the three-dimensional structure of a barley β-D-glucan glucohydrolase, solved in complex with glucose bound in the -1 subsite <cite>Vargese1999</cite>.  Whereas the catalytic nucleophile is well conserved among GH3 members from diverse species, the catalytic acid/base has been difficult to generally identify.  In many cases, this residue is borne on a flexible loop structure, which exhibits significant sequence divergence and conformational flexibility.  The following sections detail the definitive assignments of the catalytic residues in a growing number of GH3 enzymes.
 
 
 
=== Catalytic nucleophile ===
 
=== Catalytic nucleophile ===
In plant family 3 β-D-glycosidases, typified by the barley enzyme, active site consists of two glucosyl-binding subsites (-1 and +1) and the catalytic amino acid residues are located between these two subsites <cite>Vargese1999 Hrmova2001</cite>.  The [[catalytic nucleophile]], as identified in the crystal structure of a trapped 2-fluoroglycosyl enzyme, is Asp285, which is located in a highly conserved GFVISDW motif <cite>Hrmova2001</cite>.
+
Early labeling experiments of a GH3 β-glucosidases from ''Aspergillus wentii'' using conduritol B-epoxide by Bause and Legler in 1974 <cite>Bause1974</cite> identified an aspartate residue within the sequence VMS'''D'''W as the putative catalytic nucleophile.  This was later supported by the crystal structure of barley β-D-glucan glucohydrolase, which identified Asp285 within the sequence GFVIS'''D'''W as being appropriately positioned to act as a nucleophile during catalysis <cite>Vargese1999</cite>.  Direct identification of this aspartate as the catalytic nucleophile via active-site labeling using 2-deoxy-2-fluoro-β-D-glycosides of GH3 β-glucosidases from ''A. niger ''<cite>Dan2000</cite> and ''F. meningosepticum ''<cite>Chir2002</cite>, a glucosylceramidase from ''Paenibacillus ''sp <cite>Paal2004</cite>, as well as the barley β-D-glucan glucohydrolase <cite>Hrmova2001</cite>, and labeling of NagZ enzymes from ''V. furnisii'' <cite>Vocadlo2000 </cite> and ''S.'' ''typhimerium'' <cite>Bacik2012</cite> using 2-acetamido-2-Deoxy-5-Fluoro-β-D-glucopyranosyl fluoride, have confirmed that this catalytic residue is well conserved in GH3 enzymes from diverse species. The conservation of this aspartate becomes clear when known crystal structures of a number of GH3 enzymes are superposed (Fig. 3). <br>
  
The [[catalytic nucleophile]] for ''Vibrio furnisii'' ExoII (a NagZ) has been identified chemically as the corresponding Asp242, which is conserved throughout the family 3 NagZ enzymes <cite>Vocadlo2000</cite>.
+
=== Catalytic acid/base ===
 +
Though sequence alignments can now be used to easily identify the catalytic nucleophile of GH3 members, the position of type of general acid/base residue is phylogenetically variable and less readily identifiable. An increase in 3D structural information for GH3 enzymes has recently provided insight into identity of the elusive general acid-base residue. The barley β-D-glucan glucohydrolase structure is a two-domain enzyme consisting of an N-terminal (b/a)<sub>8</sub> barrel domain housing the active site pocket and catalytic aspartate nucleophile (Asp285), and a C-terminal domain with a glutamate residue (Glu 491) that projects into the active site of (b/a)<sub>8</sub> barrel domain to act as the catalytic acid/base <cite>Vargese1999</cite>.  Structural studies of additional GH3 enzymes have since shown that while insertion of additional domains is possible (which shifts the location of the glutamic acid general acid-base within the primary sequence with respect to the barely enzyme), the unusual two-domain active site architecture observed for the barley enzyme appears to be a core feature of multidomain GH3 b-glucanases <cite>Pozzo2010 Nakatani2012 Yoshida2010</cite>. Superposition of available GH3 β-D-glucan glucohydrolase structures clearly reveals the conservation of this architecture (Fig. 3).
  
=== Catalytic acid/base ===
+
GH3 NagZ enzymes depart significantly from the above two-domain active site architecture. A crystal structure of NagZ from ''B. subtilis'' recently revealed that the catalytic acid-base is an unusual histidine/aspartate dyad that resides within a flexible loop that is present on the catalytic (b/a)<sub>8</sub> barrel and not on a separate domain <cite>Litzinger2010b</cite>. Though the enzyme adopts a two-domain fold similar to barely β-D-glucan glucohydrolase, the C-terminal domain does not participate in catalysis (Fig. 4).  In fact, most NagZ enzymes from Gram-negative bacteria are single domain enzymes comprised solely of a catalytic (b/a)<sub>8</sub> barrel that contains the conserved aspartate nucleophile and a catalytic histidine/aspartate dyad on flexible loop as seen for ''B. subtilis ''NagZ <cite>Bacik2012</cite> (Fig. 5).  The flexible loop containing the His/Asp dyad is unique to GH3 NagZ enzymes and can be identified by the consensus motif [KH(F/I)PG(H/L)GXXXX'''D'''(S/T)'''H'''] (catalytic dyad highlighted in boldface) <cite>Vocadlo2000 </cite>.
The [[general acid]], E491, is highly conserved in plant family 3 enzymes but is more difficult to locate in more distantly related members of the family <cite>Harvey2000</cite>.
 
  
A [[general acid]] residue has not been identified for family 3 NagZ enzymes.
 
  
 
== Three-dimensional structures ==
 
== Three-dimensional structures ==
Line 109: Line 106:
 
#Nakatani2012 pmid=22129429
 
#Nakatani2012 pmid=22129429
 
#Zmudka2013 pmid=23225731
 
#Zmudka2013 pmid=23225731
 +
#Yamaguchi2012 pmid=22844551
 
</biblio>
 
</biblio>
  
 
[[Category:Glycoside Hydrolase Families|GH003]]
 
[[Category:Glycoside Hydrolase Families|GH003]]

Revision as of 12:13, 10 April 2013

Approve icon-50px.png

This page has been approved by the Responsible Curator as essentially complete. CAZypedia is a living document, so further improvement of this page is still possible. If you would like to suggest an addition or correction, please contact the page's Responsible Curator directly by e-mail.


Glycoside Hydrolase Family GH3
Clan none
Mechanism retaining
Active site residues known
CAZy DB link
http://www.cazy.org/GH3.html


Substrate specificities

The GH3 glycoside hydrolase family currently group together β-D-glucosidases, α-L-arabinofuranosidases, β-D-xylopyranosidases and N-acetyl-β-D-glucosaminidases [1]. Widely distributed in bacteria, fungi and plants, GH3 enzymes carry out a range of functions including cellulosic biomass degradation, plant and bacterial cell wall remodeling, energy metabolism and pathogen defense. In many cases the enzymes have dual or broad substrate specificities with respect to monosaccharide residues, linkage position and chain length of the substrate. For example, there are several well-characterized ‘bifunctional’ enzymes in the family that have both α-L-arabinofuranosidase and β-D-xylopyranosidase activity [2], and one characterized example of an N-acetyl-β-D-glucosaminidase/β-glucosidase from Cellulomonas fimi (Nag3) [3]. GH3 β-D-glucosidases from barley, which are more precisely referred to as β-D-glucan glucohydrolases, are also broad specificity exo-hydrolases that remove single glucosyl residues from the non-reducing ends of a range of β-D-glucans, β-D-oligoglucosides and aryl β-D-glucosides, including (1,3)-β-D-glucans, (1,4)-β-D-glucans, (1,3;1,4)-β-D-glucans and (1,6)-β-D-glucans, 4-nitrophenyl β-D-glucoside, certain cyanogenic β-D-glucosides and some β-D-oligoxyloglucosides [4].

In contrast to the broad substrate specificities observed for the GH3 enzymes described above, GH3 N-acetyl-β-D-glucosaminidases are selective for N-acetyl-β-D-glucosamine (GlcNAc) [5] (though exceptions exist such as Cellulomonas fimi Nag3 [3]). A notable GH3 N-acetyl-β-D-glucosaminidase of prokaryotes is NagZ, which participates in bacterial cell wall recycling (for review see [6]) by removing GlcNAc from 1,6-anhydroMurNAc-peptides in Gram-negative bacteria [7], or GlcNAc from GlcNAc-MurNAc-peptides in Gram-positive bacteria[8], . The NagZ product 1,6-anhydroMurNAc-peptide is an important activator of AmpC β-lactamase hyper-production in several Gram-negative pathogens, making the enzyme of interest as a potential therapeutic target[9].

As more plant genome sequences are published, it is becoming clear that the GH3 glycoside hydrolases in plants are encoded by multi-gene families. In both monocot and dicot species, there are 15-18 genes for the GH3 enzymes, while in lower plants such as the Physcomitrella and Selaginella mosses 6-7 genes encoding GH3 glycoside hydrolases have been identified. Furthermore, phylogenetic analyses of the plant GH3 enzymes reveal their separation into two major groups (Fig. 1). Existing annotations suggest that one plant group might include β-D-glucosidases and the other α-L-arabinofuranosidases, β-D-xylopyranosidase and other enzymes. However, this suggestion is limited to plant GH3 enzymes and must be viewed with caution until further confirmation is obtained, because relatively few members of the enzyme family have been purified for detailed characterization of their substrate specificities. There do not appear to be any confirmed plant β-N-acetylglucosaminidases in either clade.


Kinetics and Mechanism

GH3 glycoside hydrolases remove single glycosyl residues from the non-reducing ends of their substrates. Catalysis occurs via a classical Koshland double-displacement mechanism with the β-anomeric configuration of the released glycose being retained. The retention of anomeric configuration has been established experimentally for several enzymes. The active site of GH3 enzymes consists of two glucosyl-binding subsites (-1 and +1) with an enzymic nucleophile and general acid/base residue flanking the junction of these two subsites (see [10] for subsite nomenclature). This arrangement of catalytic residues was first visualized in 1999 when the crystal structure of barley β-D-glucan glucohydrolase was determined in complex with glucose [11] (Fig. 2).

Several studies have contributed to the understanding of the kinetics and mechanism of GH3 enzymes, including detailed analyses of β-glucosidases from fungi (Aspergillus wentii [12] and Aspergillus niger [13, 14]) and Gram-negative bacteria (Flavobacterium meningosepticum [15, 16], Thermotoga neapolitana [17]), as well as a GH3 glucosylceramidase from the Gram-positive microbe Paenibacillus sp. TS12 [18]. Kinetic and mechanistic analyses of β-D-glucan glucohydrolases and two ‘bifunctional’ α-L-arabinofuranosidase/β-D-xylopyranosidases from plants (barley) have also been carried out [2, 4]. These studies, combined with kinetic and mechanistic analyses of N-acetyl-β-D-glucosaminidases from the Gram-positive microbe Bacilus subtillus [19, 20], and Gram-negative microbes Vibrio furnisii [21, 22], Vibrio cholerea [23] and Salmonella typhimerium [20] confirm that while the catalytic nucleophile of GH3 enzymes is well conserved, the location and identity of the general acid/base residue is not.

Catalytic Residues

Catalytic nucleophile

Early labeling experiments of a GH3 β-glucosidases from Aspergillus wentii using conduritol B-epoxide by Bause and Legler in 1974 [12] identified an aspartate residue within the sequence VMSDW as the putative catalytic nucleophile. This was later supported by the crystal structure of barley β-D-glucan glucohydrolase, which identified Asp285 within the sequence GFVISDW as being appropriately positioned to act as a nucleophile during catalysis [11]. Direct identification of this aspartate as the catalytic nucleophile via active-site labeling using 2-deoxy-2-fluoro-β-D-glycosides of GH3 β-glucosidases from A. niger [13] and F. meningosepticum [15], a glucosylceramidase from Paenibacillus sp [18], as well as the barley β-D-glucan glucohydrolase [24], and labeling of NagZ enzymes from V. furnisii [21] and S. typhimerium [20] using 2-acetamido-2-Deoxy-5-Fluoro-β-D-glucopyranosyl fluoride, have confirmed that this catalytic residue is well conserved in GH3 enzymes from diverse species. The conservation of this aspartate becomes clear when known crystal structures of a number of GH3 enzymes are superposed (Fig. 3).

Catalytic acid/base

Though sequence alignments can now be used to easily identify the catalytic nucleophile of GH3 members, the position of type of general acid/base residue is phylogenetically variable and less readily identifiable. An increase in 3D structural information for GH3 enzymes has recently provided insight into identity of the elusive general acid-base residue. The barley β-D-glucan glucohydrolase structure is a two-domain enzyme consisting of an N-terminal (b/a)8 barrel domain housing the active site pocket and catalytic aspartate nucleophile (Asp285), and a C-terminal domain with a glutamate residue (Glu 491) that projects into the active site of (b/a)8 barrel domain to act as the catalytic acid/base [11]. Structural studies of additional GH3 enzymes have since shown that while insertion of additional domains is possible (which shifts the location of the glutamic acid general acid-base within the primary sequence with respect to the barely enzyme), the unusual two-domain active site architecture observed for the barley enzyme appears to be a core feature of multidomain GH3 b-glucanases [17, 25, 26]. Superposition of available GH3 β-D-glucan glucohydrolase structures clearly reveals the conservation of this architecture (Fig. 3).

GH3 NagZ enzymes depart significantly from the above two-domain active site architecture. A crystal structure of NagZ from B. subtilis recently revealed that the catalytic acid-base is an unusual histidine/aspartate dyad that resides within a flexible loop that is present on the catalytic (b/a)8 barrel and not on a separate domain [19]. Though the enzyme adopts a two-domain fold similar to barely β-D-glucan glucohydrolase, the C-terminal domain does not participate in catalysis (Fig. 4). In fact, most NagZ enzymes from Gram-negative bacteria are single domain enzymes comprised solely of a catalytic (b/a)8 barrel that contains the conserved aspartate nucleophile and a catalytic histidine/aspartate dyad on flexible loop as seen for B. subtilis NagZ [20] (Fig. 5). The flexible loop containing the His/Asp dyad is unique to GH3 NagZ enzymes and can be identified by the consensus motif [KH(F/I)PG(H/L)GXXXXD(S/T)H] (catalytic dyad highlighted in boldface) [21].


Three-dimensional structures

Figure 1. Ribbon representation of barley β-glucan exohydrolase isoenzyme ExoI. Domain 1, domain 2, and the linker region of the enzyme are coloured in magenta, cyan, and yellow, respectively. Figure from [1].
Figure 2. NagZ from Vibrio cholerae in complex with PUGNAc (PDB ID: 2OXN) [23]. NagZ enzymes are single domain proteins that adopt a TIM barrel fold. Active site residues are located within the loops that extend from the C-termini of the strands of the β-barrel.

The family 3 β-D-glycosidases are globular monomeric enzymes of molecular mass around 60-70 kDa. The 3D structure of the β-D-glucan glucohydrolase isoenzyme ExoI from barley, determined by X-ray crystallography to 2.2 Å resolution, shows a two-domain, globular protein of 605 amino acid residues that is N-glycosylated at three sites [11]. The two domains are connected by a 16-amino acid helix-like linker. The first 357 residues constitute a (β/α)8 TIM barrel domain. The second domain consists of residues 374 to 559 arranged in a six-stranded β-sandwich, which contains a β-sheet of five parallel β-strands and one antiparallel β-strand, with 3 α-helices on either side of the sheet. A long antiparallel loop of 42 amino acid residues is found at the COOH-terminus of the enzyme. In some bacterial GH3 enzymes the order of the domains is reversed [1].

The active site of the barley β-D-glucan glucohydrolase consists of a relatively shallow substrate-binding pocket that is located at the interface of the two domains of the enzyme [11]. The active site pocket can accommodate the two glucosyl residues at the non-reducing terminus of the substrate and aligns the non-reducing terminal glycosidic linkage of the substrate with the catalytic amino acid residues Asp285 and Glu491. Thus, the catalytic amino acid residues are located on domains 1 and 2, respectively.

The broad specificity of the barley β-D-glucan glucohydrolase can be rationalized from the X-ray crystallographic data and from molecular modelling of enzyme-substrate complexes [11, 27]. The glucosyl residue occupying binding subsite –1 is tightly locked into a relatively fixed position through interactions with six amino acid residues near the bottom of the shallow active site pocket. In contrast, the glucosyl residue at subsite +1 is located between two tryptophan residues at the entrance of the pocket, where it is less tightly constrained. The flexibility of binding at subsite +1, coupled with the projection of the remainder of bound substrate away from the enzyme’s surface, means that the overall active site is largely independent of substrate conformation and will therefore accommodate a range of substrates in which the spatial dispositions of adjacent β-D-glucosyl residues vary as a result of glycosidic linkages between different C atoms of the adjacent β-D-glucosyl residues [27].

Family 3 NagZ enzymes are also globular, yet have a mass of ~ 36 kDa, which is a distinctive feature of NagZ enzymes from others within the family. NagZ from Vibrio cholerae has been determined in complex with GlcNAc (PDB ID: 1Y65) and with the N-acetyl-β-glucosaminidase inhibitor PUGNAc [23] and NagZ selective PUGNAc derivatives [28]. The enzyme is comprised of 340 amino acids and adopts a (β/α)8 TIM barrel fold. The active site pocket is shallow and accommodates the 2-N-acetyl group of a terminal GlcNAc sugar in a solvent accessible groove alongside the binding site for the pyranose ring. This active site architecture is different from the active site architectures the functionally-related GH20 N-acetyl-β-hexosaminidases and GH84 O-GlcNAcases. These latter families, which also remove β-1,4-linked GlcNAc residues from glycoconjugates, use a mechanism involving neighboring group participation wherein the carbonyl oxygen of the 2-acetamido group of the terminal GlcNAc acts as a nucleophile, yielding an oxazolinium ion intermediate [29, 30, 31]. Thus, unlike family 3 NagZ enzymes, GH20 and GH84 do not possess an enzymic catalytic nucleophile; however, they do have an appropriately positioned catalytic acid residue. Together, these mechanistic differences have allowed for the development of 2-N-acyl derivatives of PUGNAc that are selective for family 3 NagZ over GH20 and GH84 enzymes [23, 28].

Family Firsts

First 3D Structure
Barley β-D-glucan glucohydrolase [11].
First Catalytic Residues
Barley β-D-glucan glucohydrolase [11].

References

  1. Harvey AJ, Hrmova M, De Gori R, Varghese JN, and Fincher GB. (2000). Comparative modeling of the three-dimensional structures of family 3 glycoside hydrolases. Proteins. 2000;41(2):257-69. DOI:10.1002/1097-0134(20001101)41:2<257::aid-prot100>3.0.co;2-c | PubMed ID:10966578 [Harvey2000]
  2. Lee RC, Hrmova M, Burton RA, Lahnstein J, and Fincher GB. (2003). Bifunctional family 3 glycoside hydrolases from barley with alpha -L-arabinofuranosidase and beta -D-xylosidase activity. Characterization, primary structures, and COOH-terminal processing. J Biol Chem. 2003;278(7):5377-87. DOI:10.1074/jbc.M210627200 | PubMed ID:12464603 [Lee2003]
  3. Mayer C, Vocadlo DJ, Mah M, Rupitz K, Stoll D, Warren RA, and Withers SG. (2006). Characterization of a beta-N-acetylhexosaminidase and a beta-N-acetylglucosaminidase/beta-glucosidase from Cellulomonas fimi. FEBS J. 2006;273(13):2929-41. DOI:10.1111/j.1742-4658.2006.05308.x | PubMed ID:16762038 [Mayer2006]
  4. Hrmova, M. and Fincher, G.B. (1998) Barley ß-D-glucan exohydrolases. Substrate specificity and kinetic properties. Carbohydr. Res. 305, 209-221.

    [Hrmova1998]
  5. Chitlaru E and Roseman S. (1996). Molecular cloning and characterization of a novel beta-N-acetyl-D-glucosaminidase from Vibrio furnissii. J Biol Chem. 1996;271(52):33433-9. DOI:10.1074/jbc.271.52.33433 | PubMed ID:8969206 [Chitlaru1996]
  6. Johnson JW, Fisher JF, and Mobashery S. (2013). Bacterial cell-wall recycling. Ann N Y Acad Sci. 2013;1277(1):54-75. DOI:10.1111/j.1749-6632.2012.06813.x | PubMed ID:23163477 [Johnson2013]
  7. Cheng Q, Li H, Merdek K, and Park JT. (2000). Molecular characterization of the beta-N-acetylglucosaminidase of Escherichia coli and its role in cell wall recycling. J Bacteriol. 2000;182(17):4836-40. DOI:10.1128/JB.182.17.4836-4840.2000 | PubMed ID:10940025 [Cheng2000]
  8. Litzinger S, Fischer S, Polzer P, Diederichs K, Welte W, and Mayer C. (2010). Structural and kinetic analysis of Bacillus subtilis N-acetylglucosaminidase reveals a unique Asp-His dyad mechanism. J Biol Chem. 2010;285(46):35675-84. DOI:10.1074/jbc.M110.131037 | PubMed ID:20826810 [Litzinger2010a]
  9. Mark BL, Vocadlo DJ, and Oliver A. (2011). Providing β-lactams a helping hand: targeting the AmpC β-lactamase induction pathway. Future Microbiol. 2011;6(12):1415-27. DOI:10.2217/fmb.11.128 | PubMed ID:22122439 [Mark2011]
  10. Davies GJ, Wilson KS, and Henrissat B. (1997). Nomenclature for sugar-binding subsites in glycosyl hydrolases. Biochem J. 1997;321 ( Pt 2)(Pt 2):557-9. DOI:10.1042/bj3210557 | PubMed ID:9020895 [Davies1997]
  11. Varghese JN, Hrmova M, and Fincher GB. (1999). Three-dimensional structure of a barley beta-D-glucan exohydrolase, a family 3 glycosyl hydrolase. Structure. 1999;7(2):179-90. DOI:10.1016/s0969-2126(99)80024-0 | PubMed ID:10368285 [Vargese1999]
  12. Bause E and Legler G. (1974). Isolation and amino acid sequence of a hexadecapeptide from the active site of beta-glucosidase A3 from Aspergillus wentii. Hoppe Seylers Z Physiol Chem. 1974;355(4):438-42. DOI:10.1515/bchm2.1974.355.1.438 | PubMed ID:4611895 [Bause1974]
  13. Dan S, Marton I, Dekel M, Bravdo BA, He S, Withers SG, and Shoseyov O. (2000). Cloning, expression, characterization, and nucleophile identification of family 3, Aspergillus niger beta-glucosidase. J Biol Chem. 2000;275(7):4973-80. DOI:10.1074/jbc.275.7.4973 | PubMed ID:10671536 [Dan2000]
  14. Chir J, Withers S, Wan CF, and Li YK. (2002). Identification of the two essential groups in the family 3 beta-glucosidase from Flavobacterium meningosepticum by labelling and tandem mass spectrometric analysis. Biochem J. 2002;365(Pt 3):857-63. DOI:10.1042/BJ20020186 | PubMed ID:11978178 [Chir2002]
  15. Li YK, Chir J, Tanaka S, and Chen FY. (2002). Identification of the general acid/base catalyst of a family 3 beta-glucosidase from Flavobacterium meningosepticum. Biochemistry. 2002;41(8):2751-9. DOI:10.1021/bi016049e | PubMed ID:11851422 [Li2002]
  16. Pozzo T, Pasten JL, Karlsson EN, and Logan DT. (2010). Structural and functional analyses of beta-glucosidase 3B from Thermotoga neapolitana: a thermostable three-domain representative of glycoside hydrolase 3. J Mol Biol. 2010;397(3):724-39. DOI:10.1016/j.jmb.2010.01.072 | PubMed ID:20138890 [Pozzo2010]
  17. Litzinger S, Duckworth A, Nitzsche K, Risinger C, Wittmann V, and Mayer C. (2010). Muropeptide rescue in Bacillus subtilis involves sequential hydrolysis by beta-N-acetylglucosaminidase and N-acetylmuramyl-L-alanine amidase. J Bacteriol. 2010;192(12):3132-43. DOI:10.1128/JB.01256-09 | PubMed ID:20400549 [Litzinger2010b]
  18. Bacik JP, Whitworth GE, Stubbs KA, Vocadlo DJ, and Mark BL. (2012). Active site plasticity within the glycoside hydrolase NagZ underlies a dynamic mechanism of substrate distortion. Chem Biol. 2012;19(11):1471-82. DOI:10.1016/j.chembiol.2012.09.016 | PubMed ID:23177201 [Bacik2012]
  19. Vocadlo DJ, Mayer C, He S, and Withers SG. (2000). Mechanism of action and identification of Asp242 as the catalytic nucleophile of Vibrio furnisii N-acetyl-beta-D-glucosaminidase using 2-acetamido-2-deoxy-5-fluoro-alpha-L-idopyranosyl fluoride. Biochemistry. 2000;39(1):117-26. DOI:10.1021/bi991958d | PubMed ID:10625486 [Vocadlo2000]
  20. Vocadlo DJ and Withers SG. (2005). Detailed comparative analysis of the catalytic mechanisms of beta-N-acetylglucosaminidases from families 3 and 20 of glycoside hydrolases. Biochemistry. 2005;44(38):12809-18. DOI:10.1021/bi051121k | PubMed ID:16171396 [Vocadlo2005]
  21. Stubbs KA, Balcewich M, Mark BL, and Vocadlo DJ. (2007). Small molecule inhibitors of a glycoside hydrolase attenuate inducible AmpC-mediated beta-lactam resistance. J Biol Chem. 2007;282(29):21382-91. DOI:10.1074/jbc.M700084200 | PubMed ID:17439950 [Stubbs2007]
  22. Hrmova M, Varghese JN, De Gori R, Smith BJ, Driguez H, and Fincher GB. (2001). Catalytic mechanisms and reaction intermediates along the hydrolytic pathway of a plant beta-D-glucan glucohydrolase. Structure. 2001;9(11):1005-16. DOI:10.1016/s0969-2126(01)00673-6 | PubMed ID:11709165 [Hrmova2001]
  23. Nakatani Y, Cutfield SM, Cowieson NP, and Cutfield JF. (2012). Structure and activity of exo-1,3/1,4-β-glucanase from marine bacterium Pseudoalteromonas sp. BB1 showing a novel C-terminal domain. FEBS J. 2012;279(3):464-78. DOI:10.1111/j.1742-4658.2011.08439.x | PubMed ID:22129429 [Nakatani2012]
  24. Yoshida E, Hidaka M, Fushinobu S, Koyanagi T, Minami H, Tamaki H, Kitaoka M, Katayama T, and Kumagai H. (2010). Role of a PA14 domain in determining substrate specificity of a glycoside hydrolase family 3 β-glucosidase from Kluyveromyces marxianus. Biochem J. 2010;431(1):39-49. DOI:10.1042/BJ20100351 | PubMed ID:20662765 [Yoshida2010]
  25. Hrmova M, De Gori R, Smith BJ, Fairweather JK, Driguez H, Varghese JN, and Fincher GB. (2002). Structural basis for broad substrate specificity in higher plant beta-D-glucan glucohydrolases. Plant Cell. 2002;14(5):1033-52. DOI:10.1105/tpc.010442 | PubMed ID:12034895 [Hrmova2002]
  26. Balcewich MD, Stubbs KA, He Y, James TW, Davies GJ, Vocadlo DJ, and Mark BL. (2009). Insight into a strategy for attenuating AmpC-mediated beta-lactam resistance: structural basis for selective inhibition of the glycoside hydrolase NagZ. Protein Sci. 2009;18(7):1541-51. DOI:10.1002/pro.137 | PubMed ID:19499593 [Balcewich2009]
  27. Tews I, Perrakis A, Oppenheim A, Dauter Z, Wilson KS, and Vorgias CE. (1996). Bacterial chitobiase structure provides insight into catalytic mechanism and the basis of Tay-Sachs disease. Nat Struct Biol. 1996;3(7):638-48. DOI:10.1038/nsb0796-638 | PubMed ID:8673609 [Tews1996]
  28. Mark BL, Vocadlo DJ, Knapp S, Triggs-Raine BL, Withers SG, and James MN. (2001). Crystallographic evidence for substrate-assisted catalysis in a bacterial beta-hexosaminidase. J Biol Chem. 2001;276(13):10330-7. DOI:10.1074/jbc.M011067200 | PubMed ID:11124970 [Mark2001]
  29. Dennis RJ, Taylor EJ, Macauley MS, Stubbs KA, Turkenburg JP, Hart SJ, Black GN, Vocadlo DJ, and Davies GJ. (2006). Structure and mechanism of a bacterial beta-glucosaminidase having O-GlcNAcase activity. Nat Struct Mol Biol. 2006;13(4):365-71. DOI:10.1038/nsmb1079 | PubMed ID:16565725 [Dennis2006]
  30. Vötsch W and Templin MF. (2000). Characterization of a beta -N-acetylglucosaminidase of Escherichia coli and elucidation of its role in muropeptide recycling and beta -lactamase induction. J Biol Chem. 2000;275(50):39032-8. DOI:10.1074/jbc.M004797200 | PubMed ID:10978324 [Votsch2000]
  31. Asgarali A, Stubbs KA, Oliver A, Vocadlo DJ, and Mark BL. (2009). Inactivation of the glycoside hydrolase NagZ attenuates antipseudomonal beta-lactam resistance in Pseudomonas aeruginosa. Antimicrob Agents Chemother. 2009;53(6):2274-82. DOI:10.1128/AAC.01617-08 | PubMed ID:19273679 [Asgarali2009]
  32. Thongpoo P, McKee LS, Araújo AC, Kongsaeree PT, and Brumer H. (2013). Identification of the acid/base catalyst of a glycoside hydrolase family 3 (GH3) beta-glucosidase from Aspergillus niger ASKU28. Biochim Biophys Acta. 2013;1830(3):2739-49. DOI:10.1016/j.bbagen.2012.11.014 | PubMed ID:23201198 [Thongpoo2012]
  33. Li H, Zhao G, Miyake H, Umekawa H, Kimura T, Ohmiya K, and Sakka K. (2006). Identification of a catalytic residue of Clostridium paraputrificum N-acetyl-beta-D-glucosaminidase Nag3A by site-directed mutagenesis. Biosci Biotechnol Biochem. 2006;70(5):1127-33. DOI:10.1271/bbb.70.1127 | PubMed ID:16717412 [Li2006]
  34. Paal K, Ito M, and Withers SG. (2004). Paenibacillus sp. TS12 glucosylceramidase: kinetic studies of a novel sub-family of family 3 glycosidases and identification of the catalytic residues. Biochem J. 2004;378(Pt 1):141-9. DOI:10.1042/BJ20031028 | PubMed ID:14561218 [Pall2004]
  35. Zmudka MW, Thoden JB, and Holden HM. (2013). The structure of DesR from Streptomyces venezuelae, a β-glucosidase involved in macrolide activation. Protein Sci. 2013;22(7):883-92. DOI:10.1002/pro.2204 | PubMed ID:23225731 [Zmudka2013]
  36. Yamaguchi T, Blázquez B, Hesek D, Lee M, Llarrull LI, Boggess B, Oliver AG, Fisher JF, and Mobashery S. (2012). Inhibitors for Bacterial Cell-Wall Recycling. ACS Med Chem Lett. 2012;3(3):238-242. DOI:10.1021/ml2002746 | PubMed ID:22844551 [Yamaguchi2012]

All Medline abstracts: PubMed