Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2020 Sep 10.
Published in final edited form as: Biochemistry. 2019 Aug 27;58(36):3755–3766. doi: 10.1021/acs.biochem.9b00582

Structure and Function of the Acetylpolyamine Amidohydrolase from the Deep Earth Halophile Marinobacter subterrani

Jeremy D Osko 1, Benjamin W Roose 1, Stephen A Shinsky 1,, David W Christianson 1,*
PMCID: PMC6736730  NIHMSID: NIHMS1047609  PMID: 31436969

Abstract

Polyamines are small organic cations essential for cellular function in all kingdoms of life. Polyamine metabolism is regulated by enzyme-catalyzed acetylation-deacetylation cycles in similar fashion to the epigenetic regulation of histone function in eukaryotes. Bacterial polyamine deacetylases are particularly intriguing, since these enzymes share the fold and function of eukaryotic histone deacetylases. Recently, acetylpolyamine amidohydrolase from the deep earth halophile Marinobacter subterrani (msAPAH) was described. This Zn2+-dependent deacetylase shares 53% amino acid sequence identity with the acetylpolyamine amidohydrolase from Mycoplana ramosa (mrAPAH) and 22% amino acid sequence identity with the catalytic domain of histone deacetylase 10 from Danio rerio (zebrafish; zHDAC10), the eukaryotic polyamine deacetylase. The X-ray crystal structure of msAPAH, determined in complexes with 7 different inhibitors as well as the acetate coproduct, shows how the chemical strategy of Zn2+-dependent amide hydrolysis and the catalytic specificity for cationic polyamine substrates is conserved in a subterranean halophile. Structural comparisons with mrAPAH reveal that an array of aspartate and glutamate residues unique to msAPAH enable the binding of one or more Mg2+ ions in the active site and elsewhere on the protein surface. Notwithstanding these differences, activity assays with a panel of acetylpolyamine and acetyllysine substrates confirm that msAPAH is a broad-specificity polyamine deacetylase, much like mrAPAH. Broad substrate specificity contrasts with the narrow substrate specificity of zHDAC10, which is highly specific for N8-acetylspermidine hydrolysis. Notably, quaternary structural features govern the substrate specificity of msAPAH and mrAPAH, whereas tertiary structural features govern the substrate specificity of zHDAC10.

Graphical Abstract

graphic file with name nihms-1047609-f0001.jpg

Introduction

The arginase-deacetylase superfamily of metalloenzymes has been extensively studied in recent years, but much of the phylogenetic tree remains uncharacterized in terms of structure and function.13 Rat arginase I was the first member of this superfamily to yield an X-ray crystal structure,4 and the subsequent crystal structure determination of the structurally homologous histone deacetylase-like protein from Aquifex aeolicus5 led to the designation of their common α/β-metallohydrolase topology as the arginase-deacetylase fold. Additional crystal structures of arginase-related metalloenzymes have since been reported, including proclavaminic acid amidino hydrolase,6 agmatinase,7 and formiminoglutamase.8 Among the deacetylases, crystal structures of class I and class II histone deacetylases (HDACs),3,916 as well as acetylpolyamine amidohydrolase from Mycoplana ramosa (mrAPAH),17,18 have been described.

Structural comparisons indicate that the arginases and deacetylases divergently evolved from a common ancestral α/β-metallohydrolase despite sharing low amino acid sequence identities generally ranging 10–15%.1921 The stoichiometry of and selectivity for catalytic metal ion(s) have also diverged: arginases generally require 2 Mn2+ ions for function,22 whereas deacetylases require a single Zn2+ ion (or possibly a single Fe2+ ion) for function.23 The Zn2+ binding site in the deacetylases is conserved as the Mn2+B binding site of the arginases.1921

Among the diversity of hydrolytic functions represented in the arginase-deacetylase phylogenetic tree, enzymes of polyamine biosynthesis are particularly intriguing. Arginase itself functions in polyamine biosynthesis to generate product L-ornithine, which undergoes decarboxylation to yield putrescine, a building block in the biosynthesis of spermidine and spermine.2427 Generally present at millimolar concentrations in the cell, polyamines serve myriad biological functions such as the stabilization of nucleic acid structure2830 and the regulation of transcription and translation.3133 Notably, polyamines undergo reversible acetylation in cellular trafficking and function, and enzymes that catalyze polyamine deacetylation are structurally and functionally related to HDACs.17,18 Indeed, HDAC10 was recently discovered to be a highly specific N8-acetylspermidine deacetylase.3,34 Although this eukaryotic cytosolic activity was discovered more than 40 years ago,35,36 the responsible enzyme was not identified at the time.

In recent years, Marinobacter species have been identified in harsh marine and non-marine environments, including the deep subsurface based on their isolation in hydraulic fracturing effluent.3739 These species are well adapted to the high salinity and Fe2+ concentrations found in mines, wells, and other deep subsurface locations. The halophile Marinobacter subterrani was recently identified 714 m below the surface in the Soudan iron mine in Minnesota,40 and genomic analysis reveals an acetylpolyamine amidohydrolase (msAPAH, UniProt A0A0J7JFD7) similar to mrAPAH (53% amino acid sequence identity). Both msAPAH and mrAPAH retain approximately 22% amino acid sequence identity with the polyamine deacetylase domain of HDAC10 from Danio rerio (zebrafish; zHDAC10).

Here, we report the X-ray crystal structure of msAPAH, and we show that msAPAH is a broad-specificity polyamine deacetylase (Figure 1a). To illustrate the structural basis of substrate specificity and the catalytic mechanism, crystal structures of msAPAH complexed with the acetate product as well as seven inhibitors (Figure 1b) are presented. These studies provide important insight regarding the relationship between eukaryotic histone deacetylases and prokaryotic polyamine deacetylases.

Figure 1.

Figure 1.

(a) The polyamine deacetylase reaction catalyzed by msAPAH illustrated for substrate acetylputrescine. (b) Inhibitors studied in complex with msAPAH: 1, 5-[(3-aminopropyl)amino]pentylboronic acid; 2, 7-[(3-aminopropyl)amino]-1,1,1-trifluoroheptan-2-one; 3, 5-[(3-aminopropyl)amino]pentane-1-thiol; 4, 6-amino-N-hydroxyhexanamide; 5, 8-amino-N-hydroxyoctanamide; 6, 6-[(3-aminopropyl)amino]-N-hydroxyhexanamide; 7, 4-(dimethylamino)-N-[7-hydroxyamino)-7-oxoheptyl]benzamide (also known as M344). Inhibitor syntheses have been described.18,41

Materials and Methods

Reagents.

In general, all chemicals were purchased from Fisher Scientific, Millipore Sigma, or Hampton Research and used without further purification. Inhibitory polyamine analogues 16 were synthesized as previously described.18,41 Inhibitor 7, 4-(dimethylamino)-N-[7-hydroxyamino)-7-oxoheptyl]benzamide (also known as M344), was purchased from Cayman Chemical.

Protein Preparation.

Full-length msAPAH was recombinantly expressed using a pGEX-6P-1 vector by Genscript. This construct utilizes a BAMHI/Xhol cloning site and contains ampicillin bacterial resistance. An N-terminal GST tag is attached, followed by a PreScission protease cleavage site prior to the start of the protein sequence.

Protein was expressed using Escherichia coli One Shot BL21(DE3) cells (Invitrogen) and grown in 2xYT medium in the presence of 100 μg/mL ampicillin. Cells were grown at 37° C and 250 RPM in an Innova 40 incubator shaker until OD600 reached approximately 0.80. The temperature was then reduced to 18° C until OD600 reached 1.0. At this point, cells were supplemented with 200 μM isopropyl β-L-1-thiogalactopyranoside (IPTG; Gold Biotechnology) and were grown for an additional 18 h at 250 RPM. Cells were then centrifuged for 20 min at 5,000 RPM using a Sorvall LYNX 6000 centrifuge. Cell pellets were stored at −80° C until needed.

The cell pellet was thawed and re-suspended in 100 mL of Buffer A [50 mM 4-(2-hydroxyethyl)-1-piperazineethanesulfonic acid (HEPES) sodium salt (pH 7.5), 250 mM NaCl, 1 mM tris(2-carboxyethyl)phosphine (TCEP), 10% glycerol, and 10 μM ZnCl2]. Additionally, 2 protease inhibitor tablets, 0.1 mg/mL lysozyme, and 50 μg/mL DNAse were added to the solution. The cells were then lysed by sonication. Cell lysate was centrifuged for 1 h at 15,000 RPM using a Sorval LYNX 6000 centrifuge. Cell lysate was then applied to a 5-mL pre-equilibrated GSTrap HP column.

The GST-msAPAH fusion protein bound to the GSTrap HP column and was eluted using Buffer B [50 mM HEPES sodium salt (pH 8.0), 250 mM NaCl, 1 mM TCEP, 10% glycerol,10 μM ZnCl2, and 10 mM reduced glutathione]. Protein containing fractions were collected and digested overnight using 5 mg/mL recombinant PreScission protease in 4 L Buffer A through dialysis (RC dialysis tubing, 6–8 kD molecular weight cut-off).

After protease cleavage for 12 h, the protein digest was applied to the 5-mL GSTrap HP column. The GST tag remained bound to the column and the msAPAH protein flowed through the column. The msAPAH-containing fractions were concentrated to 5 mL using a 15-mL centrifugal filter unit with a molecular weight cut-off of 10 kDa. The protein was then filtered using a 0.22-μM Millex-GV filter unit prior to being loaded onto a HiLoad Superdex 26/600 200 pg column. The column was pre-equilibrated with 360 mL of Buffer C [50 mM HEPES sodium salt (pH 7.5), 200 mM KCl, 1 mM TCEP, and 5% glycerol]. The 5-mL sample of msAPAH was injected at a rate of 1 mL/min, and 5-mL fractions were collected. Pure msAPAH protein was confirmed by SDS-PAGE, concentrated to approximately 10 mg/mL, and stored at −80° C.

Enzyme Kinetics.

To measure the rate of acetylpolyamine hydrolysis catalyzed by msAPAH, the generation of the acetate coproduct was measured using a colorimetric assay kit (Sigma-Aldrich, catalogue number MAK086). Briefly, the reaction mixture (50 μL) was prepared as described in the kit protocol: 37 μL Acetate Assay Buffer, 2 μL Acetate Enzyme Mix, 2 μL ATP, 2 μL Acetate Substrate Mix, 2 μL Probe, and 5 μL of 10 μM msAPAH. Separately, 50 μL substrate solutions were prepared at concentrations ranging 0–5,000 μM in Acetate Assay Buffer.

To initiate the reaction, 50 μL of the reaction mixture was combined with 50 μL of substrate solution in a Greiner 96 well flat-bottom transparent plate, which was immediately placed in a Tecan Infinite M1000Pro plate reader and monitored at 450 nm. The pH of the assay solution was measured to be 7.75. Measurements were made every 10 s for a total of 40 min, with shaking every 5 s. The plate was covered at all times due to the photosensitivity of the probe.

To calculate msAPAH activity, absorbance measurements at 450 nm were converted into product formation using an acetate standard curve according to the kit protocol. Data were analyzed using GraphPad Prism version 7.00 for MAC OS X (GraphPad software, La Jolla California USA, www.graphpad.com). Assays were performed in triplicate at 25° C. Nonlinear regression fits to the Michaelis-Menten equation were used to determine steady-state parameters.

Inhibitory Activity Measurements.

Inhibitory potencies of compounds 17 against msAPAH were determined using the colorimetric acetate assay described above. Briefly, the reaction mixture (50 μL) was prepared as described above, except that the protein was omitted: 42 μL Acetate Assay Buffer, 2 μL Acetate Enzyme Mix, 2 μL ATP, 2 μL Acetate Substrate Mix, and 2 μL Probe. Separately, 20 μL of 2 μM msAPAH protein was prepared with 10 μL inhibitor (prepared in Buffer C) with final inhibitor concentrations ranging from 0–2,000 μM. The inhibitor was incubated with enzyme for 15 minutes prior to the addition of substrate.

The reaction was initiated by the addition of 20 μL of 4,000 μM acetylcadaverine substrate. Immediately after substrate addition, 50 μL of reaction mixture was added to the 50 μL of mixture containing the msAPAH protein, acetylcadaverine substrate, and inhibitor. The reaction was analyzed using a Tecan Infinite M1000Pro plate reader as described above. Assays were run at 25°C, with the absorbance being measured at 450 nm for a total of 100 min.

The IC50 values were calculated from raw data using Prism 7. Data were first transformed using the function x = log(x). Data was then normalized by setting the 0% identity definition to be the smallest mean in each dataset, while the 100% identity definition was set to be the largest mean in each dataset. Finally, the data were fit to a nonlinear regression curve of log(inhibitor) vs. normalized response. All measurements were run in duplicate. Goodness of Fit statistics ranged from 88%−97%. IC50 measurements are recorded in Figure S1.

Isothermal Titration Calorimetry (ITC).

Given the somewhat modest inhibitory potencies for some of the inhibitors studied, we additionally performed direct measurements of enzyme-inhibitor complexation using ITC. Dissociation constants (Kd) for inhibitors 17 to msAPAH were measured using a MicroCal iTC 200 isothermal titration calorimeter (GE Healthcare). For each measurement, 300 μM of inhibitor was titrated into 30 μM msAPAH in 50 mM HEPES (pH 7.5), 150 mM NaCl, 1 mM TCEP-HCl, and 10% glycerol (v/v) at 25 °C. In the case of inhibitor 6, 600 μM inhibitor was titrated into 60 μM of msAPAH. For inhibitor 7, both inhibitor and msAPAH solutions included 5% (v/v) DMSO. Nineteen 2-μL injections were made over 60 min. Enthalpogram data were visualized and analyzed using Origin software (OriginLab, Northampton, MA). Enthalpograms for inhibitors 1, 2, 4, and 7 could not be fit to binding curves. Regardless, the direct calorimetric measurement of enzyme-inhibitor affinity for inhibitors 5 and 6 yielded micromolar dissociation constants while inhibitor 3 yielded a nanomolar dissociation constant (Figure S2).

Crystallization.

All msAPAH-inhibitor complexes were crystallized using the sitting-drop vapor diffusion method. For cocrystallization of all msAPAH complexes, a 350-nL drop of protein solution [10 mg/mL msAPAH in Buffer C, and 2 mM inhibitor in phosphate-buffered saline (PBS)] was added to a 350-nL drop of precipitant solution as outlined below, and equilibrated against 80 μL of precipitant solution in the well reservoir. All crystals appeared within approximately two days at 4° C, except those of the msAPAH-3 complex which grew at 21° C. Ethylene glycol (15% v/v) was added to the mother liquor as a cryoprotectant prior to flash cooling crystal for X-ray diffraction data collection.

For cocrystallization of the msAPAH-acetate complex, the precipitant was 0.2 M calcium acetate hydrate, 20% w/v PEG 3350, which yielded thick cubic crystals. For cocrystallization of the msAPAH-1 complex, the precipitant was 0.2 M magnesium chloride hexahydrate, 0.1 M HEPES (pH 7.5), 25% w/v PEG 3350, which yielded long rod-like crystals. For cocrystallization of the msAPAH-2 complex, the precipitant was 0.2 M magnesium acetate tetrahydrate, 20% w/v PEG 3350, which yielded thick plate-like crystals. For cocrystallization of the msAPAH-3 complex, the precipitant was 20% v/v 2-propanol, 0.1 M MES monohydrate (pH 6.0), 20% w/v polyethylene glycol monomethyl ether 2000, which yielded long rod-like crystals. For cocrystallization of the msAPAH-4 complex, the precipitant was 0.2 M magnesium acetate tetrahydrate, 20% w/v PEG 3350, which yielded thick plate-like crystals. For cocrystallization of the msAPAH-5 complex, the precipitant consisted of 0.2 M magnesium acetate tetrahydrate, 20% w/v PEG 3350, which yielded thick plate-like crystals. For cocrystallization of the msAPAH-6 complex, the precipitant was 0.2 M calcium chloride dihydrate, 20% w/v PEG 3350, which yielded thick plate-like crystals. For cocrystallization of the msAPAH-7 complex, the precipitant was 0.2 M calcium chloride dihydrate, 20% w/v PEG 3350, which yielded thick plate-like crystals.

Data Collection and Structure Determination.

X-ray diffraction data were collected at Northeastern Collaborative Access Team beamline 24-ID-E, Advanced Photon Source, Argonne National Laboratory, from crystals of msAPAH complexed with 1 and 3. X-ray diffraction data were collected at Northeastern Collaborative Access Team beamline 24-ID-C, Advanced Photon Source, Argonne National Laboratory, from crystals of msAPAH complexed with 2, 5, and 7. X-ray diffraction data were collected at beamline 12–2, Stanford Synchrotron Radiation Laboratory, Stanford University, from crystals of msAPAH complexed with 4 and 6. X-ray diffraction data were collected at Frontier Macromolecular Crystallography beamline 17-ID-2 (FMX), National Synchrotron Light Source II, Brookhaven National Laboratory, from crystals of msAPAH complexed with acetate. Data were indexed and integrated using iMosflm42 and scaled using Aimless in the CCP4 program suite.43 Data reduction statistics are recorded in Table 1.

Table 1:

Crystallographic data collection and refinement statistics.

Data Collection msAPAH-1 msAPAH-2 msAPAH-3 msAPAH-4 msAPAH-5 msAPAH-6 msAPAH-7
msAPAH-Acetate
Space Group P21 P21 P21 P1 P21 P21 P21 P21
Unit Cell Dimensionsa
a,b,c (Å) 65.6 163.5 65.7 47.4 82.4 90.7 45.9 120.7 65.4 66.3 67.0 167.5 52.3 119.1 66.2 65.7 163.2 65.7 65.6 163.0 65.6 52.4 121.1 66.6
α,β,γ (°) 90 94 90 90 98 90 90 109 90 90 90 93 90 109 90 90 94 90 90 95 90 90 109 90
Rmergeb 0.161 (0.456) 0.180 (0.597) 0.163 (1.018) 0.205 (0.474) 0.172 (0.742) 0.092 (0.531) 0.105 (0.370) 0.104 (0.826)
Rpimc 0.102 (0.312) 0.180 (0.595) 0.106 (0.675) 0.158 (0.403) 0.172 (0.742) 0.085 (0.470) 0.100 (0.353) 0.066 (0.531)
CC1/2d 0.958 (0.706) 0.952 (0.525) 0.991 (0.657) 0.830 (0.588) 0.969 (0.537) 0.992 (0.657) 0.986 (0.811) 0.997 (0.703)
Redundancy 3.6 (3.6) 3.4 (3.5) 6.2 (6.1) 1.9 (1.9) 3.4 (3.4) 3.2 (3.0) 3.4 (3.1) 6.7 (6.6)
Completeness (%) 98.9 (99.7) 90.2 (91.4) 99.1 (99.2) 83.2 (86.5) 86.7 (88.7) 95.0 (95.6) 90.8 (81.8) 99.9 (99.7)
I/σ 4.7 (2.8) 10.2 (6.1) 6.7 (2.1) 4.4 (2.2) 5.6 (2.1) 7.3 (1.9) 6.8 (2.5) 10.2 (2.1)
Refinement
Resolution (Å) 65.44–1.65 (1.68–1.65) 89.79–2.00 (2.07–2.00) 61.77–1.65 (1.71–1.65) 66.97–2.03 (2.10–2.03) 59.53–1.54 (1.60–1.55) 46.27–1.75 (1.81–1.75) 60.71–1.70 (1.76–1.70) 27.85–1.64 (1.69–1.64)
No. Reflections 125734
(12518)
41498
(4163)
79658
(7965)
154543
(15944)
95821
(7699)
131376
(12851)
136205
(13035)
95961
(9560)
Rwork/Rfreee 0.193/0.225 (0.216/0.264) 0.161/0.212 (0.158/0.227) 0.179/0.208 (0.254/0.284) 0.220/0.247 (0.278/0.315) 0.244/0.280 (0.415/0.415) 0.175/0.212 (0.242/0.293) 0.162/0.199 (0.202/0.247) 0.168/0.194 (0.277/0.333)
No. Atomsf
Protein 10580 5332 5325 21126 5343 10689 10673 5386
Ligand 69 44 54 112 33 85 105 25
Solvent 598 354 294 648 254 774 953 502
Average B Factors Å2
Protein 12 11 15 10 12 17 13 19
Ligand 15 21 19 8 16 20 24 19
Solvent 15 14 19 9 14 21 20 26
R.m.s. Deviations
Bond Lengths (Å) 0.007 0.007 0.008 0.004 0.006 0.006 0.006 0.006
Bond Angles (°) 0.9 0.9 0.9 0.7 0.8 0.8 0.8 0.9
Ramachandran Plotg
Favored 96.0 97.0 96.0 97.0 97.0 96.0 96.0 97.0
Allowed 4.00 3.00 4.00 3.00 3.00 4.00 4.00 3.00
Outliers 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00
PDB code 6PHT 6PHZ 6PHR 6PIC 6PID 6PIA 6PI1 6PI8
a

Values in parentheses refer to the highest-resolution shell indicated.

b

Rmerge = ∑hkli|Ii,hkl − ⟨I⟩hkl|/∑hkliIi,hkl, where ⟨I⟩hkl is the average intensity calculated for reflection hkl from replicate measurements.

c

Rp.i.m.= (∑hkl(1/(N-1))1/2i|Ii,hkl − ⟨I⟩hkl|)/∑hkli Ii,hkl, where ⟨I⟩hkl is the average intensity calculated for reflection hkl from replicate measurements and N is the number of reflections.

d

Pearson correlation coefficient between random half-datasets.

e

Rwork = ∑||Fo| − |Fc||/∑|Fo| for reflections contained in the working set. |Fo| and |Fc| are the observed and calculated structure factor amplitudes, respectively. Rfree is calculated using the same expression for reflections contained in the test set held aside during refinement.

f

Per asymmetric unit.

g

Calculated with PROCHECK.

Molecular replacement using the atomic coordinates of mrAPAH [Protein Data Bank (PDB) entry 4ZUM]18 was used as a search model to phase the initial electron density map of the first msAPAH structure. Subsequent msAPAH structures were determined in similar fashion using either mrAPAH or msAPAH as search probes for molecular replacement. Autobuild44 was used to register the msAPAH sequence in the initial electron density map. The graphics program COOT45 was used to manually adjust residue conformations and structure refinement was performed using PHENIX.46 The quality of each structure was assessed using MolProbity47 and PROCHECK.48 Refinement statistics are recorded in Table 1.

Results

Catalytic activity.

Recombinant msAPAH exhibits broad specificity for short and long acetylpolyamine substrates, with a moderate preference for putrescine or N8-acetylspermidine (depending on the concentration of KCl) based on catalytic efficiency (kcat/KM) (Figure 2, Table 2). Even so, catalytic efficiency is modest; at best, kcat/KM = 470 M−1s−1 for N8-acetylspermidine in the presence of 10 mM KCl. We were unable to determine steady-state kinetic parameters in higher concentrations of KCl. No significant catalytic activity was observed for the hydrolysis of Aly-NH2 or the MSH2-based peptide substrates Ac-Ala-Aly, Ac-Gly-Ala-Aly, Ac-Gly-Ala-Aly-Asn-Leu-Gln-NH2, or Ac-Aly-Asn-Leu-NH2 (Aly = acetyllysine) using the liquid chromatography-mass spectrometry assay employed for the assay of HDAC6 activity.16 Thus, while msAPAH is not a particularly efficient polyamine deacetylase, it is clearly not a lysine deacetylase.

Figure 2.

Figure 2.

Michaelis-Menten plot showing the steady-state kinetics of msAPAH-catalyzed acetylpolyamine hydrolysis.

Table 2:

Steady-state kinetic parameters for acetylpolyamine hydrolysis by msAPAH

Substrate kcat
(s−1)
KM
(μM)
kcat/KM
(M−1s−1)
graphic file with name nihms-1047609-t0014.jpg
acetylcadaverine
0.026 ± 0.004 800 ± 300 32 ± 6
graphic file with name nihms-1047609-t0015.jpg
acetylputrescine
0.0078 ± 0.0003 44 ± 10 180 ± 30
graphic file with name nihms-1047609-t0016.jpg
acetylspermine
0.007 ± 0.002 1300 ± 600 5 ± 1
graphic file with name nihms-1047609-t0017.jpg
N1-acetylspermidine
0.0038 ± 0.0006 410 ± 200 9 ± 4
graphic file with name nihms-1047609-t0018.jpg
N8-acetylspermidine
(0 mM KCl and 10 mM KCl)
0.0092 ± 0.0009

0.008 ± 0.003
280 ± 80

17 ± 5
32 ± 20

470 ± 40

All measurements represent mean ± average of absolute deviations from the mean in a given set of data. All measurements were run in triplicate at 25° C.

Structure of the msAPAH–1 Complex.

The 1.65 Å-resolution crystal structure of msAPAH complexed with the boronic acid analogue of N8-acetylspermidine (1, IC50 = 390 μM) contains 4 monomers in the asymmetric unit. The structure of msAPAH reveals the arginase-deacetylase fold, as previously observed for the broad-specificity polyamine deacetylase mrAPAH.17 Like mrAPAH, msAPAH is a dimer with identical quaternary structure (Figure 3). There are no major conformational changes between the monomers or dimers of msAPAH and mrAPAH (PDB 4ZUM): the root-mean-square (rms) deviation is 0.53 Å for 268 Cα atoms (monomer A) and 0.75 Å for 566 Cα atoms for the dimer, respectively.

Figure 3.

Figure 3.

Crystal structure of the msAPAH-1 complex (top) illustrating the dimer interface between monomer A (blue) and monomer B (red), looking down the two-fold symmetry axis. There are no major conformational differences between monomers and least-squares superposition yields an rms deviation of 0.15 Å for 300 Cα atoms. The cutaway view of the active site (bottom) shows how the dimer interface constricts the approach to the catalytic Zn2+ ions (gray spheres). Only long and slender acetylpolyamine substrates, and not bulky acetyllysine-containing peptide substrates, can bind in this constricted active site.

Electron density corresponding to the boronic acid moiety of 1 clearly indicates trigonal planar geometry (Figure 4), with the hydroxyl groups coordinated to the catalytic Zn2+ ion with asymmetric bidentate coordination geometry (average Zn2+---O1 distance = 2.1 Å and Zn2+---O2 distance = 2.4 Å). Notably, inhibitor binding displaces the Zn2+-bound solvent molecule expected for the unliganded enzyme. Surprisingly, the electrophilic boronic acid does not undergo nucleophilic attack to form a tetrahedral boronate anion, which would mimic the tetrahedral transition state for acetylpolyamine hydrolysis. Such chemistry is typically observed, for example, in the binding of boronic acids in the arginase active site,4952 but not exclusively so.53 Boronic acid hydroxyl group O1 hydrogen bonds with Y323, and hydroxyl group O2 hydrogen bonds with H158 and H159. These catalytic residues are unique to metal-dependent deacetylases. In catalysis, Y323 assists the Zn2+ ion in polarizing the amide carbonyl group of the substrate for nucleophilic attack by a Zn2+-bound solvent molecule, and tandem histidine residues H158-H159 serve general base-general acid functions.

Figure 4.

Figure 4.

Polder omit map of the msAPAH-1 complex (contoured at 3.0σ). Atoms are color-coded as follows: C = light blue (msAPAH monomer D), dark gray (msAPAH monomer B) or wheat (inhibitor), B = light green, N = blue, O = red, Zn2+ = gray sphere, Mg2+ = large, dark red sphere, and solvent = small red spheres. Metal coordination and hydrogen bond interactions are indicated by solid and dashed black lines, respectively.

Additional hydrogen bond interactions stabilize the binding of 1 in the msAPAH active site. The primary amino group of 1 forms water-mediated hydrogen bonds with the side chain of E17, the backbone carbonyl of L18, and the side chain of D19. The binding conformation of 1 is slightly different in monomer B, such that there is a direct hydrogen bond with the side chain of Y168. The secondary amino group of 1 forms water-mediated hydrogen bonds with the side chains of D117, Y168, and the primary amino group of 1.

Although the quaternary structure of msAPAH is identical to that of mrAPAH, there are several key differences in residues that define the active site cleft at the dimer interface of msAPAH. Residue Y19 in mrAPAH corresponds to D19 in msAPAH, a residue that is critical for accepting hydrogen bonds from the polyamine substrates. Residue F27 in mrAPAH corresponds to H27 in msAPAH and is located at the surface of the active site pocket. The most noticeable difference at the dimer interface is the substitution of residues Y83, S91 and E106 in mrAPAH by K83, F91 and D106 in msAPAH. This difference enables the binding of a Mg2+ ion at the dimer interface of msAPAH. This Mg2+ ion is observed in all msAPAH crystal structures described herein and is coordinated by the carboxylate side chains of D104 and D106; the Mg2+ coordination polyhedron is completed by water molecules. Additional residues important for dimer assembly of mrAPAH include F92, which engages in offset π stacking interactions with F92 in the adjacent monomer. In msAPAH, this interaction is maintained by W92-W92 at the dimer interface.

In addition to the Mg2+ ion bound in the active site at the dimer interface, an additional Mg2+ ion is coordinated by E288 in monomer A. The coordination sphere of this Mg2+ ion is completed by 4 water molecules, yielding a pentacoordinate metal ion. This Mg2+ ion is absent in monomers B, C, and D.

Structure of the msAPAH–2 Complex.

The 2.00 Å-resolution crystal structure of msAPAH complexed with trifluoroketone analogue of N8-acetylspermidine (2, IC50 = 350 μM) contains 2 monomers in the asymmetric unit. The trifluoroketone moiety exists predominantly as the gem-diol in solution;54 accordingly, it binds as the gem-diol or perhaps the ionized gem-diolate in the active site of msAPAH (Figure 5). This binding mode mimics the binding of the tetrahedral intermediate and its flanking transition states in the hydrolysis of an acetylpolyamine substrate. The gem-diol(ate) coordinates to the catalytic Zn2+ ion with nearly symmetric bidentate coordination geometry, with average Zn2+---O1 and Zn2+---O2 distances of 2.2 Å and 2.3 Å, respectively. The gem-diol O1 group also hydrogen bonds with Y323, which adopts the “in” conformation in monomer B. This hydrogen bond is absent in monomer A, where Y323 adopts the “out” conformation. The gem-diol O2 group hydrogen bonds with H158 and H159. The binding of trifluoroketone inhibitors as gem-diols has similarly been observed in the active sites of other deacetylases.3,16,18

Figure 5.

Figure 5.

Polder omit map of the msAPAH-2 complex, monomer B (contoured at 3.0σ). Atoms are color-coded as follows: C = light blue (msAPAH monomer B), dark gray (msAPAH monomer A) or wheat (inhibitor), F = magenta, S = yellow, N = blue, O = red, Zn2+ = gray sphere, Mg2+ = large, dark red sphere, and solvent = small red spheres. Metal coordination and hydrogen bond interactions are indicated by solid and dashed black lines, respectively.

Each of the C-F bonds of the trifluoro moiety of 2 engages in hydrogen bond interactions with protein atoms. One fluorine atom accepts a hydrogen bond from Y323, and a second fluorine atom accepts a hydrogen bond from the backbone NH group of G321. A third fluorine atom accepts a hydrogen bond from C169; this fluorine atom also makes a van der Waals contact with the backbone carbonyl of G167.

The primary amino group of 2 donates hydrogen bonds to the side chain of E17, the side chain of D19, and the backbone carbonyl oxygen of L18. The secondary amino group makes water-mediated hydrogen bond interactions with the side chains of D117 (monomer B) or Y168 (monomer A).

In addition to the Mg2+ ion bound in the active site, magnesium hexahydrate (Mg2+(OH2)6) binds near D275 in monomer A, and the side chain carboxylate of D275 makes hydrogen bond interactions with metal-bound water molecules. In monomer B, E288 coordinates to a Mg2+ ion along with 5 water molecules, yielding octahedral metal coordination geometry.

Structure of the msAPAH-Acetate Complex.

The 1.64 Å-resolution crystal structure of msAPAH complexed with the acetate coproduct contains 2 monomers in the asymmetric unit. Absent the polyamine coproduct bound in the active site, this structure represents a partially dissociated product complex. Acetate coordinates to the active site Zn2+ ion with nearly symmetric bidentate geometry (average Zn2+---O1 and Zn2+---O2 distances = 2.2 Å and 2.4 Å, respectively) (Figure 6). The carboxylate O1 atom additionally accepts a hydrogen bond from Y323, and the carboxylate O2 atom accepts hydrogen bonds from H158 and H159.

Figure 6.

Figure 6.

Polder omit map of the msAPAH-acetate complex (contoured at 3.0σ). Atoms are color-coded as follows: C = light blue (msAPAH monomer A), dark gray (msAPAH monomer B), wheat (inhibitor), or orange (ethylene glycol), N = blue, O = red, Zn2+ = gray sphere, Mg2+ = large, dark red sphere, and solvent = small red spheres. Metal coordination and hydrogen bond interactions are indicated by solid and dashed black lines, respectively.

In addition to the Mg2+ ion bound in the active site, an additional Mg2+ ion is coordinated by E288, D324, D326, and 3 water molecules in monomer A. However, this Mg2+ ion is absent in monomer B.

Structure of the msAPAH–3 Complex.

The 1.65 Å-resolution crystal structure of msAPAH complexed with the thiol analogue of N8-acetylspermidine (3, IC50 = 160 μM, Kd = 0.2 μM) contains 2 monomers in the asymmetric unit. The thiol group is presumably ionized to the negatively charged thiolate anion as it coordinates to the active site Zn2+ ion, with an average Zn2+---S distance of 2.4 Å (Figure 7). With Zn2+---S–C angles of 116° and 100°, and Zn2+---S–C–C dihedral angles of −7° in monomer A and 59° in monomer B, respectively, thiolate-metal coordination parameters deviate somewhat from ideal values for thiolate-metal interactions first outlined for the side chain of cysteine (Zn2+---S–C angle of 90° and Zn2+---S–C–C dihedral angle of ± 90° or 180°).55 The phenolic hydroxyl group of Y323 additionally donates a hydrogen bond to the Zn2+-bound thiolate.

Figure 7.

Figure 7.

Polder omit map of the msAPAH-3 complex (contoured at 3.0σ). Atoms are color-coded as follows: C = light blue (msAPAH, monomer B), dark gray (msAPAH, monomer A), or wheat (inhibitor and MES), S = yellow, N = blue, O = red, Zn2+ = gray sphere, Mg2+ = large, dark red sphere, and solvent = small red spheres. Metal coordination and hydrogen bond interactions are indicated by solid and dashed black lines, respectively.

As observed in the msAPAH-2 complex, the primary amino group of 3 hydrogen bonds to E17, D19, and the backbone carbonyl of L18. Also similar are the hydrogen bonds between the secondary amino group of 3 and D117 and Y168.

Unique to this structure is the binding of a buffer molecule, 2-(N-morpholino)-ethanesulfonic acid (MES), in the active site tunnel (Figure 7). The sulfonate group of MES accepts a hydrogen bond from the backbone NH group of F225 and a Mg2+-bound water molecule. It also makes a hydrogen bond with H197.

Structure of the msAPAH–4 Complex.

The 2.03 Å-resolution crystal structure of msAPAH complexed with hydroxamate inhibitor 4 (IC50 = 410 μM) contains 8 monomers in the asymmetric unit. Inhibitor 4 is a hydroxamate analogue of N-acetylputrescine. Intermolecular interactions of the hydroxamate group provide useful structural inferences on catalysis (Figure 8). Specifically, just as the hydroxamate carbonyl group coordinates to Zn2+ and accepts a hydrogen bond from Y323, so too would the scissile carbonyl group of N-acetylputrescine in the precatalytic enzyme-substrate complex. Both Zn2+ coordination and Y323 hydrogen bond interactions are required to activate the amide carbonyl of N-acetylputrescine for nucleophilic attack by a Zn2+-bound solvent molecule in catalysis.

Figure 8.

Figure 8.

Polder omit map of the msAPAH-4 complex (contoured at 3.0σ). Atoms are color-coded as follows: C = light blue (msAPAH, monomer A), dark gray (msAPAH, monomer C, symmetry mate), or wheat (inhibitor), N = blue, O = red, Zn2+ = gray sphere, Mg2+ = large, dark red sphere, and solvent = small red spheres. Metal coordination and hydrogen bond interactions are indicated by solid and dashed black lines, respectively.

The primary amino group of inhibitor 4 donates hydrogen bonds to D117 and Y168 (Figure 8), in a manner similar to the interactions of the secondary amino group of inhibitor 3 (Figure 7). The primary amino group of 4 also engages in water-mediated hydrogen bonds with E17, D19, and the backbone carbonyl of L18.

Structure of the msAPAH–5 Complex.

The 1.55 Å-resolution crystal structure of msAPAH complexed with hydroxamate inhibitor 5 (IC50 = 380 μM, Kd = 29 μM) contains 2 monomers in the asymmetric unit. Hydroxamate-metal coordination interactions and hydrogen bond interactions (Figure 9) are identical to those observed for the binding of hydroxamate inhibitor 4. While inhibitor 5 is not isosteric with any naturally occurring polyamine substrate, its intermolecular interactions reveal that alternative hydrogen bond interactions can be achieved for the primary amino group of a long, slender polyamine analogue in the active site of msAPAH. Inhibitor 5 is two methylene groups longer than inhibitor 4 and accordingly extends further out of the active site tunnel. The additional length allows for the primary amino group to engage in direct hydrogen bonds rather than water-mediated hydrogen bonds with E17, D19, and the backbone carbonyl of L18. The D19 hydrogen bond only occurs in monomer A while a hydrogen bond to Y168 occurs in monomer B.

Figure 9.

Figure 9.

Polder omit map of the msAPAH-5 complex (contoured at 3.0σ). Atoms are color-coded as follows: C = light blue (msAPAH, monomer A), dark gray (msAPAH, monomer B), or wheat (inhibitor), N = blue, O = red, Zn2+ = gray sphere, Mg2+ = large, dark red sphere, and solvent = small red spheres. Metal coordination and hydrogen bond interactions are indicated by solid and dashed black lines, respectively.

In addition to the Mg2+ ion bound in the active site, an additional Mg2+ ion is coordinated by E288, D324, D326, and 3 water molecules in monomer A. However, this Mg2+ ion is absent in monomer B.

Structure of the msAPAH–6 Complex.

The 1.75 Å-resolution crystal structure of the complex between msAPAH and hydroxamate inhibitor 6 (IC50 = 470 μM, Kd = 11 μM) contains 4 monomers in the asymmetric unit. Hydroxamate-metal coordination interactions and hydrogen bond interactions (Figure 10) are identical to those observed for other hydroxamate inhibitors described above. Inhibitor 6 is the hydroxamate analogue of the polyamine substrate N8-acetylspermidine. The primary amino group of 6 donates hydrogen bonds to E17, D19, and the backbone carbonyl of L18. The secondary amino group of 6 donates hydrogen bonds to D117 in monomers A and B, and Y168 in all four monomers.

Figure 10.

Figure 10.

Polder omit map of the msAPAH-6 complex (contoured at 3.0σ). Atoms are color-coded as follows: C = light blue (msAPAH, monomer A), dark gray (msAPAH, monomer B), or wheat (inhibitor), N = blue, O = red, Zn2+ = gray sphere, Mg2+ = large, dark red sphere, and solvent = small red spheres. Metal coordination and hydrogen bond interactions are indicated by solid and dashed black lines, respectively.

In addition to the Mg2+ ion bound in the active site, an additional Mg2+ ion is coordinated by E288 and 5 water molecules in monomer B; D324 and D326 form hydrogen bonds with metal-bound water molecules. However, this Mg2+ ion is absent in monomers A, C, and D.

Structure of the msAPAH–7 Complex.

The 1.70 Å-resolution crystal structure of msAPAH complexed with inhibitor 7 (IC50 = 160 μM) contains 4 monomers in the asymmetric unit. Inhibitor 7 is also known as M344 and is characterized by a m-dimethylaminobenzamide “capping group”. Hydroxamate-metal coordination interactions and hydrogen bond interactions (Figure 11) are identical to those observed for other hydroxamate inhibitors described above. The aromatic capping group is nestled in the dimer interface and is oriented away from the Mg2+ ion. The amide group forms water-mediated hydrogen bonds with H197, the backbone amide of F225, the backbone carbonyl of I291, and a Mg2+-bound water molecule. The amide group also forms water-mediated hydrogen bonds with E17, D117, and Y168.

Figure 11.

Figure 11.

Polder omit map of the msAPAH-7 complex (contoured at 3.0σ). Atoms are color-coded as follows: C = light blue (msAPAH, monomer B), dark gray (msAPAH, monomer A), or wheat (inhibitor), N = blue, O = red, Zn2+ = gray sphere, Mg2+ = large, dark red sphere, and solvent = small red spheres. Metal coordination and hydrogen bond interactions are indicated by solid and dashed black lines, respectively.

In addition to the Mg2+ ion bound in the active site, an additional Mg2+ ion is coordinated by E288 and 5 water molecules in monomer B. However, this Mg2+ ion is absent in monomers A, C, and D.

Discussion

Crystal structures of msAPAH–inhibitor complexes reveal key insight on substrate recognition and catalysis by a bacterial polyamine deacetylase. Common to all metal-dependent deacetylases in the arginase-deacetylase superfamily, catalytically important residues in the msAPAH active site include tandem histidine residues H158 and H159. We speculate that H158 serves as a general base in catalysis by deprotonating a Zn2+-bound water molecule for nucleophilic attack at the scissile amide carbonyl of the substrate, which in turn is polarized by coordination to Zn2+ and also by a hydrogen bond with Y323. This mechanistic proposal is consistent with intermolecular interactions observed for the binding of trifluoroketone transition state analogue 2 (Figure 5) as well as the acetate coproduct (Figure 6). Based on the observation of two conformations for Y323 in the msAPAH–2 complex – the “in” conformation oriented into the active site, and the “out” conformation oriented toward bulk solvent – it appears likely that Y323 undergoes a substrate-induced conformational change in catalysis. Conformational flexibility of the catalytic tyrosine has also been observed in crystal structures of mrAPAH,17,18 as well as molecular dynamics simulations of HDAC3 and HDAC8.5658 This flexibility appears to be enhanced by the location of Y323 in a glycine-rich loop.58

Molecular recognition of cationic polyamine substrates by msAPAH and mrAPAH is facilitated by substantial negative electrostatic potential on the protein surface surrounding the mouth of the active site (Figure 12). The negatively charged protein surface provides electrostatic attraction to facilitate the binding of positively charged polyamine substrates. Once bound in the active site of msAPAH, the backbone carbonyl of L18 and the side chains of E17, D19, D117 and Y168 provide direct or water-mediated hydrogen bond interactions to the positively charged amino groups of the substrate, as observed in the array of crystal structures presented in Figures 411.

Figure 12.

Figure 12.

Electrostatic surfaces of msAPAH, mrAPAH, and zHDAC10. Arrows indicate the mouth of each enzyme active site.

Unique to msAPAH is the binding of a Mg2+ ion in the substrate binding cleft in the structures of all enzyme-inhibitor complexes. This metal ion is coordinated by D104 and D106 of the alternate subunit of the dimer, and water molecules usually complete an octahedral coordination complex. The binding of this Mg2+ ion further constricts the active site and likely contributes to the catalytic preference for acetylpolyamine substrates. The Mg2+ ligand D106 is not conserved in mrAPAH and Mg2+ binding has not been observed in this related bacterial deacetylase. The binding of additional Mg2+ ions as observed in all msAPAH structures may reflect the functional adaptation of this deacetylase as M. subterrani evolved in the high-salt subterranean environment.

A characteristic feature of halophilic proteins is an increase in the number of acidic residues on the protein surface.5962 Accordingly, msAPAH has increased aspartate and glutamate content (15% total) compared with mrAPAH (12% total). Increased negative charged accounts for the lower pI of msAPAH compared with that of mrAPAH: based on pI values calculated from their amino acid sequences, the pI of msAPAH is 5.0 and the pI of mrAPAH is 5.3. Additional aspartate residues present in msAPAH also account for additional Mg2+ binding sites observed in the crystal structures of complexes with different inhibitors.

The structural features described above likely contribute to the substrate specificity of mrAPAH and msAPAH. Our previous steady-state kinetic studies utilizing a liquid chromatography-mass spectrometry assay for the direct measurement of dansylated polyamine products3 show that mrAPAH exhibits broad substrate specificity and catalyzes the hydrolysis of N-acetylcadaverine, N-acetylputrescine, N1-acetylspermidine, N8-acetylspermidine, and N-acetylspermine with turnover numbers ranging 0.20–2.5 s−1 and catalytic efficiencies ranging from 2.5 × 103 M−1s−1 to 6.8 × 104 M−1s−1. These results are consistent with previous observations of broad substrate specificity based on specific activity measurements using native and recombinant mrAPAH.63,64 Similarly, msAPAH exhibits broad substrate specificity but lower catalytic activity using an assay that quantitates the generation of coproduct acetate: turnover numbers range 0.0038–0.026 s−1 and catalytic efficiencies range from 5 M−1s−1 to 470 M−1s−1. We cannot rule out the possibility that msAPAH assay conditions do not fully reflect the cytosolic conditions of the deep earth halophile from which msAPAH derives, nor can we rule out the possibility that the biological substrate of msAPAH in vivo is an as-yet unidentified acetylated small molecule. The biological substrate is unlikely to be acetyllysine, since a peptide or protein substrate would not fit in the constricted approach to the active site (Figure 3).

Recently, eukaryotic HDAC10 was discovered to be a highly specific cytosolic polyamine deacetylase that utilizes N8-acetylspermidine as a substrate.3 Although this activity was first observed in partially purified cell extracts more than 40 years ago, the responsible enzyme had not been identified.35,36 It is interesting to contrast the structural basis of catalytic specificity in the bacterial polyamine deacetylases msAPAH and mrAPAH with that of HDAC10.34 This specificity is rooted in structural and electrostatic features in each enzyme active site. First, each enzyme active site is sterically constricted so as to favor the binding of long, slender polyamine substrates over bulky, acetyllysine-containing peptide substrates. However, the active sites of the bacterial enzymes are constricted by quaternary structure (Figure 3), whereas the active site of the eukaryotic enzyme is constricted by tertiary structure: a 310 helix unique to HDAC10 and no other HDAC isozymes.3,34 Second, each enzyme is characterized by significant negative electrostatic potential on the protein surface flanking the active site (Figure 12), as well as specific aspartate and/or glutamate residues in the active site cleft that engage in electrostatic interactions with the positively charged amino groups of polyamine substrates. As noted by Hai and colleagues,3 polyamine substrate specificity in HDAC10 and APAH enzymes appears to have evolved convergently.

In closing, it is interesting to consider the biology and ecology of the halophile Marinobacter subterrani from which msAPAH derives. As noted in the introduction, these species are well adapted to the high Fe2+ concentrations found in deep subsurface locations.40 Given that HDAC enzymes can utilize Fe2+ for catalytic activity slightly exceeding that measured with Zn2+, and the implication that metal-dependent deacetylases could utilize Fe2+, Zn2+, or perhaps a mixture of both in vivo,23 it is interesting to speculate that msAPAH could potentially function as an Fe2+-dependent polyamine deacetylase in the iron-rich environment of the deep terrestrial biosphere. Future structure-function studies of msAPAH will shed further light on this possibility.

Supplementary Material

Supporting Information

Figure S1: IC50 plots; Figure S2: ITC enthalpograms.

Acknowledgments

We thank the Northeastern Collaborative Access Team (NE-CAT) beamlines, which are funded by the National Institute of General Medical Sciences from the National Institutes of Health (P30 GM124165). The Pilatus 6M detector on beamline 24-ID-C is funded by a NIH-ORIP HEI grant (S10 RR029205). This research used resources of the Advanced Photon Source, a U.S. Department of Energy (DOE) Office of Science User Facility operated for the DOE Office of Science by Argonne National Laboratory under Contract No. DE-AC02-06CH11357. Additionally, we thank the Stanford Synchrotron Radiation Lightsource (SSRL), SLAC National Accelerator Laboratory, which is supported by the DOE Office of Science, Office of Basic Energy Sciences, under Contract No. DE-AC02-76SF00515. The SSRL Structural Molecular Biology Program is supported by the DOE Office of Biological and Environmental Research, and by the NIH, National Institute of General Medical Sciences (NIGMS) (including P41GM103393). Finally, we thank the FMX beamline at the National Synchrotron Light Source II, a DOE Office of Science User Facility operated for the DOE Office of Science by Brookhaven National Laboratory under Contract No. DE-SC0012704. The Life Science Biomedical Technology Research resource is primarily supported by the National Institute of Health, National Institute of General Medicine Sciences (NIGMS) through a Biomedical Technology Research Resource P41 grant (P41GM111244), and by the DOE Office of Biological and Environmental Research (KP1605010).

Funding

We thank the National Institutes of Health for grants GM49758 (D.W.C.) and F32GM125141 (S.A.S.) in support of this research.

Footnotes

Supporting Information

The Supporting Information is available free of charge on the ACS Publications website.

Accession Codes

The atomic coordinates and crystallographic structure factors of msAPAH complexes with the inhibitors shown in Figure 1b have been deposited in the Protein Data Bank (www.rcsb.org) with accession codes as follows: 1, 6PHT; 2, 6PHZ; 3, 6PHR; 4, 6PIC; 5, 6PID; 6, 6PIA; 7, 6PI1; acetate, 6PI8.

The authors declare no competing financial interests.

References

  • 1.Ouzounis CA, and Kyrpides NC (1994) On the evolution of arginases and related enzymes. J. Mol. Evol 39, 101–104. [DOI] [PubMed] [Google Scholar]
  • 2.Liepe DD, and Landsman D (1997) Histone deacetylases, acetoin utilization proteins and acetylpolyamine amidohydrolases are members of an ancient protein superfamily. Nucleic Acids Res 25, 3693–3697. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.Hai Y, Shinsky SA, Porter NJ, and Christianson DW (2017) Histone deacetylase 10 structure and molecular function as a polyamine deacetylase. Nat. Commun 8, 15368. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Kanyo ZF, Scolnick LR, Ash DE, and Christianson DW (1996) Structure of a unique binuclear manganese cluster in arginase. Nature 383, 554–557. [DOI] [PubMed] [Google Scholar]
  • 5.Finnin MS, Donigina JR, Cohen A, Richon VM, Rifkind RA, Marks PA, Breslow R, and Pavletich NP (1996) Structures of a histone deacetylase homologue bound to the TSA and SAHA inhibitors. Nature 401, 188–193. [DOI] [PubMed] [Google Scholar]
  • 6.Elkins JM, Clifton IJ, Hernández H, Doan LX, Robinson CV, Schofield CJ, and Hewitson KS (2002) Oligomeric structure of proclavaminic acid amidino hydrolase: evolution of a hydrolytic enzyme in clavulanic acid biosynthesis. Biochem. J 366, 423–434. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.Ahn HJ, Kim KH, Lee J, Ha J-Y, Lee HH, Kim D, Yoon H-J, Kwon A-R, and Suh SW (2004) Crystal structure of agmatinase reveals structural conservation and inhibition mechanism of the ureohydrolase superfamily. J. Biol. Chem 279, 50505–50513. [DOI] [PubMed] [Google Scholar]
  • 8.Hai Y, Dugery RJ, Healy D, and Christianson DW (2013) Formiminoglutamase from Trypanosoma cruzi is an arginase-like manganese metalloenzyme. Biochemistry 52, 9294–9309. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Somoza JR, Skene RJ, Katz BA, Mol C, Ho JD Jennings AJ, Luong C, Arvai A, Buggy JJ, Chi E, Tang J, Sang BC, Verner E, Wynands R, Leahy EM, Dougan DR, Snell G, Navre M, Knuth MW, Swanson RV, McRee DE, and Tari LW (2004) Structural snapshots of human HDAC8 provide insights into the class I histone deacetylases. Structure 12, 1325–1334. [DOI] [PubMed] [Google Scholar]
  • 10.Vannini A, Volpari C, Filocamo G, Casavola EC, Brunetti M, Renzoni D, Chakravarty P, Paolini C, De Francesco R, Gallinari P, Steinkuhler C, and Di Marco S (2004) Crystal structure of a eukaryotic zinc-dependent histone deacetylase, human HDAC8, complexed with a hydroxamic acid inhibitor. Proc. Natl. Acad. Sci. USA 101, 15064–15069. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Bottomley MJ, Lo Surdo P, Di Giovine P, Cirillo A, Scarpelli R, Ferrigno F, Jones P, Neddermann P, De Francesco R, Steinkuhler C, Gallinari P, and Carfi A (2008) Structural and functional analysis of the human Hdac4 catalytic domain reveals a regulatory zinc-binding domain. J. Biol. Chem 283, 26694–26704. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Schuetz A, Min J, Allali-Hassani A, Schapira M, Shuen M, Loppnau P, Mazitschek R, Kwiatkowski NP, Lewis TA, Maglathin RL, McLean TH, Bochkarev A, Plotnikov AN, Vedadi M, and Arrowsmith CH (2008) Human HDAC7 harbors a class IIa histone deacetylase-specific zinc binding motif and cryptic deacetylase activity. J. Biol. Chem 283, 11355–11363. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Bressi JC, Jennings AJ, Skene R, Wu Y, Melkus R, De Jong R, O’Connell S, Grimshaw CE, Navre M, and Gangloff AR (2010) Exploration of the HDAC2 foot pocket: Synthesis and SAR of substituted N-(2-aminophenyl)Benzamides. Bioorg. Med. Chem. Lett 20, 3142–3145. [DOI] [PubMed] [Google Scholar]
  • 14.Watson PJ, Fairall L, Santos GM, and Schwabe JWR (2012) Structure of Hdac3 bound to co-repressor and inositol tetraphosphate. Nature 481, 335–340. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Millard CJ, Watson PJ, Celardo I, Gordiyenko Y, Cowley SM, Robinson CV, Fairall L and Schwabe JW (2013) Class I HDACs share a common mechanism of regulation by inositol phosphates. Mol Cell 51, 57–67. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Hai Y, and Christianson DW (2016) Histone deacetylase 6 structure and molecular basis of catalysis and inhibition. Nat. Chem. Biol 12, 741–747. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Lombardi PM, Angell HD, Whittington DA, Flynn EF, Rajashankar KR, and Christianson DW (2011) Structure of prokaryotic polyamine deacetylase reveals evolutionary functional relationships with eukaryotic histone deacetylases. Biochemistry 50, 1808–1817. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Decroos C, and Christianson DW (2015) Design, synthesis, and evaluation of polyamine deacetylase inhibitors, and high-resolution crystal structures of their complexes with acetylpolyamine amidohydrolase. Biochemistry 54, 4692–4703. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Christianson DW (2005) Arginase: structure, mechanism, and physiological role in male and female sexual arousal. Acc. Chem. Res 38, 191–201. [DOI] [PubMed] [Google Scholar]
  • 20.Lombardi PM, Cole KE, Dowling DP, and Christianson DW (2011) Structure, mechanism, and inhibition of histone deacetylases and related metalloenzymes. Curr. Opin. Struct. Biol 21, 735–743. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Porter NJ, and Christianson DW (2019) Structure, mechanism, and inhibition of the zinc-dependent histone deacetylases. Curr. Opin. Struct. Biol 59, 9–18. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Ash DE (2004) Structure and function of arginases. J. Nutrition 134, 2760S–2764S. [DOI] [PubMed] [Google Scholar]
  • 23.Gantt SL, Gattis SG, and Fierke CA (2006) Catalytic activity and inhibition of human histone deacetylase 8 is dependent on the identity of the active site metal ion. Biochemistry 45, 6170–6178. [DOI] [PubMed] [Google Scholar]
  • 24.Tabor CW, and Tabor H (1984) Polyamines. Annu. Rev. Biochem 53, 749–790. [DOI] [PubMed] [Google Scholar]
  • 25.Thomas T, and Thomas TJ (2001) Polyamines in cell growth and cell death: molecular mechanisms and therapeutic applications. Cell Mol. Life Sci 58, 244–258. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Bachmann AS, and Geerts D (2018) Polyamine synthesis as a target of MYC oncogenes. J. Biol. Chem 293, 18757–18769. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Casero RA Jr., Murray Stewart T, and Pegg AE (2018) Polyamine metabolism and cancer: treatments, challenges and opportunities. Nat. Rev. Cancer 18, 681–695. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Bloomfield VA (1997) DNA condensation by multivalent cations. Biopolymers 44, 269–282. [DOI] [PubMed] [Google Scholar]
  • 29.Thomas TJ, and Thomas T (2017) Collapse of DNA in packaging and cellular transport. Int. J. Biol. Macromol 109, 36–48. [DOI] [PubMed] [Google Scholar]
  • 30.Trachman RJ III, and Draper DE (2017) Divalent ion competition reveals reorganization of an RNA ion atmosphere upon folding. Nucleic Acids Res 45, 4733–4742. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Childs AC, Mehta DJ, and Gerner EW (2003) Polyamine-dependent gene expression. Cell Mol. Life Sci 60, 1394–1406. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Ivanov IP, Atkins JF, and Michael AJ (2010) A profusion of upstream open reading frame mechanisms in polyamine-responsive translational regulation. Nucleic Acids Res 38, 353–359. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Dever TE, and Ivanov IP (2018) Roles of polyamines in translation. J. Biol. Chem 293, 18719–18729. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Shinsky SA, and Christianson DW (2018) Polyamine deacetylase structure and catalysis: prokaryotic acetylpolyamine amidohydrolase and eukaryotic HDAC10. Biochemistry 57, 3105–3114. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Blankenship J (1978) Deacetylation of N8-acetylspermidine by subcellular fractions of rat tissue. Arch. Biochem. Biophys 189, 20–27. [DOI] [PubMed] [Google Scholar]
  • 36.Libby PR (1978) Properties of an acetylspermidine deacetylase from rat liver. Arch. Biochem. Biophys 188, 360–363. [DOI] [PubMed] [Google Scholar]
  • 37.Handley KM, and Lloyd JR (2013) Biogeochemical implications of the ubiquitous colonization of marine habitats and redox gradients by Marinobacter species. Front. Microbiol 4, 136. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Fathepure BZ (2014) Recent studies in microbial degradation of petroleum hydrocarbons in hypersaline environments. Front. Microbiol 5, 173. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Cluff MA, Hartsock A, MacRae JD, Carter K, and Mouser PJ (2014) Temporal changes in microbial ecology and geochemistry in produced water from hydraulically fractured Marcellus shale gas wells. Environ. Sci. Technol 48, 6508–6517. [DOI] [PubMed] [Google Scholar]
  • 40.Bonis BM, and Gralnick JA (2015) Marinobacter subterrani, a genetically tractable neutrophilic Fe(II)-oxidizing strain isolated from the Soudan Iron Mine. Frontiers Microbiol 6, 719. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Decroos C, Bowman CM, and Christianson DW (2013) Synthesis and evaluation of N8-acetylspermidine analogues as inhibitors of bacterial acetylpolyamine amidohydrolase. Bioorg. Med. Chem 21, 4530–4540. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Battye TGG, Kontogiannis L, Johnson O, Powell HR, and Leslie AGW (2011) iMosflm: a new graphical interface for diffraction-image processing with Mosflm. Acta Crystallogr., Sect. D: Biol. Crystallogr 67, 271–281. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Winn MD, Ballard CC, Cowtan KD, Dodson EJ, Emsley P, Evans PR, Keegan RM, Krissinel EB, Leslie AG, W., McCoy A, McNicholas SJ, Murshudov GN, Pannu NS, Potterton EA, Powell HR, Read RJ, Vagin A, and Wilson KS (2011) Overview of the CCP4 suite and current developments. Acta Crystallogr., Sect D: Biol. Crystallogr 67, 235–242. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Terwillinger TC, Grosse-Kunnstleve RW, Afonine PV, Moriarty NM, Zwart PH, Hung LW, Read RJ, and Adams PD (2008) Iterative model building, structure refinement and density modification with the PHENIX AutoBuild wizard. Acta Crystallogr., Sect D: Biol. Crystallogr 64, 61–69. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Emsley P, Lohkamp B, Scott WG, and Cowtan K (2010) Features and development of Coot. Acta Crystallogr., Sect. D: Biol. Crystallogr 66, 486–501. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Adams PD, Afonine PV, Bunkóczi G, Chen VB, Davis IW, Echols N, Headd JJ, Hung L,Kapral GJ, Grosse-Kunstleve RW, McCoy AJ, Moriarty NW, Oeffner R, Read RJ,Richardson DC, Richardson JS, Terwilliger TC, and Zwart PH (2010) PHENIX: a comprehensive Python-based system for macromolecular structure solution. Acta Crystallogr., Sect. D: Biol. Crystallogr 66,213–221. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Chen VB, Arendall WB III, Headd JJ, Keedy DA, Immormino RM, Kapral GJ, Murray LW,Richardson JS, and Richardson DC (2010) MolProbity: all-atom structure validation for macromolecular crystallography. Acta Crystallogr., Sect. D: Biol. Crystallogr 66, 12–21. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Laskowski RA, MacArthur MW, Moss DS, and Thornton JM (1993) PROCHECK: A program to check the stereochemical quality of protein structures. J. Appl. Crystallogr 26, 283–291. [Google Scholar]
  • 49.Cox JD, Kim NN, Traish AM, and Christianson DW (1999) Arginase-boronic acid complex highlights a physiological role in erectile function. Nat. Struct. Biol 6, 1043–1047. [DOI] [PubMed] [Google Scholar]
  • 50.Ash DE, Cox JD, and Christianson DW (1999) Arginase: a binuclear manganese metalloenzyme. In Manganese and Its Role in Biological Processes, Vol. 37 of Metal Ions in Biological Systems, Sigel A, Sigel H, Eds. New York: M. Dekker, 407–428. [PubMed] [Google Scholar]
  • 51.Di Costanzo L, Sabio G, Mora A, Rodriguez PC, Ochoa AC, Centeno F, and Christianson DW (2005) Crystal structure of human arginase I at 1.29-Å resolution and exploration of inhibition in the immune response. Proc. Natl. Acad. Sci. USA 102, 13058–13063. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Hai Y, Edwards JE, Van Zandt MC, Hoffmann KF, and Christianson DW (2014) Crystal structure of Schistosoma mansoni arginase, a potential drug target for the treatment of schistosomiasis. Biochemistry 53, 4671–4684. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53.Ilies M, Di Costanzo L, Dowling DP, Thorn KJ, and Christianson DW (2011) Binding of α,α-disubstituted amino acids to arginase suggests new avenues for inhibitor design. J. Med. Chem 54, 5432–5443. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Gelb MH, Svaren JP, and Abeles RH (1985) Fluoro ketone inhibitors of hydrolytic enzymes. Biochemistry 24, 1813–1817. [DOI] [PubMed] [Google Scholar]
  • 55.Chakrabarti P (1989) Geometry of interaction of metal ions with sulfur-containing ligands in protein structures. Biochemistry 28, 6081–6085. [DOI] [PubMed] [Google Scholar]
  • 56.Arrar M, Turnham R, Pierce L, de Oliviera CAF, and McCammon JA (2013) Structural insight into the separate roles of inositol tetraphosphate and deacetylase-activating domain in activation of histone deacetylase 3. Prot. Sci 22, 83–92. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Decroos C, Christianson NH, Gullet LE, Bowman CM, Christianson KE, Deardorff MA, and Christianson DW (2015) Biochemical and structural characterization of HDAC8 mutants associated with Cornelia de Lange Syndrome spectrum disorders. Biochemistry 54, 6501–6513. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58.Porter NJ, Christianson NH, Decroos C, and Christianson DW (2016) Structural and functional influence of the glycine-rich loop G302GGGY on the catalytic tyrosine of histone deacetylase 8. Biochemistry 55, 6718–6729. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59.Lanyi JK (1974) Salt-dependent properties of proteins from extremely halophilic bacteria. Bacteriol. Rev 38, 272–290. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Dym A, Mavaerech M, and Sussman JL (1995) Structural features that stabilize halophilic malate dehydrogenase from an archaebacterium. Science 267, 1344–1346. [DOI] [PubMed] [Google Scholar]
  • 61.Graziano G, and Merlino A (2014) Molecular basis of protein halotolerance. Biochim. Biophys. Acta 1844, 850–858. [DOI] [PubMed] [Google Scholar]
  • 62.DasSarma S, and DasSarma P (2015) Halophiles and their enzymes: negativity put to good use. Curr. Opin. Microbiol 25, 120–126. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63.Fujishiro K, Ando M, and Uwajima T (1988) Crystallization and some properties of acetylpolyamine amidohydrolase from Mycoplana bullata. Biochem. Biophys. Res. Commun 157, 1169–1174. [DOI] [PubMed] [Google Scholar]
  • 64.Sakurada K, Ohta T, Fujishiro K, Hasegawa M, and Aisaka K (1996) Acetylpolyamine amidohydrolase from Mycoplana ramosa: gene cloning and characterization of the metal-substituted enzyme. J. Bacteriol 178, 5781–5786. [DOI] [PMC free article] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Supporting Information

Figure S1: IC50 plots; Figure S2: ITC enthalpograms.

RESOURCES