15
Article Nanomolar inhibition of human OGA by 2- acetamido-2-deoxy-D-glucono-1,5-lactone semicarbazone derivatives Kiss, Mariann, Szabó, Erna, Bocska, Boglárka, Sinh, Luu Thanh, Fernandes, Conceicao Piedade, Timári, István, Hayes, Joseph, Somsák, László and Barna, Teréz Available at http://clok.uclan.ac.uk/38274/ Kiss, Mariann, Szabó, Erna, Bocska, Boglárka, Sinh, Luu Thanh, Fernandes, Conceicao Piedade, Timári, István, Hayes, Joseph ORCID: 0000-0002-7745- 9616, Somsák, László and Barna, Teréz (2021) Nanomolar inhibition of human OGA by 2-acetamido-2-deoxy-D-glucono-1,5-lactone semicarbazone derivatives. European Journal of Medicinal Chemistry, 223 . p. 113649. ISSN 0223-5234 It is advisable to refer to the publisher’s version if you intend to cite from the work. http://dx.doi.org/10.1016/j.ejmech.2021.113649 For more information about UCLan’s research in this area go to http://www.uclan.ac.uk/researchgroups/ and search for <name of research Group>. For information about Research generally at UCLan please go to http://www.uclan.ac.uk/research/ All outputs in CLoK are protected by Intellectual Property Rights law, including Copyright law. Copyright, IPR and Moral Rights for the works on this site are retained by the individual authors and/or other copyright owners. Terms and conditions for use of this material are defined in the policies page. CLoK Central Lancashire online Knowledge www.clok.uclan.ac.uk

Nanomolar inhibition of human OGA by 2-acetamido-2-deoxy …

  • Upload
    others

  • View
    4

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Nanomolar inhibition of human OGA by 2-acetamido-2-deoxy …

Article

Nanomolar inhibition of human OGA by 2-acetamido-2-deoxy-D-glucono-1,5-lactonesemicarbazone derivatives

Kiss, Mariann, Szabó, Erna, Bocska, Boglárka, Sinh, Luu Thanh, Fernandes, Conceicao Piedade, Timári, István, Hayes, Joseph, Somsák, László and Barna, Teréz

Available at http://clok.uclan.ac.uk/38274/

Kiss, Mariann, Szabó, Erna, Bocska, Boglárka, Sinh, Luu Thanh, Fernandes, Conceicao Piedade, Timári, István, Hayes, Joseph ORCID: 0000-0002-7745-9616, Somsák, László and Barna, Teréz (2021) Nanomolar inhibition of human OGA by 2-acetamido-2-deoxy-D-glucono-1,5-lactone semicarbazone derivatives. European Journal of Medicinal Chemistry, 223 . p. 113649. ISSN 0223-5234

It is advisable to refer to the publisher’s version if you intend to cite from the work.http://dx.doi.org/10.1016/j.ejmech.2021.113649

For more information about UCLan’s research in this area go to http://www.uclan.ac.uk/researchgroups/ and search for <name of research Group>.

For information about Research generally at UCLan please go to http://www.uclan.ac.uk/research/

All outputs in CLoK are protected by Intellectual Property Rights law, includingCopyright law. Copyright, IPR and Moral Rights for the works on this site are retainedby the individual authors and/or other copyright owners. Terms and conditions for useof this material are defined in the policies page.

CLoKCentral Lancashire online Knowledgewww.clok.uclan.ac.uk

Page 2: Nanomolar inhibition of human OGA by 2-acetamido-2-deoxy …

lable at ScienceDirect

European Journal of Medicinal Chemistry 223 (2021) 113649

Contents lists avai

European Journal of Medicinal Chemistry

journal homepage: http: / /www.elsevier .com/locate /ejmech

Nanomolar inhibition of human OGA by 2-acetamido-2-deoxy-D-glucono-1,5-lactone semicarbazone derivatives

Mariann Kiss a, Erna Szab�o b, Bogl�arka Bocska a, Luu Thanh Sinh b,Conceicao Piedade Fernandes c, Istv�an Tim�ari a, Joseph M. Hayes c, **, L�aszl�o Soms�ak a, *,Ter�ez Barna b, ***

a Department of Organic Chemistry, University of Debrecen, H-4002, POB 400, Debrecen, Hungaryb Department of Genetics and Applied Microbiology, University of Debrecen, H-4002, POB 400, Debrecen, Hungaryc School of Pharmacy & Biomedical Sciences, University of Central Lancashire, Preston, PR1 2HE, United Kingdom

a r t i c l e i n f o

Article history:Received 12 March 2021Received in revised form11 June 2021Accepted 12 June 2021Available online 15 June 2021

Keywords:hOGAExpressionInhibitorGlyconolactone semicarbazoneQM/MM-PBSAHalogenHalogen�Hydrogen bond donor

* Corresponding author.** Corresponding author.*** Corresponding author.

E-mail addresses: [email protected] (J.M. Haunideb.hu (L. Soms�ak), [email protected]

https://doi.org/10.1016/j.ejmech.2021.1136490223-5234/© 2021 The Authors. Published by Elseviernd/4.0/).

a b s t r a c t

O-GlcNAcylation is a dynamic post-translational modification mediated by O-linked b-N-acetylglucos-amine transferase (OGT) and O-GlcNAc hydrolase (OGA), that adds or removes a single b-N-acetylglu-cosamine (GlcNAc) moiety to or from serine/threonine residues of nucleocytosolic and mitochondrialproteins, respectively. The perturbed homeostasis of O-GlcNAc cycling results in several pathologicalconditions. Human OGA is a promising therapeutic target in diseases where aberrantly low levels of O-GlcNAc are experienced, such as tauopathy in Alzheimer's disease. A new class of potent OGA inhibitors,2-acetamido-2-deoxy-D-glucono-1,5-lactone (thio)semicarbazones, have been identified. Eight inhibitorswere designed and synthesized in five steps starting from D-glucosamine and with 15e55% overall yields.A heterologous OGA expression protocol with strain selection and isolation has been optimized thatresulted in stable, active and full length human OGA (hOGA) isomorph. Thermal denaturation kinetics ofhOGA revealed environmental factors affecting hOGA stability. From kinetics experiments, the synthe-sized compounds proved to be efficient competitive inhibitors of hOGA with Ki-s in the range of ~30e250 nM and moderate selectivity with respect to lysosomal b-hexosaminidases. In silico studies con-sisting of Prime protein-ligand refinements, QM/MM optimizations and QM/MM-PBSA binding freeenergy calculations revealed the factors governing the observed potencies, and led to design of the mostpotent analogue 2-acetamido-2-deoxy-D-glucono-1,5-lactone 4-(2-naphthyl)-semicarbazone 6g(Ki ¼ 36 nM). The protocol employed has applications in future structure based inhibitor design targetingOGA.© 2021 The Authors. Published by Elsevier Masson SAS. This is an open access article under the CC BY-

NC-ND license (http://creativecommons.org/licenses/by-nc-nd/4.0/).

1. Introduction

O-GlcNAcylation has emerged as a major posttranslationalmodification (PTM) that indicates the energy, nutritional and stressstatus of the cell, and involves transfer of a single N-acetylglucos-amine moiety from UDP-GlcNAc to a serine or threonine side chainof proteins [1]. The two most striking features of O-GlcNAcylation,which distinguish this PTM from N- and O-glycosylations are its

yes), somsak.laszlo@science.(T. Barna).

Masson SAS. This is an open access

localisation and its dynamic nature. It is localized to cytosol, nu-cleus and mitochondria rather than the endoplasmic reticulum orGolgi apparatus [2]. Similar to phosphorylation, two opposingenzyme actions determine the level of O-GlcNAcylation of a pro-tein: OGT (O-GlcNAc transferase), the transferase, which transfersthe N-acetylglucosamine moiety and creates the b-glycosidic link-age between the protein and GlcNAc, and the hydrolase OGA (O-GlcNAcase), which cleaves the bond [3]. UDP-GlcNAc, produced bythe hexosamine pathway, provides the GlcNAc moiety for OGT totransfer to the aliphatic hydroxyl groups of the proteins to bemodified. This UDP-GlcNAc is the nutrient sensor of the cell and O-GlcNAcylation, in response to the hexosamine flux, modulatesmetabolism and affects numerous cellular processes such as tran-scription regulatory pathways, cell cycle progression, embryonic

article under the CC BY-NC-ND license (http://creativecommons.org/licenses/by-nc-

Page 3: Nanomolar inhibition of human OGA by 2-acetamido-2-deoxy …

M. Kiss, E. Szab�o, B. Bocska et al. European Journal of Medicinal Chemistry 223 (2021) 113649

development, the regulation of the epigenome, respiration andproteosomal degradation [4e8]. O-GlcNAcylation does not onlyswitch the biological function of key regulatory proteins on and off,but finely tunes their activities and interaction partners in theintracellular network. Imbalance in the GlcNAcylation cycle hasbeen shown to manifest itself in diseases, where the hyperactivityof either OGA or OGT is observed. In nearly all cancers, an elevatedlevel of O-GlcNAcylation was reported [9]. The trend is opposite inthe brain, in multiple neurodegenerative diseases where decreasedlevels of O-GlcNAc modifications on numerous proteins have beenreported, such as the proteins of the cytoskeleton [10]. The shift ofhomeostasis of O-GlcNAcylation to the increased hydrolysis ofGlcNAc from the modified proteins is one of the main mechanismscontributing to neurodegeneration [11]. As a consequence of thecomplementary and competitive nature of phosphorylation and O-GlcNAcylation at many serine/threonine sites, low levels of O-GlcNAcylation leads to hyperphosphorylation of cytoskeleton pro-teins in the brain, including the microtubule protein tau. Thehyperphosphorylated form of tau has been found in neurofibrillarytangles, characterized as tauopathy. An increase of O-GlcNAcylationlevels by inhibiting the activity of OGA has significantly decreasedthe aggregation propensity of tau [12,13]. The considerable thera-peutic relevance of dissecting the mechanism of O-GlcNAcylation,motivates biological studies, and design and synthesis of potentOGA inhibitors.

Human OGA (hOGA) is a neutral hexosaminidase and belongs tofamily 84 of glycoside hydrolases (GH84) [14]. The enzyme followsa two-step substrate-assisted mechanism, through a dissociativeoxocarbenium ion-like transition state, resulting in the net reten-tion of the anomeric configuration [15]. This mechanism is sharedby the acidic N-acetylhexosaminidases, a member of GH20 [16]. Inthe GH84 family two neighbouring aspartates and in the GH20enzymes, one aspartate and one glutamate, function as the keycatalytic nucleophile and the acid/base residues [16e18]. There arespecific human N-acetylhexosaminidases that reside in lysosomesand are expressed in several tissues, mostly as a mixture of twoisoenzymes: N-acetylhexosaminidase A (HexA) and N-acetylhex-osaminidase B (HexB) [19]. GH20 N-acetylhexosaminidases catal-yse the hydrolysis of terminal, non-reducing N-acetyl-b-D-glucosamine and N-acetyl-b-D-galactosamine residues from pro-teoglycans [16,20]. Both isoforms appear in plasma and in synovialfluid and become elevated in inflammatory processes. In the sy-novial fluid of patients with osteoarthritis, HexA and HexB arereleased by chondrocytes into the extracellular compartment,mainly responsible for the degradation of glycosaminoglycans(GAG) [21,22]. Development of potent inhibitors against HexA andHexB are also highly relevant in studying the physiological role ofthese enzymes and their potential therapeutic applications.

PUGNAc (Phenyl Urethane of GlcNAc, Chart 1) was one of thefirst inhibitors of OGA [23] with a Ki value of around 50 nM, how-ever, it is not selective, since it inhibits GH20 N-acetylhex-osaminidases in a similar manner. The N-phenylcarbamate moiety

Chart 1. The structure of some OGA inhibitors

2

can form hydrogen bonds between the oxime nitrogen atom andcatalytic site residues, and the phenyl moiety, which protrudesfrom the active site pocket, takes advantage of hydrophobic in-teractions with aromatic side chains [24]. 2-Acetamido-2-deoxy-D-gluconhydroximo-1,5-lactone (LOGNAc), which lacks the phenylcarbamate group, is an around 30 times less potent inhibitor [23],showing that this moiety makes an important contribution tobinding. Several other inhibitors are known in the literature, mostlytransition state analogue compounds (NAG-thiazoline, NButGT,Thiamet-G), glucoimidazoles (GlcNacstatins) and non-carbohy-drate-based inhibitors have also been reported [25]. Surprisingly,there are only a few examples of PUGNAc analogues where thearomatic moiety is changed [26]. In addition, to the best of ourknowledge, the effects of modifications of the urethane linker onOGA inhibition have not yet been studied.

Though PUGNAc is awell established OGA inhibitor, its synthesishas some drawbacks. The first eight-step synthetic route [27](overall yield: 14%) included two critical steps: a) the incompleteand low yielding transformation of the starting compound in theoxidation step towards the hydroximo-1,5-lactone (50% with 35%unreacted starting material); b) a debenzylidenation with Na/NH3.The scale-up of both transformations to multigram synthesis led tolower yields and involved inconvenient workup procedures [28]. Inan improved six-step synthesis [28] (overall yield: 15%) thehydroximo-1,5-lactone was obtained in an 59% yield only, and forthis step keeping the �40 �C temperature was crucial. At lowertemperature the conversion was incomplete, while at higher tem-perature the 1,4-lactone oxime was formed.

With these preliminaries in mind we envisaged the synthesis ofaryl substituted (thio)semicarbazones of 2-acetamido-2-deoxy-D-glucono-1,5-lactone (target compounds in Chart 1) whose prepa-ration was expected to lack critical steps such as those for PUGNAc.Kinetics experiments were performed to determine the potenciesfor hOGA inhibition, as well as selectivity studies considering alsoN-acetylhexosaminidases. High level in silico calculations usingquantum mechanics/molecular mechanics (QM/MM) methodswere performed to understand the hOGA binding properties andrelative inhibitory efficiencies of the compounds.

2. Results and discussion

2.1. Synthesis

A single representative of our target compounds (Chart 1, Ar ¼Ph, X¼O) was located in the literature, therefore, the synthesis wasbased on the adaptation of this published procedure [29] (Table 1).The starting material 2-acetamido-3,4,6-tri-O-acetyl-2-deoxy-D-glucopyranose (1) was obtained by a known two-step procedure[30,31] from D-glucosamine hydrochloride. Compound 1 was thentransformed by a 4-arylsemicarbazide (2, either commercialchemicals or obtained from the corresponding aromatic amines asdescribed in the supporting material) in the presence of 10 mol% of

and the target compounds of this study.

Page 4: Nanomolar inhibition of human OGA by 2-acetamido-2-deoxy …

Table 1Synthesis of the test compounds 6.

Reagents and conditions: a) 1. Ac2O, abs. Py, r.t.; 2. 4 Å molecular sieves, abs. MeOH, r.t. (83% for two steps) [30,31]; b) NH2NHR (2), 0.1 equiv. pTsOH$H2O, CHCl3, reflux; c)with a-g: CrO3-Py, CH2Cl2, 0 �C / r.t.; d) with h: MnO2, abs. CH2Cl2, reflux; e) NH3/MeOH, r.t.

R Products and yields (%)

3 þ 4 Ratioa of 3 : 4 5 6 Overall yield of 6 from D-glucosamine

a CONHPh 90 5:1 83 79 49b CONH(p-MePh) 83 7:1 92 86 55c CONH(p-CF3Ph) 81 5:1 26 87 15d CONH(p-OMePh) 87 7:3 86 83 52e CONH(p-ClPh) 77 4:1 78 74 37f CONH(1-naphthyl) 94 b 64 83 41g CONH(2-naphthyl) 80 b 93 75 46h CSNHPh 87c e 78 83 47

a Determined from the 1H NMR spectra of the worked-up reaction mixtures.b Formation of the open-chain products 4f and 4gwas detected by TLC but due to signal overlaps and small amounts of these compounds the 3: 4 ratio was not established.c Compound 3 formed as a single product.

M. Kiss, E. Szab�o, B. Bocska et al. European Journal of Medicinal Chemistry 223 (2021) 113649

p-toluenesulfonic acid as catalyst to get the N1-glycosyl semi-carbazides 3. In place of the originally proposed 1,2-dichloroethanethe use of dichloromethane, chloroform and toluene was tested forthis transformation. In toluene the solubility of the starting mate-rial was insufficient, while the reaction in boiling CH2Cl2 resulted inlower yields than in CHCl3 at reflux temperature, therefore, thelatter solvent was used in the reactions. With the semicarbazidederivatives 2a-g the equilibrating open-chain tautomers 4a-g werealso observed in the 1H NMR spectra, while the reaction with thi-osemicarbazide 2h was devoid of this product. No attempt wasmade to separate compounds 3 and 4. The NMR spectra of thesemixtures contained signals which were analogous to those ob-tained for pure 3a and 4a in the literature [29]. The b(D) configu-ration of compounds 3 followed from the large 3JH-1 e H-2 couplingsof 9e10 Hz (c.f. Experimental section). Oxidation of the mixturescontaining 3 and 4with CrO3-pyridine complex gave good yields ofthe expected lactone semicarbazones 5a-g. Regrettably, theseoxidation conditions yielded complex reaction mixtures in case ofthe thio derivative 3h. Several oxidizing agents were tested (I2/NEt3, PIDA, ICl/Cs2CO3) and finally MnO2 proved to be efficient tofurnish the required compound 5h. The lactone semicarbazones 5were O-deprotected by a solution of NH3 in MeOH to result in thetest compounds 6. The overall yields of these five-step syntheticsequences to give target compounds 6 from D-glucosamine fell inthe 15e55% range.

To verify the configuration of the C]N bond in compounds 6, aNuclear Overhauser Effect (NOE) based NMR approach was appliedto 6c (Fig. 1). The E-Z isomeric forms of compounds 6 are expectedto be distinguished on the basis of NOE cross-peaks between theN(2)H and distinctive protons of the sugar part, since themaximumspatial distance which can provide an NOE signal is about 5 Å. First,

3

a 2D 1He13C HMBC spectrum [32] (Fig. S1) was recorded for theunambiguous assignment of N(2)H (9.8 ppm). Then a 2D 1He1HEASY ROESY experiment [33] was performed, which is a powerfulalternative of NOE measurements for small molecules. In the1He1H ROESY spectrum (Fig. S2), the N(2)H resonance providedNOE cross-peaks with the H(50) and O(6’)H resonances (Fig. 1),which indicated the vicinity in space of these protons, and so, thepresence of the Z isomeric form of compound 6c. Since both the 1Hand 13C NMR spectra for the other compounds 6 closely resemblethose of 6c, one can safely conclude the presence of the Z isomeralso in those compounds.

2.2. Heterologous expression of hOGA

A heterologous expression system for hOGA in E. coli was

Fig. 1. Nuclear Overhauser Effects (NOEs) observed in compound 6c.

Page 5: Nanomolar inhibition of human OGA by 2-acetamido-2-deoxy …

M. Kiss, E. Szab�o, B. Bocska et al. European Journal of Medicinal Chemistry 223 (2021) 113649

established in order to gain soluble, active, full-length wild-typeenzyme with considerable thermostability for the kinetic evalua-tion of the inhibitory potency of new inhibitors 6a-h against hOGA.Bacteriophage T7 polymerase promoter based expression systemsuch as BL21(DE3)/pET system was chosen for this purpose. ThepET-28 based vector construct carrying the N-terminally His-tagged hOGA encoded gene was transformed into five E. coliBL21(DE3) strains and the cultivation conditions were optimized, aswell as the environmental parameters of the downstream pro-cesses. The small scale expression trials of all the transformedBL21(DE3) strains carried out at 23 �C unravelled two main ob-jectives to be solved in this expression system. Namely, the hightendency of the recombinant hOGA to form inclusion body aggre-gates during the protein production phase and the significantproteolytic degradation of the recombinant enzyme suffered in vivoand in vitro (Fig. 2). We did experience the latter phenomenon,despite all the strains usedwere deficient in two proteases, an outermembrane protease (ompT) and an ATP dependent protease (Lon),both of which have been reported as responsible for degradation ofmisfolded proteins and interference with production of some full-length recombinant proteins [34e36]. In this respect, the worstperforming strain was BL21(DE3), where full length recombinantenzyme was hardly produced (Fig. 2). Beside the non-specificproteolysis, significant production of a 70 kDa long truncatedhOGA fragment was observed with BL21 Star and pLysS strainsreferring to an existence of a sequence specific proteolytic cleavagesite. This fragment bound to the affinity column and exhibited b-N-acetylglucosaminidase activity (see supporting material, Fig. S3).The size and activity of this comparatively stable hOGA fragmentsuggests the presence of caspase-3-like protease activity of E. colicells that might be activated upon the overproduction of hOGA,causing a proteotoxic stress for the cell [37e40]. Two strains, RIPLand Rosetta, where the translation was optimized for genes thatdiffer in codon usage from that of E. coli by carrying rare tRNA geneson plasmids, showed increased production of recombinant enzyme.However, only the Rosetta strain could increase the synthesis of theenzyme in the active, soluble form (Fig. 2). The productivity of theStar and pLysS strains was very similar with a relatively high yieldof active hOGA compared to the aggregate form. This highlights theimportance of the increased life-time of mRNA in Star and the verytight control of expression in pLysS (Fig. 2). However, Star and pLysSexhibited the highest site-specific proteolytic activity producing

Fig. 2. Gel electrophoretic (8% SDS-PAGE) analysis of the soluble and insoluble frac-tions of the cell lysates produced by different E. coli BL21(DE3) strains grown at 23 �C,induced at mid-exponential phase with 0.4 mM IPTG and harvested at early stationaryphase.Lanes: 1. protein marker; 2. E. coli BL21(DE3) insoluble fraction; 3. E. coli BL21(DE3)soluble fraction; 4. E. coli BL21(DE3) RIPL insoluble fraction; 5. E. coli BL21(DE3) RIPLsoluble fraction; 6. E. coli BL21(DE3) Rosetta insoluble fraction; 7. E. coli BL21(DE3)Rosetta soluble fraction; 8. E. coli BL21 Star (DE3) insoluble fraction; 9. E. coli BL21 Star(DE3) soluble fraction; 10. E. coli BL21(DE3) pLysS insoluble fraction; 11. E. coliBL21(DE3) pLysS soluble fraction.

4

the 70 kDa long hOGA fragment. Table S1 summarizes the b-N-acetylglucosaminidase activity of the cell lysates of the selectedE. coli strain. It can be misleading to rely only on these results andrank the productivity of the strains according to these activityvalues. For instance, the BL21 (DE3) strain, which yielded the sec-ond highest b-N-acetylglucosaminidase activity, produced thelowest level of soluble, full-length recombinant enzyme. From allthese observations it can be concluded that even though the strainshave the same origin, they displayed different tolerance to theoverproduction of misfolded protein, due to the activation of anextensive network of proteolytic housekeeping and regulating en-zymes [41,42].

Since the Rosetta strain proved to be the best performingexpression host, systematic cultivation parameters were optimizedfor this strain. The effect of temperature on the recombinant pro-tein production was investigated below 30 �C, at three tempera-tures (16, 23 and 28 �C). The low protein concentrations in theoryshould decrease the partial folding andmisfolding of proteins, thus,preventing the aggregation to inclusion body [43]. Decreasing thecultivation temperature only favoured full-length and activeenzyme production at 23 �C. At lower temperature (16 �C), themisfolding of the human enzyme was accelerated and the chap-erone systems of E. coli were not able to prevent the aggregateformation (see supplementary material, Fig. S4, Table S2).

The induction phase of the culture growth also influenced theyield of hOGA production. When the inducer was employed at laterphase than the mid-exponential one (at OD600 0.9 instead of 0.6),we could achieve 50% higher yield (expressed in total b-N-acetyl-glucosaminidase activity: 35 units for the mid-exponential phaseinduction and 53 units for the later phase induction in 250 ml cellculture) without compromising greater proteolytic degradation orformation of inclusion-body aggregates (Fig. S5).

All the optimized conditions were built in the final heterologousexpression protocol for hOGA, in which aggregate formation of theoverproduced enzyme was minimized and an efficient full-length,active hOGA production was ensured.

2.3. Stability studies on wild-type hOGA

Themost important issue during the downstream processes wasundoubtedly minimization of the action of E. coli proteases.Nevertheless, the identification of stabilizing and destabilizingfactors involved in maintaining hOGA in its native state was alsoessential. Therefore, we carried out stability studies on hOGA byvarying buffer compositions and we followed the heat-induceddenaturation kinetics of the enzyme at 50 �C. The time-course ofthe enzyme residual activities was determined and the thermaldenaturation kinetic constants (kD) or the half-lifes (t1/2) werecalculated. By increasing the ionic strength, the thermal denatur-ation rate of the enzyme increased too, as was detected in phos-phate buffer upon increasing KCl concentration (Fig. S6a andTable S3). There was no significant difference in the rate of thermaldenaturation of hOGA in the presence of either potassium or so-dium cation (Fig. S6b and Table S3). Treating hOGA solution withreducing agents such as mercaptoethanol and DTT proved to be animportant protective factor increasing the protein half-life by up to25% (Table S3 and Fig. S7b). An even greater stabilizing effect wasobserved when phosphate buffer was supplemented with 15%glycerol to increase the enzyme half-life by 80% (Table S3). On thecontrary, detergent Triton X-100 greatly destabilized hOGA nativestate and accelerated its thermal denaturation rate from kD ¼ 0.05to kD ¼ 0.17 min�1, whereas the presence of Tween 20 proved to bean innocent factor in the structural disintegration of the enzyme(Fig. S7a and Table S3). The effect of the majority of environmentalfactors on thermal denaturation kinetics of hOGA could be

Page 6: Nanomolar inhibition of human OGA by 2-acetamido-2-deoxy …

M. Kiss, E. Szab�o, B. Bocska et al. European Journal of Medicinal Chemistry 223 (2021) 113649

modelled by a single-transient process between the native andunfolded states that obeys a first-order kinetics. When phosphatebuffer alone was present, the heat denaturation of hOGA followedsigmoidal kinetics, which applies for an autocatalytic process(indicating the autocatalytic nature of the unfolding event). We didnot investigate the mechanistic details further, however suggestedexplanation for the existence of cooperativity in domain confor-mational changes is discussed in the supplementary material.

We applied the stabilizing supplementations including thepresence of reducing agent, Tween 20 and the preferable envi-ronmental factors such as low ionic strength, during the celldisruption process. It is also worth emphasizing the strict use ofbroad spectrum protease inhibitors during the cell lysis. During theaffinity chromatography step, high salt concentration was neces-sary to avoid unspecific protein binding. The collected fractions ofhigh b-N-acetylglucosaminidase activity were immediately ultra-filtrated to change the elution buffer into the ‘long storage buffer’containing all the evaluated stabilizing supplementations includingglycerol and kept at �20 �C. The concentration of the enzymesamples was not larger than 2 mg/ml. The single chromatographicpurification step resulted in higher than 90% homogenous, full-length and active hOGA sample (Fig. S8). Following the quality ofthe enzyme samples during the long time storage, their enzymeactivity and homogeneity were maintained even after 12 months(Fig. S8).

2.4. Inhibitory potency of 2-acetamido-2-deoxy-D-glucono-1,5-lactone semicarbazone derivatives on hOGA and HexA/HexBactivities

Michaelis-Menten kinetic parameters for p-nitrophenyl-GlcNAc(pNP-GlcNAc) as substrate in the presence of hOGA established inpotassium-phosphate buffer at pH 7.5 resulted in 20 mM for KM and1.6 s�1 for kcat values (Table 2). The determined kcat value correlatedwell with the previously published data [44]. However, the enzymein this study displayed much higher binding affinity toward pNP-GlcNAc than was previously reported for different conditions(Table 2). Kinetic analysis of hOGA and HexA/HexB inhibition by 2-acetamido-2-deoxy-D-glucono-1,5-lactone semicarbazone de-rivatives was performed by using Dixon method [45] and animproved method presented by Cornish-Bowden [46] (Fig. 3).Plotting the reciprocal initial rate against inhibitor concentration,the intersection point of the obtained curves at three differentsubstrate concentrations was located in the fourth quadrant andgave the inhibitor constants (Ki) for all the investigated compounds(Table 3). The location of the intersection point also refers to themode of inhibition, either competitive or mixed. Since Dixon plotcannot distinguish unambiguously between these inhibitorymodes, we also plotted the kinetic data using the Cornish-Bowdenmethod (Fig. 3). The resulting lines are parallel in the plot of [S]/vagainst inhibitor concentration indicating competitive inhibition ofhOGA and HexA/HexB enzymes by each of the 2-acetamido-2-deoxy-D-glucono-1,5-lactone semicarbazone derivatives 6.

Inhibitor constants (Ki) for most of the compounds 6 againsthOGA are in the submicromolar, in some cases in the nanomolarrange, and slightly higher Ki values have been obtained for HexA/

Table 2Michaelis-Menten kinetic parameters of hOGA for pNP-GlcNAc.

Buffer 50 mMK2HPO4 - KH2PO4 buffer, pH 7.5

PBS (10 mM ph137 mM NaCl, 2

KM(mM) 20 ± 2 170 ± 10kcat (s�1) 1.6 ± 0.1 1.8 ± 0.1

5

HexB (Table 3). The most potent inhibitors of hOGA are semi-carbazones 6e and 6g, while thiosemicarbazone 6h proved to bethe weakest one.

For the determination of the inhibitor constants of the mostpotent inhibitors (6e and 6g), the Dixon method could not be usedin the presence of the pNP-GlcNAc substrate because the inhibitorconcentration was comparable to the enzyme concentration.Therefore, the fluorogenic 4-methylumbelliferyl-GlcNAc (4-MU-GlcNAc) substrate was used to enable the application of the moresensitive fluorometric detection method. This allowed a significantdecrease in the applied enzyme concentration for the inhibitionstudies. In a separate experiment with 6f we demonstrated thatsubstituting the chromogenic substrate (pNP-GlcNAc) with a fluo-rogenic one (4-MU-GlcNAc) did not influence the binding affinity.The obtained Ki values for 6f in the presence of pNP-GlcNAc and 4-MU-GlcNAc were 0.506 ± 0.020 mM and 0.509 ± 0.036 mM,respectively (Fig. S9).

The selectivity of inhibition of the two enzymes was moderate,and this feature resembles PUGNAc. It is worth noting the consid-erable difference in the inhibition by naphthyl derivatives 6f and6g, of which the 2-naphthyl 6g is more than one order of magni-tude more potent than the 1-naphthyl 6f counterpart. On the otherhand, 6f is more selective towards HexA/HexB.

2.5. In silico predictions analysis

In order to gain an understanding of the factors governing theobserved potencies for hOGA inhibition, computations were per-formed in the form of Prime protein-ligand refinements [48], fol-lowed by higher level QM/MM optimizations and QM/MM-PBSAbinding free energy calculations (DGbind, Eq. (1)).

DGbind ¼ DEQM/MM þ DGsolv e TDSMM [1]

Eq. (1) includes contributions from gas-phase protein-ligandinteraction energies (DEQM/MM), changes in solvation free energy onbinding (DGsolv), as well as the ligand entropy changes (TDSMM). Inthese calculations, side chains Tyr219, Phe223, Val254, Val255 andTyr286 in the vicinity of 6a-g ligand phenyl or naphthyl rings wereincluded as part of the QM region, selected due to the importance ofthese protein-ligand interactions to the observed potency differ-ences, and to better describe p-p interactions with Tyr286. Becauseof the potential for a better description of binding site interactionsusing QM methods, QM/MM-PB(GB)SA approaches have recentlybeen applied in a number of studies [49], including where protein-ligand p-p stacking interactions play a key role [50,51]. Here, theQM region was modelled using DFT at the M06-2X/6-31þG** levelof theory, with theM06-2X functional recognised for its descriptionof non-covalent interactions [52e54]. We should note that nativeligand docking calculations with Glide [48] failed to reproduce thePUGNAc ligand crystallographic conformation (PDB code: 5UHO);more specifically the critical orientation of the phenylurethane (-O-CO-NH-Ph) moiety, so that the Prime refinement, QM/MM opti-mization and QM/MM-PBSA approach as fully outlined in theexperimental details was applied.

The results of the QM-MM/PBSA calculations for PUGNAc and

osphate buffer,.7 mM KCl) pH 7.4 [44]

0.5 M citrate/phosphate buffer, pH 6.5 [47]

1100 ±1003.4 ± 0.3

Page 7: Nanomolar inhibition of human OGA by 2-acetamido-2-deoxy …

Fig. 3. Plotting hOGA inhibition data in the presence of 6a at three different pNP-GlcNAc concentrations (- 75 mM; � 40 mM;: 20 mM) by using Dixon method (A) and Cornish-Bowden method (B).

Table 3The binding affinities of 2-acetamido-2-deoxy-D-glucono-1,5-lactone semicarbazone derivatives toward hOGA and HexA/HexB enzymes (Ki [mM]).

Compound hOGA HexA/HexB

6a 0.190 ± 0.008 0.205 ± 0.014 (lit.: 0.130 ± 0.020) [29]

6b 0.155 ± 0.003 0.332 ± 0.017

6c 0.167 ± 0.006 0.125 ± 0.004

6d 0.270 ± 0.010 0.413 ± 0.014

6e 0.083 ± 0.004a 0.170 ± 0.004

6f 0.506 ± 0.020 0.154 ± 0.006

6g 0.036 ± 0.001a 0.047 ± 0.002

6h 27 ± 3 30 ± 2

a 4-MU-GlcNAc was used as substrate.

M. Kiss, E. Szab�o, B. Bocska et al. European Journal of Medicinal Chemistry 223 (2021) 113649

the eight inhibitors 6a-h studied in this work are shown in Table 4.The final predicted DGbind values for each inhibitor are broken downinto the different contributions fromDEQM/MM, DGsolv, and TDSMM. In

6

terms of reproducing the relative potencies, good agreement withexperiment was obtained (the absolute values of predicted DGbind

are larger than experiment, an expected consequence of the

Page 8: Nanomolar inhibition of human OGA by 2-acetamido-2-deoxy …

Table 4QuantumMechanics ⁄MolecularMechanicse PoissoneBoltzmann Surface Area (QM ⁄MM-PBSA) results for the estimation of hOGAe ligand binding free energies using Eq. (1).The QM region was modelled at the M06-2X/6-31þG** level of theory; MM region using the OPLS-AA (2005) forcefield. All energy values are in kcal/mol.

Ligand Predicted Experimental

DEQM/MM DGsolv TDSMMa DGbind DGbind

PUGNAc �136.2 96.2 �21.1 �18.9 �10.4

6a �140.7 101.9 �21.8 �17.0 �9.5

6b �143.1 103.3 �23.0 �16.8 �9.7

6c �152.7 111.8 �23.3 �17.6 �9.6

6d �147.7 109.2 �22.9 �15.6 �9.3

6e �148.4 107.8 �22.2 �18.4 �10.0

6f �152.2 105.8 �23.0 �23.4 �8.9

6g �153.8 106.8 �22.7 �24.3 �10.6

6h �145.9 119.0 �19.8 �7.1 �6.5

a Values calculated at 310 K to be consistent with the performed kinetic experiments. Experimental DGbind ¼ RT ln Ki.

M. Kiss, E. Szab�o, B. Bocska et al. European Journal of Medicinal Chemistry 223 (2021) 113649

method approximations [50,51,55]). PUGNAc was correctly pre-dicted as more potent than the phenylsemicarbazone analogues6a-e. PUGNAc and 6a differ structurally only in terms of their -O-C(O)-NH- and -NH-C(O)eNH- moieties, respectively. Compared toPUGNAc, 6a (and the other semicarbazone analogues, c.f. 6e com-plex in Fig. 4 (A)) is able to exploit a hydrogen bond interactionfrom N2H (c.f. numbering scheme in Table 4) with the Tyr219 sidechain hydroxyl O atom. While protein-ligand interactions based onDEQM/MM are predicted as stronger for 6a (�140.7 kcal/mol)compared to PUGNAc (�136.2 kcal/mol), the desolvation costs for6a are greater, as well as entropy costs accounted for as loss ofligand vibration, rotation and translation on binding. Thio-semicarbazone 6h (a C]O to C]S replacement in the linker) wascorrectly predicted as the least potent ligand. Although its protein-

7

ligand (DEQM/MM) interactions are predicted as strong (�145.9 kcal/mol), the desolvation cost of binding (119.0 kcal/mol) was thehighest of all inhibitors in Table 4.

Analyzing the DGbind values for all the phenyl analogues 6a-e,compounds 6d and 6e are correctly predicted as the least and mostpotent inhibitors within this series, respectively; the remaininginhibitors in this series (6a-c) displayed similar experimental po-tencies (0.155e0.190 mM) and had predicted potency ranks in be-tween. Considering the electron donating or withdrawingproperties of 6a-e para-substituents, there is no obvious correlationbetween these properties and the observed potencies. Indeed,considering the contributions to DGbind, the relative potencies of6a-e are predicted to be a sensitive balance between protein-ligandinteractions (DEQM/MM) and desolvation effects; entropy to a lesser

Page 9: Nanomolar inhibition of human OGA by 2-acetamido-2-deoxy …

Fig. 4. Predicted model complexes from Prime, QM/MM and QM/MM-PBSA calculations: (A) hOGA-6e and (B) hOGA-6g.

M. Kiss, E. Szab�o, B. Bocska et al. European Journal of Medicinal Chemistry 223 (2021) 113649

degree. Given the good agreement between relative DGbind pre-dicted and experimental values for the phenyl substituted com-pound series (Spearman rank correlation Rs ¼ 0.893 (p < 0.01);Pearson correlation Rp ¼ 0.993 (p < 0.01) e for PUGNAc, 6a-e and6h), the predicted binding interactions could be analysed withconfidence.

The most potent phenyl analogue 6e has the strongest protein-

8

ligand interactions (DEQM/MM ¼ �148.4 kcal/mol), and its predictedprotein-ligand complex is shown in Fig. 4 (A). The crystallographicN-acetyl-glucosaminemoiety interactions in hOGAe PUGNAc (PDBcode: 5UHO) are, as expected, conserved in the predicted com-plexes for the newly designed ligands 6a-h. These include keyhydrogen bond interactions of inhibitor O40 and O60 hydroxyls (c.f.Table 4) with Asp285 side chain carboxylate; O40-hydroxyl O with

Page 10: Nanomolar inhibition of human OGA by 2-acetamido-2-deoxy …

M. Kiss, E. Szab�o, B. Bocska et al. European Journal of Medicinal Chemistry 223 (2021) 113649

the Asn313 side chain amide; and O30-hydroxyl with both Gly67backbone O and Lys98 side chain. The N-acetyl group is involved inhydrogen bonds with Asp174 side chain carboxylate and Asn280side chain amide. Considering the binding interactions of thephenyl substituents of 6a-6e (and 6h), there are favourable in-teractions with the hydrophobic side chains of Tyr219, Phe223,Val254, Val255 and Tyr286. Of particular significance are edge-to-face T-shaped p-p stacking with Tyr286. Although the rings arenot ideally aligned, the p-p stacking ring centroid e ring centroiddistances for 6a e 6e were in the range 5.3e5.5 Å (5.4 Å for 6e inFig. 4 (A)), close to theoretical optimum intermonomer distances(~4.8e5.0 Å) in simple model systems [56] and within the observedgeometric preferences (5.0e5.6 Å) of protein e ligand (Phe e

phenyl) T-shaped p-p interactions from Protein Data Bank analysis[57]. The protein-ligand ring planes had intersection angles83.5�e86.8�, that are also consistent with the T-stacking ProteinData Bank analysis [57]. Accounting for phenyl substituent effectsfor T-shaped configurations can be somewhat more complex thanfor sandwich arrangements [58]; in our situation (6a-e), the ligandp e p interactions with Tyr286, are accompanied by additionaldirect interactions between the ligand phenyl substituents and thephenol ring of Tyr286.

Of further significance, we observed the potential for bindingsite loop Val255 (backbone) NH e p interactions with the designedcompounds. For the 6e complex (Fig. 4 (A)), the Val255 NH tocentroid of 6e phenyl ring predicted distance was relatively long at3.8 Å. This analysis, in part, was the basis of our decision to explorenaphthyl substituents that could better exploit such interactions(vide infra). Compound 6e, however, had the additional potential toform halogen Cl e hydrogen bond donor (HBD) interactions withVal255 NH. A quite recent QM study and survey of crystal structuredata demonstrated that halogen (X) e HBD interactions can makesignificant contributions to ligand�protein binding [59]. Geometricparameters considered by the authors included the angleCeXeHBD and the distance HBD e X, with the HBD defined by itsoxygen, nitrogen or sulfur atom. The survey revealed CeXeHBDangle maxima in the plots in the range ~80�e110�. While the pre-dicted CeXeHBD angle (66�) and HBD e X distance (4.9 Å) for thehOGA e 6e predicted model complex are also not optimum forhalogen e HBD interactions, these interactions can occur over abroad range of angles and wemust also consider that Val255 is partof a loop. Additionally, from the crystal structure data analysis,60�e70� angles for nitrogen as HBD were associated with mainlylonger 4.0e4.5 Å distances. We can therefore postulate that theVal255 NHeCl interaction may contribute to the improved potencyof 6e compared to the other analogues 6a-6d.

As mentioned, it was decided to explore whether substitution ofthe phenyls of 6a-6e with naphthyl moieties could better exploitpotential NH e p contacts with Val255, but also the hydrophobicnature of other surrounding residues. Indeed, predictions for both1- (6f) and 2-naphthyl (6g) suggested improved potency, withpredicted DGbind values of �23.4 and �24.3 kcal/mol, respectively.And while the prediction was incorrect for 6f, potentially due toconformational entropy effects (over-prediction of 1-naphthalenecompound analogues extends to other methods and systems[60]), 6g was experimentally the most potent of all ligands tested.The predicted binding of 6g with a 2-naphthyl moiety did asanticipated lead to better Val255 NHe p contacts (3.4 Å to centroidof ring extension), as well as preserving the edge-to-face T-shapedcontacts with Tyr286 (Fig. 4 (B)). The DEQM/MM value of�153.9 kcal/mol was considerably more favourable (>5 kcal/mol) to those ob-tained for the substituted phenyls. An alternative parallel confor-mation p-p binding mode with Tyr286 was also predicted for 6g,but was less favourable (DGbind ¼ �11.4 kcal/mol) compared to theT-shaped conformation (DGbind ¼�25.0 kcal/mol); in agreement, T-

9

shaped binding conformations are reported as more predominantfor solved protein-ligand complexes [57].

3. Conclusion

A series of 2-acetamido-2-deoxy-D-glucono-1,5-lactone 4-aryl(thio)semicarbazone derivatives was obtained by a five-stepsynthetic sequence from D-glucosamine with overall yields of15e55%. These compounds represent the first analogues of PUGNAcwith modified linkers between the sugar and the aromatic moiety.An efficient heterologous expression protocol for full length andactive recombinant hOGA was developed by E. coli strain selectionand optimization of cultivation conditions. Thermal denaturationstudies revealed environmental factors that most affected hOGAstability, so that hOGA samples displayed long-term stability andensured reproducibility in the evaluation of inhibitory potencies.Kinetic studies with hOGA and hexosaminidases A and B showedcompetitive inhibition for all synthesized semicarbazone de-rivatives (hOGA: range of Ki-s ~30e250 nM; HexA/HexB: range ofKi-s ~50e400 nM) with a modest selectivity towards hOGA. Theselactone semicarbazones are therefore alternatives of the OGA in-hibitor PUGNAc in terms of both synthetic efficiency and inhibitoryefficacy. A computational approach consisting of Prime protein-ligand refinements, post-processed by QM/MM optimizations andcalculation of binding free energies using QM/MM-PBSA provedeffective, for the most part, to describe the interactions governingthe observed ligand potencies. The most potent inhibitor 6g(Ki ¼ 36 nM) had a 2-naphthyl substituent that better exploitedinteractions (including NH e p and T-shaped p e p stacking), withsurrounding hydrophobic residues. Further computational andsynthetic work to enhance hOGA selectivity is in progress and willbe disclosed in due course.

4. Experimental

4.1. Syntheses

4.1.1. General methodsMelting points were measured on a Kofler hot-stage and are

uncorrected. Optical rotations were determined with aPerkineElmer 241 or Jasco P-2000 (Easton, MD, USA) polarimetersat room temperature. NMR spectra were recorded by Bruker AMAvance 360 (360/90 MHz for 1H/13C) spectrometer. Chemical shiftsare referenced to TMS as the internal reference (1H), or to the re-sidual solvent signals (13C). 2D 1He13C HMBC experiment [32]optimized to long-range heteronuclear coupling of 8 Hz, and 2D1He1H EASY ROESYexperiment [33] with 300msmixing timewereperformed on a Bruker Avance II 500 MHz spectrometer equippedwith a 5 mm z-gradient TXI probe. Mass spectra were recordedwith MicroTOF-Q type Qq-TOF MS and maXis II UHR ESI-QTOF MS(Bruker Daltonik, Bremen, Germany) instruments in positive ionmode with electrospray ionization or atmospheric pressurechemical ionization technique. TLC plates were visualized under UVlight, and by spray reagent with gentle heating (the plate wassprayed with the following solution: abs. EtOH (95 ml), cc H2SO4(5 ml), anisaldehyde (1 ml)). For column chromatography Kieselgel60 (Merck, particle size (0.063e0.200 mm) was applied. Dichloro-methane was distilled from P2O5 and stored over 4 Å molecularsieves. Pyridine was distilled from KOH and stored over KOH. CrO3was stored over P2O5. 4-Phenylsemicarbazide (2a) and 4-phenylthiosemicarbazide (2h) were purchased from Sigma-Aldrich Ltd. (Budapest, Hungary). 2-acetamido-3,4,6-tri-O-acetyl-2-deoxy-a,b-D-glucopyranose (1) [30,31], substituted phenyl-semicarbazides (2b-e) and 1- or 2- naphthylsemicarbazides (2f-g)[61] were synthesized according to literature procedures. Unless

Page 11: Nanomolar inhibition of human OGA by 2-acetamido-2-deoxy …

M. Kiss, E. Szab�o, B. Bocska et al. European Journal of Medicinal Chemistry 223 (2021) 113649

otherwise indicated, all chemicals used in the biochemical experi-ments were analytical-grade and purchased from Sigma-AldrichLtd. (Budapest, Hungary).

4.1.2. General procedure I. for the synthesis of 1-(2-acetamido-3,4,6-tri-O-acetyl-2-deoxy-b-D-glucopyranosyl)-(thio)semicarbazides (3a-h)

A solution of 200 mg (0.58 mmol) 2-acetamido-3,4,6-tri-O-acetyl-2-deoxy-a,b-D-glucopyranose (1), 0.87 mmol (thio)semi-carbazide and 0.058 mmol p-toluenesulfonic acid in 4 ml chloro-form were boiled under reflux until the reactions were complete(monitored by TLC hexane: acetone ¼ 3 : 2). The mixtures werecooled in ice-bath, the white precipitates were filtered. If the re-agents contaminated the products, column chromatography wasused (hexane: acetone 1 : 1).

4.1.3. General procedure II. for the synthesis of 2-acetamido-3,4,6-tri-O-acetyl-2-deoxy-D-glucono-1,5-lactone semicarbazones (5a-h)

Method A: A solution of 4 equivalents of anh. pyridine in 2 ml ofanh. CH2Cl2 was stirred at 25 �C as 2 equivalents of dry CrO3 wereadded in one portion. The reactionwas slightly exothermic, and thesolution turned burgundy. After 20 min at 25 �C, the solution wascooled to 0 �C and a suspension of 100mg compound in 3ml of anh.CH2Cl2 was added in one portion. The reaction mixture was stirred,while it slowly warmed up to room temperature. After the reactionwas complete (monitored by TLC hexane: acetone 1 : 1) it wasdiluted with 4 ml of Et2O. The mixture was filtered under vacuumand the filtrate concentrated. The pyridinewas removed under highvacuum, and the residue was purified by column chromatography(hexane: acetone 2:1).

Method B: Activated MnO2 was added to a solution of 100 mgcompound in anh. CH2Cl2. Themixturewas boiled until the reactionwas complete (hexane: acetone 1:1), the MnO2 was filtered undervacuum and the filtrate concentrated, the residue was purified bycolumn chromatography (hexane: acetone 2:1).

4.1.4. General procedure III. for the synthesis of 2-acetamido-2-deoxy-D-glucono-1,5-lactone semicarbazones (6a-h)

A solution of 100 mg compound in 1 ml of MeOH was stirred at25 �C with 1 ml of a sat. NH3/MeOH solution. After the reactionwascomplete (monitored by TLC, CHCl3: MeOH 7:3) the solvent wasremoved and the residue was purified by column chromatographyor was recrystallized (MeOH).

4.1.5. Synthesis and characterization of the new compoundsDetailed synthetic procedures and compound characterization

data can befound in the electronic supplementary information.

4.2. Biochemical materials and methods

4.2.1. Microorganisms and culture conditionsFive different Escherichia coli strains were used for the expres-

sion of the long isomorph of hOGA: E. coli BL21(DE3), E. coliBL21(DE3) pLysS (from Novagen), E. coli BL21- CodonPlus (DE3)-RIPL (from Stratagene) and E. coli BL21 STAR(DE3) (from Invitrogen)and E. coli BL21(DE3) Rosetta. hOGA encoding gene was insertedbetween NdeI and XhoI restriction site of pET28b plasmid carryingkanamycin resistance. The resulting vector construction was kindlyprovided by prof. D. J. Vocadlo from Simon Fraser University (Bur-naby, Canada) [62]. The vector construct carrying hOGA encodinggene was transformed into the five E. coli strains by the heat-shockmethods [63]. Resulting transformant cells of the different strains,were subjected to the protein expression procedure.

10

4.2.2. Heterologous expression of human OGA in E. coli4.2.2.1. Small scale protein expression. Small scale expressions wereperformed in 250 ml Luria broth (LB) growth medium supple-mented with 50 mg/ml kanamycin antibiotics in 1L Erlenmeyerflasks and inoculated with 1 ml starter cultures of the differentE. coli strains. They were grown overnight from a single colony, at37 �C. The growing temperature of the small-scale expression wasvaried between 16 and 37 �C, specifically 16 �C, 23 �C and 28 �C.OGA enzyme was expressed by induction with 0.4 mM IPTG atabout mid-log phase, OD600 ~0.6 or 0.9. Bacterial cells were har-vested at stationary phase (OD600 ~ 1.4) by centrifugation at6000 rpm for 20 min at 4 �C. The isolated pellet (~1.2e1.6 g) waskept at �20 �C for the later isolation procedures. The frozen pelletwas re-suspended in 30 ml of lysis buffer (50 mM potassium-phosphate buffer, pH 7.5, containing 1 mM DTT, 0.5% Tween 20,Protease inhibitor cocktail (P8849, Sigma Aldrich, Hungary) and0.1 mg/ml lysozyme) and kept on ice for 30 min. In addition tolysozyme, lysis was also performed by using an Ultrasonic Ho-mogenizer UW 2070 (from Bandelin Electronic), at 4 �C. Theresulting lysates were incubated with DNase in the presence of1 mMMgCl2, for 120min, at 10 �C. Soluble fraction of the lysatewasseparated by centrifugation at 15,000 rpm, at 4 �C and was kept at4 �C for the chromatographic separation which followed. Duringthe expression and lysis, samples were collected for further char-acterization by gel electrophoresis, hexosaminidase activity testingand protein content determination, by the Bradford method, usingBSA for the calibration curve [64].

4.2.2.2. Large-scale protein expression. Starter culture of E. coliRosetta (DE3) strain was prepared (7 � 2 ml) to inoculate 500 mlkanamycine (50 mg/ml) supplemented LB medium in Erlenmeyerflasks (7 � 500 ml). The cells were grown at 23 �C, until OD600reached mid-log phase (~0.9), whereupon by adding 0.4 mM IPTGthe induction was carried out. The incubation of the growth wascontinued until the stationary phase, when cells were harvested bycentrifugation at 5000 rpm for 20 min at 4 �C. The collected pelletwas weighed (~27 g) and kept frozen at �20 �C. Lysis was carriedout using the frozen sample by following the same procedurementioned above (“small scale protein expression” section) exceptless lysis buffer was used to suspend the pellet in relation to “pelletmass/lysis volume” ratio (27 g pellet was dissolved in 400 ml lysisbuffer).

4.2.3. Chromatography of hOGAThe long isomorph of hOGA protein was expressed as a fusion

protein with an N-terminal His-tag comprising of hexapeptide ofhistidine. HiPrep IMAC Fast Flow (diameter 16 mm, length 10 mm;GE-Healthcare) column was charged up with Co2þ and pre-equilibrated with 20 mM potassium-phosphate buffer, pH ¼ 7.7,containing 0.5 M KCl, 50 mM imidazole and 0.5% Tween 20 asbinding buffer. The bound hOGA protein was eluted with a50e500 mM linear imidazole gradient in 20 mM potassium-phosphate buffer, pH ¼ 7.7, containing 0.5 M KCl over 20 columnvolumes. The N-acetylglucosaminidase activity and the proteincontent of the fractions were determined. Those with the highestN-acetylglucosaminidase activity were pooled together and ultra-filtrated (Amicon Ultra e 100 kDa, Merck Millipore) for a bufferexchange, into 50 mM potassium-phosphate buffer, pH ¼ 7.5 con-taining 1 mM PMSF, 1 mM DTT, 1 mM EDTA and 30% glycerol. Theenzyme sample was stored at �20 �C in small aliquots.

4.2.4. N-acetylglucosaminidase activity assayThe N-acetylglucosaminidase activity of hOGA enzyme was

determined using pNP-GlcNAc substrate. Upon enzyme action, p-nitrophenol is liberated from the substrate in the enzyme-substrate

Page 12: Nanomolar inhibition of human OGA by 2-acetamido-2-deoxy …

M. Kiss, E. Szab�o, B. Bocska et al. European Journal of Medicinal Chemistry 223 (2021) 113649

reaction mixture. Quenching the enzyme effect by adding strongalkaline solution to the reaction mixture, the liberated p-nitro-phenol product was detected by measuring the absorbance at400 nm. For quantitative product estimation, the p-nitrophenolcalibration curve was determined in a solution that contains onevolume of the assay buffer and two volumes of the 0.2 M Na2HPO4/NaOH solution, pH ¼ 12.0. N-acetylglucosaminidase activitydetermination was performed in 200 ml of 50 mM potassiumphosphate buffer, pH ¼ 7.5, containing 0.2 mM pNP-GlcNAc as asubstrate, which was pre-incubated in 37 �C. The enzyme reactionwas started by adding a 50 ml hOGA sample to the above solution,incubating the reaction mixture at 37 �C, for 5 min. This reactionwas then stopped by adding 500 ml of 0.2 M Na2HPO4/NaOH solu-tion, pH¼ 12.0. The absorbance at 400 nm of the quenched reactionwas followed and the enzyme activity was calculated in units ofmmol/min.

4.2.5. Stability studies on hOGAThermal stability of hOGA was measured by incubating the

enzyme at 50 �C in different solutions of varying the present cationand its concentration, the type of reducing agents, the type of de-tergents and their concentrations. In all cases, the stability studieswere carried out by incubating 0.33 mg/ml hOGA at 50 �C, and, atregular intervals (5 min for the first half hour, then every 10 min)samples (10 ml) were taken and were placed on ice. For the incu-bation of hOGA samples, Thermalmixer MKR 13 (by HLC BioTech)was used by applying 200 rpm shaking. To the incubated samples,placed on ice, 40 ml distilled water was added then the N-acetyl-glucosaminidase activity was measured by following the protocoldescribed above. The sodium and potassium ion influence on theenzyme thermal stability was tested in 50 mM phosphate buffer,pH 7.5, containing 0.3 M KCl and 0.3 M NaCl. Thermal stability ofhOGA in the presence of different reducing agents such as DTT(1 mM) and b-mercaptoethanol (1 mM) was also determined in50 mM phosphate buffer, pH 7.5 as well as the effect of differentdetergents including 0.5% Tween 20, 0.5% Triton X-100 and glycerolin the same buffer. In all measurements the ratios of the N-ace-tylglucosaminidase activity determined at a given time interval (At)and at zero time (A0) were calculated as residual activities (At/A0).Residual activity was then plotted against time. The data pointswere fitted to a single step transition, from the native enzymeconformation to the denatured enzyme form (EN/ED), that fol-lowed first order kinetics: d[EN]/t¼ -kD[EN], where kD is the thermaldenaturation rate (min�1). The active enzyme concentration ([EN])is expressed as the residual enzyme activity (At/A0). Fitting residualactivity vs incubation time according to the single exponentialdecay curve, ln At/A0 will be constant and gives kD. For data fittingand graphic representation Origin-Pro-8 (OriginLab Corp. North-umpton, MA, USA) was used.

Data points for residual activity versus time, giving sigmoidalcurve, were modelled by using logistic equation (Eq. (2).):

y¼A1þ A2� A11þ 10ðlogx0�xÞp [2]

where A1 is the bottom asymptote, A2 top asymptote, logxo center,p: slope factor that decides the steepness of the curve [65]. Thelogistic function is used for describing the rate of autocatalyticprocesses [66]. In this case y is residual activity, x is time, logxo thetime needed for the denaturation of the half protein population(half-life).

Residual activity versus time, which resulted in a biphasic sig-moid curve, can be fitted by biphasic logistic equation (Eq. (3).):

11

y¼A1þ ðA2�A1Þ�

p1þ 10ðlogx01�xÞh1 þ

1� p1þ 10ðlogx02�xÞh2

�[3]

where A1 is the bottom asymptote, A2 top asymptote, logxo1 centreof the first phase, logxo2 centre of the second phase, h1 slope of thefirst phase, h2 slope of the second phase, p proportion between thephases. In this case, two-state transition was observed during thedenaturation of the protein, the inflexion points (logxo1, logxo2) ofthe biphasic sigmoid curve gives the time (t1/2) needed to reachhalf-maximal of each phase.

4.2.6. Kinetic studiesFor determining the Michaelis-Menten constants of OGA for

pNP-GlcNAc and 4-MU-GlcNAc substrates, the enzymatic reactionswere performed in 50mMpotassium phosphate buffer (pH 7.5) in afinal volume of 1 ml, at 37 �C, containing varied final substrateconcentrations. These ranged from 2 mM to 5 mM (for both sub-strates) in the presence of two hOGA concentrations, namely 33 nMand 82 nM in case of pNP-GlcNAc and 3 nM in case of 4-Mu-GlcNAc.Using pNP-GlcNAc substrate the reaction was conducted in 1 mlcuvettes held in the thermostated cell holder of AnalytikJena Spe-cord 250 plus double beam spectrophotometer. On addition of theenzyme, the reactions were monitored continuously at 400 nm.Using fluorogenic 4-MU-GlcNAc substrate, the reactions were setup in a fluorescence cuvette, held in thermostated cell holder ofJasco FP-8200 fluorescence spectrophotometer equipped with a Xelamp light source at 37 �C. The reactions were initiated by theaddition of 4-MU-GlcNAc and the liberated fluorescent product, 4-methylumbelliferone, was followed with excitation at 360 nM, andemission was detected at 450 nM every 10 s for 10 min. Fordetermining the Michaelis-Menten constants of HexA/HexB forpNP-GlcNAc substrate, the enzymatic reactions were performed in50 mM potassium phosphate buffer (pH 5.8) in a final volume of1 ml, at 37 �C. The substrate concentrations used ranged from 5mMto 60 mM in the presence of 5.0 mU of HexA/HexB (from bovinekidney, Sigma Aldrich). Initial velocities were calculated within thelinear region of the reaction progress curves obtained from usingboth the chromogenic and fluorogenic substrates, thenMichaeliseMenten kinetic parameters were calculated throughnonlinear regression of Michaelian saturation curves using Origin-Pro-8 (OriginLab Corp. Northumpton, MA, USA). Initial velocitieswere expressed in mM of liberated product in 1 s as it was deter-mined from the calibration curves of the liberated pNP. All themeasurements were performed in triplicates.

4.2.7. Inhibition studiesAll target compounds were evaluated for their inhibitory ac-

tivities against hOGA and HexA/HexB. Inhibition studies wereperformed at three substrate concentrations (20 mM, 40 mM and75 mM for pNP-GlcNAc and 7.5 mM, 12.5 mM and 25 mM for 4-MU-GlcNAc) in a reaction mixture containing 50 mM potassium phos-phate buffer (pH 7.5), final concentration of synthesized inhibitors(6a-g) ranging from 0 to 3 mM or from 0 to 200 mM for 6h, in thepresence of 33 nM hOGA. In experiments using fluorogenic sub-strate, 3 nM hOGA was applied in the presence of 6e and 6g in-hibitors, their concentrations ranged from 0 nM to 100 nM. In caseof HexA/HexB the pNP-GlcNAc concentrations were 0.5 mM,1.0 mM and 1.5 mM, except for 6e and 6h, where concentrationswere 0.8 mM, 1.0 mM and 1.5 mM. The concentration of the in-hibitors was in the 0e1 mM range, except for 6h where the rangewas from 0 to 50 mM.

In all cases the enzyme and the inhibitors were pre-incubated at37 �C. The reactions were initiated by the addition of the substratesolution and the kinetic curves were recorded as described above.

Page 13: Nanomolar inhibition of human OGA by 2-acetamido-2-deoxy …

M. Kiss, E. Szab�o, B. Bocska et al. European Journal of Medicinal Chemistry 223 (2021) 113649

Initial velocities were estimated and the reciprocal of initial ve-locities plotted against inhibitor concentrations. All measurementswere performed in triplicates. Inhibition constants (Ki) weredetermined by linear regression of data from Dixon plots [45]. Themodality of the inhibition were assessed for the investigatedcompounds according to the Cornish-Bowden method [46].

4.3. Computational details

4.3.1. Protein preparationhOGAwas prepared for the initial Prime refinement calculations

using Schr€odinger's Protein PreparationWizard [48] and the solved3.21 Å resolution co-crystallized complex of the protein (chains A-D) with PUGNAc (PDB code 5UHO). The waters within 5 Å ofPUGNAc were initially retained (deleted for subsequent calcula-tions), bond orders assigned and hydrogens added, with proton-ation states for basic and acidic residues based on calculated pKa'sat pH 7 from PROPKA [67]. Subsequent optimization of hydroxyls,histidine protonation states and C/N atom flips, as well as side chainO/N atom flips of Gln and Asn residues was based on optimizinghydrogen bonding patterns. The system was then minimized usingOPLS-AA (2005) forcefield [68] with the RMSD (heavy atoms)maintained to within 0.3 Å of the crystallographic positions.

4.3.2. Prime protein-ligand complex refinementsInitial models of protein-ligand complexes for each of the in-

hibitors 6a-h were prepared by mutation of the relevant atoms inthe PUGNAc ligand in chain A of the prepared crystallographichOGA-PUGNAc complex (PDB code: 5UHO; section 4.1.1). For thenaphthyl analogues, rings were extended by fusing at both ortho-meta positions for 1-napthalene and both meta-para positions for2-napthalene. Prime v5.4 protein-ligand refinements [48] werethen performed in local optimization mode for each new complex.Default settings were employed that included the OPLSe forcefield[69] and VSGB model of solvation [70]. Residues within 5 Å ofPUGNAc in 5UHO were selected to be free in all refinements (same360 atoms), with the rest of the protein atoms constrained. Themethod was also applied to the native PUGNAc ligand; comparingthe Prime refined protein-ligand model with the crystal structurecomplex (following backbone superimposition), small RMSDs(heavy atoms) of 0.821 Å and 0.430 Å were obtained for ligand andflexible binding site residues, respectively.

4.3.3. QM/MM and QM/MM-PBSA calculationsThe Prime refined protein-ligand complexes for each ligand

were used as input for QM/MM gas phase optimizations. In thesecalculations, the same protein residues were free and constrainedas per the Prime refinements described above. The QM regionconsisted of the respective ligand and side chains of key bindingsite residues (Try219, Phe223, Val254, Val255 and Tyr286 of chainA) surrounding the phenyl/naphthyl ligand substituents, and wasmodelled using DFT with the M06-2X functional [53] and the 6-31þG** basis set [71,72]; the rest of the systems were modelledusing MMwith the OPLS-AA(2005) forcefield [68]. For the QMwithMM interface, hydrogen caps on the QM region residue side chainswere used. No cut-off was used for non-bonded interactions. TheMM region effectively polarizes the QM region, with interactionsbetween QM and MM regions including electrostatic effects be-tween the MM point charges and the QM wavefunction, as well asvan der Waals terms between QM and MM atoms [73]. The opti-mized complexes were then used as input for single-point QM/MM-PBSA calculations [55], where the binding free energies(DGbind) were calculated as the difference in energies between thebound and the unbound states of the protein-ligand complexesusing Eq. Eq. (1).

12

In Eq. (1), EQM/MM is the gas phase QM/MM energy; and Gsolv, thesolvation free energy. For the latter, bulk solvation (water) effectswere included using Poisson-Boltzmann Surface Area (PBSA)approach [74], with the default solute (internal) dielectric constantof 1.0 used. Otherwise, the QM and MM regions were modelled thesame as for the preceding QM/MM optimizations. The final term inEq. (1) includes a MM estimate (OPLS-AA(2005) forcefield [68]) forthe loss of ligand entropy (DSMM) on binding calculated using theRigid Rotor Harmonic Oscillator (RRHO) approximation, whichconsiders the vibrational, rotational and translational (VRT) en-tropy of the ligands; RRHO calculations were run using Macro-Model v12.2 [48]. All QM/MM and QM/MM-PBSA calculations wereperformed using QSite (Jaguar v10.2; Impact v81012) [48]. Cor-rections for basis set superimposition error (BSSE) as a result ofincomplete basis sets are not included with the Schrodinger qsi-te_binding_energies.py script used for QM/MM-PBSA calculations.However, it was important to probe the potential effects of neglectof BSSE on the predicted relative potencies (reportedDGbind values).Accordingly, Jaguar v10.2 [48] gas phase QM calculations (M06-2X/6-31þG**) on simplified model complexes of the inhibitors wereperformed using an approach previously described [50]. In thesecalculations the complexes consisted only of hydrogen capped sidechains of residues from the QM region (Try219, Phe223, Val254,Val255 and Tyr286) and the ligands. Estimated BSSE effects calcu-lated using the Boys-Bernardi counterpoise (CP) method [75]revealed values ranging 2.5e3.0 kcal/mol and DEQM/MM values werecorrected accordingly using the following equation [50]:

DEQM=MM ¼DEuncorrQM=MM þ BSSE: [4]

Declaration of competing interest

The authors declare that they have no known competingfinancial interests or personal relationships that could haveappeared to influence the work reported in this paper.

Acknowledgement

This work was financially supported by the National Research,Development and Innovation Office of Hungary (Grants: 109450,FK-125067 and PD 135034) and by the EU and co-financed by theEuropean Regional Development Fund under the projects GINOP-2.3.2-15-2016-00008 and GINOP-2.3.3-15-2016-00004. I.T. ac-knowledges the support of the J�anos Bolyai Research Scholarship ofthe Hungarian Academy of Sciences (BO/00372/20/7) and theÚNKP-20-5-DE-262 New National Excellence Program of the Min-istry for Innovation and Technology from the source of the NationalResearch, Development and Innovation Fund.

Appendix A. Supplementary data

Supplementary data to this article can be found online athttps://doi.org/10.1016/j.ejmech.2021.113649.

References

[1] X. Yang, K. Qian, Protein O-GlcNAcylation: emerging mechanisms and func-tions, Nat. Rev. Mol. Cell Biol. 18 (2017) 452e465.

[2] J. Ma, C. Wu, G.W. Hart, Analytical and biochemical perspectives of protein O-GlcNAcylation, Chem. Rev. 121 (2021) 1513e1581.

[3] M.S. Macauley, D.J. Vocadlo, Enzymatic characterization and inhibition of thenuclear variant of human O-GlcNAcase, Carbohydr. Res. 344 (2009)1079e1084.

[4] G.W. Hart, Nutrient regulation of signaling and transcription, J. Biol. Chem.294 (2019) 2211e2231.

Page 14: Nanomolar inhibition of human OGA by 2-acetamido-2-deoxy …

M. Kiss, E. Szab�o, B. Bocska et al. European Journal of Medicinal Chemistry 223 (2021) 113649

[5] D. Wu, Y. Cai, J. Jin, Potential coordination role between O-GlcNAcylation andepigenetics, Protein & Cell 8 (2017) 713e723.

[6] C. Liu, J. Li, O-GlcNAc: a sweetheart of the cell cycle and DNA damageresponse, Front. Endocrinol. 9 (2018).

[7] Y.R. Yang, M. Song, H. Lee, Y. Jeon, E.-J. Choi, H.-J. Jang, H.Y. Moon, H.-Y. Byun,E.-K. Kim, D.H. Kim, M.N. Lee, A. Koh, J. Ghim, J.H. Choi, W. Lee-Kwon, K.T. Kim,S.H. Ryu, P.-G. Suh, O-GlcNAcase is essential for embryonic development andmaintenance of genomic stability, Aging Cell 11 (2012) 439e448.

[8] C. Guinez, A.-M. Mir, V. Dehennaut, R. Cacan, A. Harduin-Lepers, J.-C. Michalski, T. Lefebvre, Protein ubiquitination is modulated by O-GlcNAcglycosylation, Faseb. J. 22 (2008) 2901e2911.

[9] C.M. Ferrer, V.L. Sodi, M.J. Reginato, O-GlcNAcylation in cancer biology: linkingmetabolism and signaling, J. Mol. Biol. 428 (2016) 3282e3294.

[10] W.Y. Wani, J.C. Chatham, V. Darley-Usmar, L.L. McMahon, J. Zhang, O-GlcNAcylation and neurodegeneration, Brain Res. Bull. 133 (2017) 80e87.

[11] C.-X. Gong, F. Liu, K. Iqbal, O-GlcNAcylation: a regulator of tau pathology andneurodegeneration, Alzheimer’s Dementia 12 (2016) 1078e1089.

[12] A.F. Abdel-Magid, Inhibition of O-GlcNAcase (OGA): a potential therapeutictarget to treat alzheimer’s disease, ACS Med. Chem. Lett. 5 (2014) 1270e1271.

[13] X.H. Wang, W.P. Li, J. Marcus, M. Pearson, L.X. Song, K. Smith, G. Terracina,J.L. Lee, K.L.K. Hong, S.X. Lu, L. Hyde, S.C. Chen, D. Kinsley, J.P. Melchor,D.J. Rubins, X.J. Meng, E. Hostetler, C. Sur, L.L. Zhang, J.B. Schachter, J.F. Hess,H.G. Selnick, D.J. Vocadlo, E.J. McEachern, J.M. Uslaner, J.L. Duffy, S.M. Smith,MK-8719, a novel and selective O-GlcNAcase inhibitor that reduces the for-mation of pathological tau and ameliorates neurodegeneration in a mousemodel of tauopathy, J. Pharmacol. Exp. Therapeut. 374 (2020) 252e263.

[14] B.L. Cantarel, P.M. Coutinho, C. Rancurel, T. Bernard, V. Lombard, B. Henrissat,The Carbohydrate-Active EnZymes database (CAZy): an expert resource forGlycogenomics, Nucleic Acids Res. 37 (2009) D233eD238.

[15] Y. He, M.S. Macauley, K.A. Stubbs, D.J. Vocadlo, G.J. Davies, Visualizing thereaction coordinate of an O-GlcNAc hydrolase, J. Am. Chem. Soc. 132 (2010)1807e1809.

[16] M.J. Lemieux, B.L. Mark, M.M. Cherney, S.G. Withers, D.J. Mahuran,M.N.G. James, Crystallographic structure of human b-hexosaminidase A:interpretation of Tay-Sachs mutations and loss of GM2 ganglioside hydrolysis,J. Mol. Biol. 359 (2006) 913e929.

[17] N. Çetinbas, M.S. Macauley, K.A. Stubbs, R. Drapala, D.J. Vocadlo, Identificationof Asp174 and Asp175 as the key catalytic residues of human O-GlcNAcase byfunctional analysis of site-directed mutants, Biochemistry 45 (2006)3835e3844.

[18] B.L. Mark, D.J. Vocadlo, S. Knapp, B.L. Triggs-Raine, S.G. Withers, M.N.G. James,Crystallographic evidence for substrate-assisted catalysis in a bacterial b-hexosaminidase, J. Biol. Chem. 276 (2001) 10330e10337.

[19] S.K. Srivastava, E. Beutler, Hexosaminidase-A and hexosaminidase-B: studiesin Tay-Sachs’ and Sandhoff’s disease, Nature 241 (1973), 463-463.

[20] S.T. Hepbildikler, R. Sandhoff, M. K€olzer, R.L. Proia, K. Sandhoff, Physiologicalsubstrates for human lysosomal b-hexosaminidase S, J. Biol. Chem. 277 (2002)2562e2572.

[21] J. Popko, J. Marciniak, A. Zalewska, P. Małdyk, M. Rogalski, K. Zwierz, Theactivity of exoglycosidases in the synovial membrane and knee fluid of pa-tients with rheumatoid arthritis and juvenile idiopathic arthritis, Scand. J.Rheumatol. 35 (2006) 189e192.

[22] J. Popko, A. Zalewska, S. Olszewski, A. G�orska, S. Sierakowski, K. Zwierz,M. Urban, Activity of N-acetyl-b-hexosaminidase in serum and joint fluid ofthe knees of patients with juvenile idiopatic arthritis, Clin. Exp. Rheumatol. 21(2003) 675.

[23] D.L. Dong, G.W. Hart, Purification and characterization of an O-GlcNAc se-lective N-acetyl-b-D-glucosaminidase from rat spleen cytosol, J. Biol. Chem.269 (1994) 19321e19330.

[24] M.S. Macauley, A.K. Bubb, C. Martinez-Fleites, G.J. Davies, D.J. Vocadlo,Elevation of global O-GlcNAc levels in 3T3-L1 adipocytes by selective inhibi-tion of O-GlcNAcase does not induce insulin resistance, J. Biol. Chem. 283(2008) 34687e34695.

[25] A.A. Elbatrawy, E.J. Kim, G. Nam, O-GlcNAcase: emerging mechanism, sub-strate recognition and small-molecule inhibitors, ChemMedChem 15 (2020)1244e1257.

[26] M. Hattie, N. Cekic, A.W. Debowski, D.J. Vocadlo, K.A. Stubbs, Modifying thephenyl group of PUGNAc: reactivity tuning to deliver selective inhibitors forN-acetyl-D-glucosaminidases, Org. Biomol. Chem. 14 (2016) 3193e3197.

[27] D. Beer, J.-L. Maloisel, D.M. Rast, A. Vasella, Synthesis of 2-Acetamido-2-deoxy-D-gluconhydroximolactone- and chitobionhydroximolactone-derivedN-phenylcarbamates, potential inhibitors of b-N-acetylglucosaminidase,Helv. Chim. Acta 73 (1990) 1918e1922.

[28] H. Mohan, A. Vasella, An improved synthesis of 2-Acetamido-2-deoxy- D-gluconohydroximolactone (PUGNAc), A strong inhibitor of b-N-acetylgluco-saminidases, Helv. Chim. Acta 83 (2000) 114e118.

[29] D.R. Wolk, A. Vasella, F. Schweikart, M.G. Peter, Synthesis and enzyme-inhibition studies of phenylsemicarbazones derived from D-glucono-1,5-lactone and 2-acetamido-2-deoxy-D-glucono-1,5-lactone, Helv. Chim. Acta75 (1992) 323e334.

[30] D. Chaplin, D.H.G. Crout, S. Bornemann, D.W. Hutchinson, R. Khan, Conversionof 2-acetamido-2-deoxy-b-D-glucopyranose (N-acetylglucosamine) into 2-acetamido-2-deoxy-b-D-galactopyranose (N-acetylgalactosamine) using abiotransformation to generate a selectively deprotected substrate for SN2inversion, J. Chem. Soc., Perkin Trans. 1 (1992) 235e237.

13

[31] K.P. Ravindranathan Kartha, B. Mukhopadhyay, R.A. Field, Practical de-O-acylation reactions promoted by molecular sieves, Carbohydr. Res. 339 (2004)729e732.

[32] A. Bax, M.F. Summers, Proton and carbon-13 assignments from sensitivity-enhanced detection of heteronuclear multiple-bond connectivity by 2Dmultiple quantum NMR, J. Am. Chem. Soc. 108 (1986) 2093e2094.

[33] C.M. Thiele, K. Petzold, J. Schleucher, EASY ROESY: reliable cross-peak inte-gration in adiabatic symmetrized ROESY, Chem. Eur J. 15 (2009) 585e588.

[34] S.A. Goff, L.P. Casson, A.L. Goldberg, Heat shock regulatory gene htpR in-fluences rates of protein degradation and expression of the lon gene inEscherichia coli, Proc. Natl. Acad. Sci. Unit. States Am. 81 (1984) 6647.

[35] V. Tsilibaris, G. Maenhaut-Michel, L. Van Melderen, Biological roles of the LonATP-dependent protease, Res. Microbiol. 157 (2006) 701e713.

[36] S. Luo, M. McNeill, T.G. Myers, R.J. Hohman, R.L. Levine, Lon protease promotessurvival of Escherichia coli during anaerobic glucose starvation, Arch. Micro-biol. 189 (2008) 181e185.

[37] L. Wells, Y. Gao, J.A. Mahoney, K. Vosseller, C. Chen, A. Rosen, G.W. Hart,Dynamic O-glycosylation of nuclear and cytosolic proteins: further charac-terization of the nucleocytoplasmic b-N-acetylglucosaminidase, O-GlcNAcase,J. Biol. Chem. 277 (2002) 1755e1761.

[38] C. Butkinaree, W.D. Cheung, S. Park, K. Park, M. Barber, G.W. Hart, Charac-terization of b-N-acetylglucosaminidase cleavage by caspase-3 duringapoptosis*, J. Biol. Chem. 283 (2008) 23557e23566.

[39] S. Wadhawan, S. Gautam, Rescue of Escherichia coli cells from UV-induceddeath and filamentation by caspase-3 inhibitor, Int. Microbiol. 22 (2019)369e376.

[40] K.W. Bayles, Bacterial programmed cell death: making sense of a paradox, Nat.Rev. Microbiol. 12 (2014) 63e69.

[41] F. Baneyx, M. Mujacic, Recombinant protein folding and misfolding inEscherichia coli, Nat. Biotechnol. 22 (2004) 1399e1408.

[42] F.D. Schramm, K. Schroeder, K. Jonas, Protein aggregation in bacteria, FEMS(Fed. Eur. Microbiol. Soc.) Microbiol. Rev. 44 (2020) 54e72.

[43] D.M. Francis, R. Page, Strategies to optimize protein expression in E. coli,Current Protocols in Protein Science 61 (2010), 5.24.21-25.24.29.

[44] N.L. Elsen, S.B. Patel, R.E. Ford, D.L. Hall, F. Hess, H. Kandula, M. Kornienko,J. Reid, H. Selnick, J.M. Shipman, S. Sharma, K.J. Lumb, S.M. Soisson, D.J. Klein,Insights into activity and inhibition from the crystal structure of human O-GlcNAcase, Nat. Chem. Biol. 13 (2017) 613e615.

[45] M. Dixon, The determination of enzyme inhibitor constants, Biochem. J. 55(1953) 170e171.

[46] A. Cornish-Bowden, A simple graphical method for determining the inhibitionconstants of mixed, uncompetitive and non-competitive inhibitors, Biochem.J. 137 (1974) 143e144.

[47] E.J. Kim, D.O. Kang, D.C. Love, J.A. Hanover, Enzymatic characterization of O-GlcNAcase isoforms using a fluorogenic GlcNAc substrate, Carbohydr. Res. 341(2006) 971e982.

[48] Schr€odinger Release 2018-4, Schr€odinger, LLC, New York, NY, 2018.[49] U. Ryde, P. S€oderhjelm, Ligand-binding affinity estimates supported by

quantum-mechanical methods, Chem. Rev. 116 (2016) 5520e5566.[50] K.E. Tsitsanou, J.M. Hayes, M. Keramioti, M. Mamais, N.G. Oikonomakos,

A. Kato, D.D. Leonidas, S.E. Zographos, Sourcing the affinity of flavonoids forthe glycogen phosphorylase inhibitor site via crystallography, kinetics andQM/MM-PBSA binding studies: comparison of chrysin and flavopiridol, FoodChem. Toxicol. 61 (2013) 14e27.

[51] B.A. Chetter, E. Kyriakis, D. Barr, A.G. Karra, E. Katsidou, S.M. Koulas,V.T. Skamnaki, T.J. Snape, A.-M.G. Psarra, D.D. Leonidas, J.M. Hayes, Syntheticflavonoid derivatives targeting the glycogen phosphorylase inhibitor site:QM/MM-PBSA motivated synthesis of substituted 5,7-dihydroxyflavones,crystallography, in vitro kinetics and ex-vivo cellular experiments reveal novelpotent inhibitors, Bioorg. Chem. 102 (2020) 104003.

[52] C.D. Sherrill, T. Takatani, E.G. Hohenstein, An assessment of theoreticalmethods for nonbonded interactions: comparison to complete basis set limitcoupled-cluster potential energy curves for the benzene dimer, the methanedimer, Benzene�Methane, and Benzene�H2S, J. Phys. Chem. 113 (2009)10146e10159.

[53] Y. Zhao, D.G. Truhlar, The M06 suite of density functionals for main groupthermochemistry, thermochemical kinetics, noncovalent interactions, excitedstates, and transition elements: two new functionals and systematic testing offour M06-class functionals and 12 other functionals, Theor. Chem. Accounts120 (2008) 215e241.

[54] E.G. Hohenstein, S.T. Chill, C.D. Sherrill, Assessment of the performance of theM05�2X and M06�2X exchange-correlation functionals for noncovalent in-teractions in biomolecules, J. Chem. Theor. Comput. 4 (2008) 1996e2000.

[55] S. Manta, A. Xipnitou, C. Kiritsis, A.L. Kantsadi, J.M. Hayes, V.T. Skamnaki,C. Lamprakis, M. Kontou, P. Zoumpoulakis, S.E. Zographos, D.D. Leonidas,D. Komiotis, 30-Axial CH2OH substitution on glucopyranose does not increaseglycogen phosphorylase inhibitory potency. QM/MM-PBSA calculations sug-gest why, Chem. Biol. Drug Des. 79 (2012) 663e673.

[56] M.O. Sinnokrot, C.D. Sherrill, Substituent effects in p�p Interactions: sand-wich and T-shaped configurations, J. Am. Chem. Soc. 126 (2004) 7690e7697.

[57] Y. Zhao, J. Li, H. Gu, D. Wei, Y.-c. Xu, W. Fu, Z. Yu, Conformational preferencesof p e p stacking between ligand and protein, analysis derived from crystalstructure data geometric preference of p e p interaction, Interdiscipl. Sci.Comput. Life Sci. 7 (2015) 211e220.

[58] A.L. Ringer, M.O. Sinnokrot, R.P. Lively, C.D. Sherrill, The effect of multiple

Page 15: Nanomolar inhibition of human OGA by 2-acetamido-2-deoxy …

M. Kiss, E. Szab�o, B. Bocska et al. European Journal of Medicinal Chemistry 223 (2021) 113649

substituents on sandwich and T-shaped pep interactions, Chem. Eur J. 12(2006) 3821e3828.

[59] F.-Y. Lin, A.D. MacKerell, Do halogenehydrogen bond donor interactionsdominate the favorable contribution of halogens to ligandeprotein binding?J. Phys. Chem. B 121 (2017) 6813e6821.

[60] D. Barr, E. Szennyes, �E. Bokor, Z.H. Al-Oanzi, C. Moffatt, S. Kun, T. Docsa,�A. Sipos, M.P. Davies, R.T. Mathomes, T.J. Snape, L. Agius, L. Soms�ak, J.M. Hayes,Identification of C-b-D-Glucopyranosyl azole-type inhibitors of glycogenphosphorylase that reduce glycogenolysis in hepatocytes: in silico design,synthesis, in vitro kinetics, and ex vivo studies, ACS Chem. Biol. 14 (2019)1460e1470.

[61] M.N. Aboul-Enein, A.A. El-Azzouny, M.I. Attia, Y.A. Maklad, K.M. Amin,M. Abdel-Rehim, M.F. El-Behairy, Design and synthesis of novel stiripentolanalogues as potential anticonvulsants, Eur. J. Med. Chem. 47 (2012) 360e369.

[62] M.S. Macauley, K.A. Stubbs, D.J. Vocadlo, O-GlcNAcase catalyzes cleavage ofthioglycosides without general acid catalysis, J. Am. Chem. Soc. 127 (2005)17202e17203.

[63] A. Froger, J.E. Hall, Transformation of plasmid DNA into E. coli using the heatshock method, JoVE (2007) 253.

[64] M.M. Bradford, A rapid and sensitive method for the quantitation of micro-gram quantities of protein utilizing the principle of protein-dye binding, Anal.Biochem. 72 (1976) 248e254.

[65] W.W. Focke, I. van der Westhuizen, N. Musee, M.T. Loots, Kinetic interpreta-tion of log-logistic dose-time response curves, Sci. Rep. 7 (2017) 2234.

[66] A.M. Morris, R.G. Finke, a-Synuclein aggregation variable temperature andvariable pH kinetic data: a re-analysis using the FinkeeWatzky 2-step modelof nucleation and autocatalytic growth, Biophys. Chem. 140 (2009) 9e15.

[67] C.R. Søndergaard, M.H.M. Olsson, M. Rostkowski, J.H. Jensen, Improvedtreatment of ligands and coupling effects in empirical calculation andrationalization of pKa values, J. Chem. Theor. Comput. 7 (2011) 2284e2295.

14

[68] G.A. Kaminski, R.A. Friesner, J. Tirado-Rives, W.L. Jorgensen, Evaluation andreparametrization of the OPLS-AA force field for proteins via comparison withaccurate quantum chemical calculations on peptides, J. Phys. Chem. B 105(2001) 6474e6487.

[69] E. Harder, W. Damm, J. Maple, C. Wu, M. Reboul, J.Y. Xiang, L. Wang,D. Lupyan, M.K. Dahlgren, J.L. Knight, J.W. Kaus, D.S. Cerutti, G. Krilov,W.L. Jorgensen, R. Abel, R.A. Friesner, OPLS3: a force field providing broadcoverage of drug-like small molecules and proteins, J. Chem. Theor. Comput.12 (2016) 281e296.

[70] J. Li, R. Abel, K. Zhu, Y. Cao, S. Zhao, R.A. Friesner, The VSGB 2.0 model: a nextgeneration energy model for high resolution protein structure modeling,Proteins: Struct. Func. Bioinf. 79 (2011) 2794e2812.

[71] M.M. Francl, W.J. Pietro, W.J. Hehre, J.S. Binkley, M.S. Gordon, D.J. DeFrees,J.A. Pople, Self-consistent molecular orbital methods. XXIII. A polarization-type basis set for second-row elements, J. Chem. Phys. 77 (1982) 3654e3665.

[72] W.J. Hehre, R. Ditchfield, J.A. Pople, Selfdconsistent molecular orbitalmethods. XII. Further extensions of Gaussiandtype basis sets for use in mo-lecular orbital studies of organic molecules, J. Chem. Phys. 56 (1972)2257e2261.

[73] R.B. Murphy, D.M. Philipp, R.A. Friesner, A mixed quantum mechanics/mo-lecular mechanics (QM/MM) method for large-scale modeling of chemistry inprotein environments, J. Comput. Chem. 21 (2000) 1442e1457.

[74] B. Marten, K. Kim, C. Cortis, R.A. Friesner, R.B. Murphy, M.N. Ringnalda,D. Sitkoff, B. Honig, New model for calculation of solvation free Energies:correction of self-consistent reaction field continuum dielectric theory forshort-range hydrogen-bonding effects, J. Phys. Chem. 100 (1996)11775e11788.

[75] S.F. Boys, F. Bernardi, The calculation of small molecular interactions by thedifferences of separate total energies. Some procedures with reduced errors,Mol. Phys. 19 (1970) 553e566.