Molecular Dynamics Pinpoint the Global Fluorine Effect in Balanoids

Jan 17, 2018 - For the first time to the best of our knowledge, we found that a structurally equivalent residue in each kinase, Thr184 in PKA and Ala5...
0 downloads 8 Views 1MB Size
Subscriber access provided by READING UNIV

Article

Molecular Dynamics Pinpoint the Global Fluorine Effect in Balanoids Binding to PKC# and PKA Ari Hardianto, Fei Liu, and Shoba Ranganathan J. Chem. Inf. Model., Just Accepted Manuscript • DOI: 10.1021/acs.jcim.7b00504 • Publication Date (Web): 17 Jan 2018 Downloaded from http://pubs.acs.org on January 18, 2018

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

Journal of Chemical Information and Modeling is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 29 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Chemical Information and Modeling

1

Molecular Dynamics Pinpoint the Global Fluorine

2

Effect in Balanoids Binding to PKCε and PKA

3

Ari Hardianto, † Fei Liu, † Shoba Ranganathan *, †

4

† Department of Chemistry & Biomolecular Sciences, Macquarie University, Sydney, NSW

5

2109, Australia

6

7

ABSTRACT: (-)-Balanol is an ATP mimic that inhibits protein kinase C (PKC) isozymes and

8

cAMP-dependent protein kinase (PKA) with little selectivity. While PKA is known as a tumour

9

promoter, PKC isozymes can be tumour promoters or suppressors. In particular, PKCε is

10

frequently involved in tumorigenesis and a potential target for anticancer drugs. We recently

11

reported that stereospecific fluorination of balanol yielded a balanoid with enhanced selectivity

12

for PKCε over other PKC isozymes and PKA, although the global fluorine effect behind the

13

selectivity enhancement is not fully understood. Interestingly, in contrast to PKA, PKCε is more

14

sensitive to this fluorine effect. Here we investigate the global fluorine effect on the different

15

binding responses of PKCε and PKA to balanoids using molecular dynamics (MD) simulations.

16

For the first time to the best of our knowledge, we found that a structurally equivalent residue in

17

each kinase, Thr184 in PKA and Ala549 in PKCε, is essential for the different binding

18

responses. Furthermore, the study revealed that the invariant Lys, Lys73 in PKA and Lys437 in

19

PKCε, already known to have a crucial role in the catalytic activity of kinases, serves as the main

20

anchor for balanol binding. Overall, while Thr184 in PKA attenuates the effect of fluorination,

ACS Paragon Plus Environment

1

Journal of Chemical Information and Modeling 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 2 of 29

21

Ala549 permits remote response of PKCε to fluorine substitution, with implications for rational

22

design of future balanol-based PKCε inhibitors.

23

Keywords: balanoids; fluorine effect, PKCε; PKA; binding response; Molecular dynamics

24

simulations.

25 26

INTRODUCTION

27

(-)-Balanol (referred to as balanol) is a fungal natural product1 and a structural mimic of ATP.2

28

It has a four ring structure that completely occupies the flexible ATP site of protein kinases

29

(Figure 1).2, 3 The benzamide moiety (ring A) resides in the adenine subsite, whereas the azepane

30

ring (ring B) occupies the ribose subsite. The benzophenone moiety, comprising rings C and D,

31

fills the triphosphate subsite. As the central moiety,4 the azepane ring connects the benzamide

32

and the benzophenone moieties, using amide and ester linkages, respectively.

33

Balanol non-selectively inhibits cAMP-dependent protein kinase (PKA) and protein kinase C

34

(PKC) isozymes at their ATP sites.5 While PKA consistently shows tumour promoting

35

activities,6 most PKC isozymes can act as tumour promoters or suppressors, depending on the

36

context.7 In breast cancer, for example, PKCβI is a tumour suppressor, whereas PKCα, PKCβII,

37

and PKCδ are tumour promoters.7 On the contrary, PKCβI and PKCδ show promoter and

38

suppressor activity in prostate cancer, respectively. Nonetheless, PKCε prominently displays

39

oncogenic activities and is a potential target for anti-cancer drugs.7 Despite several decades of

40

research, the development of isozyme-specific PKC regulators (both activators and inhibitors)

41

has remained elusive.8 However, the ATP sites among PKC isozymes are highly homologous8

42

and fine tuning ATP mimics, such as balanol, for specific inhibition of PKC isozymes and PKCε

43

in particular, is essential for lead compound discovery in cancer therapy.

ACS Paragon Plus Environment

2

Page 3 of 29 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Chemical Information and Modeling

44

45 46

Figure 1. (A) Balanol structure, divided into three moieties that occupy the different subsites

47

based on structural overlay with ATP.3 (B) Balanol analogues which are fluorinated in the

48

azepane moiety.4 1a and 1c are analogues monofluorinated at C6(S) and C5(S), whereas 1d is

49

difluorinated at C6 while 1e is trifluorinated at C6 and C5(S). 1b (C6(R) – not shown) is the

50

stereoisomer of 1a.

51 52

To enhance selectivity for a specific PKC isozyme, balanol has been subjected to extensive

53

structure and activity relationship (SAR) studies,9-13 with PKA also included for comparison.10,

54

11, 13

55

(ring B),10 and benzophenone moieties (rings C and D).11,

56

moiety revealed the importance of C5′OH for PKC inhibition.9 Derivatisation studies on the

57

benzophenone ring showed that acidic functional groups are necessary for balanol activity.10, 11

58

In some SAR studies,9, 10 the azepane ring was replaced with a five-membered pyrrolidine ring,

59

but this replacement did not substantially improve target selectivity of the balanol derivatives.

These SAR studies explored chemical diversifications on the benzamide (ring A),9 azepane

ACS Paragon Plus Environment

12

The studies on the benzamide

3

Journal of Chemical Information and Modeling 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 4 of 29

60

These studies showed that as the central moiety, the azepane ring could tolerate various

61

derivatisation, but selectivity remains a challenge.11

62

Table 1. Dissociation constant values (Kd) characterising balanoids interacting with PKA or

63

PKCε4.

Protein (-)-balanol, 1 kinase

(6S)-F

(5S)-F

(6)-diF

(5S)-(6R,S)-triF

1a

1c

1d

1e

9.2 ± 0.8

43 ± 4

Kd (nM) PKA

5.9 ± 0.5

7.9 ± 0.5

PKCε

0.73 ± 0.06

19 ± 8

6.4 ± 0.1

0.4 ± 0.02 110 ± 19

38 ± 9.5

64 65

Fluorine is an important substituent in medicinal chemistry and widely utilised in drug

66

development due to its distinctive features.14, 15 It has a small atomic size of 1.47 Å which is

67

close to 1.20 Å for hydrogen (Supporting information: Note 1 and Table S1). Thus, hydrogen

68

atom substitution by fluorine does not significantly increase the size of a molecule.16, 17 As the

69

most electronegative element with Pauling electronegativity of 4.00, fluorine forms a highly

70

polarized C–F bond. The attraction of partial positive and partial negative charges on carbon and

71

fluorine atoms, respectively, makes the C-F bond short and strong with significant ionic

72

character.17 Due to its unique characters, the C–F bond can cause a perturbation to the

73

conformation of a molecule via dipole–dipole interactions, charge–dipole interactions, and

74

hyperconjugation effects (referred to as fluorine perturbation4; see Supporting information: Note

75

1 and Figure S1), which can control a molecule shape.15,

76

offers productive effects on conformation, intrinsic potency, pKa, metabolic pathways,

77

membrane permeability, and pharmacokinetic properties.15 Thus, fluorination may improve

78

balanol selectivity to a specific protein kinase.

17

ACS Paragon Plus Environment

Proper fluorination of a molecule

4

Page 5 of 29 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Chemical Information and Modeling

79

In our recent study,4 we showed that stereospecific fluorination provides a tool to achieve

80

balanol-kinase selectivity in highly homologous ATP sites. We stereospecifically incorporated

81

single fluorine (Figure 1.B, Table 1: 1a and 1c) and multiple fluorine atoms (Figure 1.B, Table 1:

82

1d and 1e) on the azepane moiety of balanol for the first time. 1b (C6(R) – not shown in Figure

83

1) is the stereoisomer of 1a. Binding affinity measurements of balanol and its fluorinated

84

analogues (referred to as balanoids) to PKA and PKC isozymes were assayed to assess the effect

85

of azepane fluorination on protein selectivity.4 We found that fluorination results in different

86

responses of the kinases to balanoids. As 1b showed very similar binding affinity values to 1a,4

87

this balanoid has not been investigated further. In particular, the C5(S)-fluorinated balanol

88

analogue, 1c, enhanced the binding affinity and selectivity for PKCε (see Table 1). This study

89

also revealed an interesting result that, in contrast to PKA, PKCε is more sensitive to the effect

90

of fluorination on balanoids.4

91

Computational analysis to understand the basis of varied binding responses of PKA and PKCε

92

to balanoids will provide useful information for further development of balanol-based ATP-

93

mimicking inhibitors. Our previous molecular docking study, which complements binding

94

affinity measurements, suggests that stereospecific fluorination provides conformational

95

adjustment of balanol in a kinase-dependent manner.4 Here, we present molecular dynamics

96

(MD) simulations, for a more detailed and insightful interaction analysis18 of balanoids to the

97

ATP sites of PKA and PKCε, such as hydrogen bond (H-bond) conservation analysis. The

98

analysis can then be used to study protein flexibility and the fluorine effect or fluorination

99

perturbation.4 This method also offers an investigation of the ligand conformational change

100

during binding and an estimate of the binding energy from ensemble conformations. Moreover,

ACS Paragon Plus Environment

5

Journal of Chemical Information and Modeling 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 29

101

since protein kinases are known to provide induce-fit binding interactions,2 MD simulations give

102

a chance to explore this plasticity in ligand binding19 using PKCε and PKA.

103

In order to carry out MD situations, the charge state (the net charge derived from the formal

104

charges on ionisable functional groups) of balanoids bound to PKA and PKCε need to be

105

correctly assigned. In our previous study,20 we computationally investigated the charge state

106

characteristics of balanoids in the ATP sites of PKA and PKCε. Our results identified the

107

charged state of different functional groups on balanoids (1-1e) that correlated with experimental

108

binding energy values (Table 1). All balanoids have consistent charge states in both PKA and

109

PKCε, except 1a (Supporting information: Figure S2). Briefly, the natural balanol 1 and the most

110

potent analogue 1c bear charges on the azepane ring (N1), the phenolic group (C6′′OH) and the

111

carboxylate (C15′′O2H) on the benzophenone moiety. The multiple fluorinated analogues (1d

112

and 1e) have charges on the C6′′OH and the C15′′O2H. Balanoid 1a can exit in two charge states.

113

One species of 1a (referred to as 1a.1) only carries a charge on the C15′′O2H, whereas the other

114

(1a.2) bears charges on the N1, C′6′OH, and C15′′OOH. In PKA, 1a.1 is the preferred state while

115

in PKCε, both charge states correlate well with the experimental binding affinity value.20 This

116

charge state information is important for conducting the MD analysis on balanoids bound to the

117

ATP sites of PKA as well as PKCε.

118

Here, we present MD analysis of balanoid binding in the ATP sites of PKA and PKCε, using

119

the charge states determined in our earlier study.20 The results suggest that stereospecific

120

fluorination influences the shape, conformational bias, and bound position of the azepane ring in

121

the ribose subsite that contribute to differential binding responses of PKA and PKCε to

122

balanoids. We also find that an invariant Lys, which is essential for the kinase activity, is the

123

major contributor in the binding energy of balanoids in both kinases. More importantly, to the

ACS Paragon Plus Environment

6

Page 7 of 29 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Chemical Information and Modeling

124

best of our knowledge, we find, for the first time, that a single amino acid residue residing in the

125

ribose subsite and close to the adenine subsites, Thr184 in PKA and Ala549 in PKCε, is essential

126

for global fluorine effect in distinct binding responses of the two kinases. This residue plays a

127

unique role in modulating the remote effect of fluorine onto other moieties of balanoids and thus

128

prescribe the different levels of sensitivities from PKA and PKCε to different fluorine

129

substitutions. Finally, this study provides pertinent information for future rational design of

130

balanol analogues to achieve improved PKC isozyme selectivity.

131 132

MATERIALS AND METHODS

133

Homology modelling

134

We built kinase domains of human PKA (UniProt ID: P17612) and PKCε (UniProt ID:

135

Q02156) which bind balanol using a homology modelling approach as described previously.20

136

The X-ray crystal structure of mouse PKA-balanol (PDB ID: 1BX6)3 was used as the template

137

for homology modelling of human PKA. In modelling human PKCε, we utilised two X-ray

138

crystal structure templates, the mouse PKA (PDB ID: 1BX6) and a human PKCη (PDB ID:

139

3TXO). To examine the structural conservation between both templates, we carried out structural

140

alignment using jCE algorithm.21 Subsequently, we used CLUSTALX 2.122 to align the sequence

141

of human PKCε to 1BX6 and 3TXO sequences. Additionally, the resulting sequence alignment

142

was manually edited to map the ‘open’ conformation23 of the Gly-rich loop (GXGXXG) from

143

1BX6 to PKCε model. Calculation of the percentage identity and similarity between human PKA

144

and PKCε sequences was carried using MatGAT.24

145

Homology modelling was performed using MODELLER 9.14.25 The modelling process

146

included the balanol transfer, from mouse PKA-balanol (PDB ID: 1BX6) to the generated query

147

structure, to maintain the three dimensional (3D) features of the ATP binding site in the resulting

ACS Paragon Plus Environment

7

Journal of Chemical Information and Modeling 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 8 of 29

148

model. The resulting models were assessed using the Discrete Optimized Protein Energy

149

(DOPE) score26 which reflects the quality of the model. Furthermore, we superimposed models

150

from multiple runs of homology modelling and selected the structure which possessed residues

151

adopting consensus conformations. The selected model was also evaluated using Ramachandran

152

plot on PROCHECK webserver.27

153

Molecular dynamics simulation preparation and protocol

154

As fully activated human PKA and PKCε are phosphorylated at specific sites,28 we added

155

phosphate groups on PKA (Thr198 and Ser339) and PKCε (Thr566, Thr710, and Ser729) using

156

Discovery Studio Visualizer.29 The starting conformations of balanoids were adopted from the

157

ligand in 1BX6 (mouse PKA with bound balanol). Atomic charges of balanoids were computed

158

using the Austin Model 1 - Bond Charge Corrections (AM1-BCC)30 protocol in

159

AmberTools16.31 To determine parameters for balanoids, we used the General Amber Force-

160

fields (GAFF)32 and the parmchk program in AmberTools16. For protein force-fields, we

161

assigned regular amino acid residues and their side chains with ff14SB33 and phosphorylated

162

residues with phosaa10.34

163

In the MD system preparation, we solvated each complex of either PKA- or PKCε-balanoid

164

with explicit water molecules using TIP3P,35 with a bounding box of 10 Å. We also added

165

sufficient Na+ and Cl- ions to achieve neutral charge and a salt concentration of 0.15 M, which is

166

the equivalent of physiological salt concentration, using the tleap utility in AmberTools16.

167

We utilised GPU-accelerated Particle-Mesh Ewald Molecular Dynamics (PMEMD), as

168

implemented in Amber16,31 throughout simulations where periodic boundary conditions were

169

applied. For every simulation, two consecutive steps of energy minimisation were performed

170

with restraints of 25 and 5 kcal.mol-1.Å-2 to the solute, the protein-ligand complex. The system

ACS Paragon Plus Environment

8

Page 9 of 29 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Chemical Information and Modeling

171

temperature was then increased to 300 K under NVT condition for 50 ps, followed by

172

equilibration steps. In the following NPT simulation over 50 ps, the system density was

173

equilibrated to 1 g.cm-1. Subsequently the system was switched to NVT. In the NVT simulation,

174

the restraint on solute was gradually decreased from 5 kcal.mol-1.Å-2 by 1 kcal.mol-1.Å-2 every 50

175

ps where in the last 50 ps simulation, the restraint was totally removed.

176

Production-phase for each system was simulated at 300 K under the NPT to yield 100 ns

177

trajectory, which was recorded every 10 ps. Long-range electrostatic interactions were treated by

178

particle-mesh Ewald (PME) method. Meanwhile, a 10 Å cut-off was used for short-range non-

179

bonded interactions which was also employed by others.36,

180

implemented to constrain all bonds involving hydrogen atoms. To maintain constant pressure

181

and temperature, we used algorithms of Berendsen barostat39 and Langevin thermostat,40

182

respectively.

37

A SHAKE algorithm38 was

183

Binding energy calculation

184

° Experimental binding energy values (∆ ) of balanoids to PKA or PKCε were calculated

185

from dissociation constant (Kd) values (Table 1).4 At equilibrium and under standard conditions,

186

the binding energy is directly related to the equilibrium constants and can be computed using the

187

following formula (eq. 1):

188 189

° ∆ = −   =  

(1)

190 191

° where ∆ is experimental binding energy, Ka and Kd are association and dissociation constant

192

respectively, R is the universal gas constant, and T is absolute temperature.

ACS Paragon Plus Environment

9

Journal of Chemical Information and Modeling 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 10 of 29

193

Estimated binding energy values of balanoids to PKA or PKCε as well as their per-residue

194

decomposition were determined using Molecular Mechanics Generalised Born Surface Area

195

(MMGBSA) method41 as implemented in MMPBSA.py.42 The MMGBSA binding free energy

196

° (∆ ) is calculated as follows in eq. 2:

197 198

° ∆ = 〈 〉 − 〈 〉 − 〈 〉

(2)

199 200

° where 〈 〉 , 〈 〉 , and 〈 〉 are the average value of ∆ for complex, enzyme,

201

and ligand, respectively, over snapshots i extracted from MD trajectories.  can be broken

202

down as shown in eq. 3:

203 204

   = ! + #$ + #$

(3)

205 206

 where ! is the gas phase energy, #$ the electrostatic portion of solvation energy

207

 determined using Generalised Born (GB) implicit solvent model, and #$ the hydrophobic

208

contribution to the solvation energy. The hydrophobic contribution is approximated by the

209

Linear Combination of Pairwise Overlaps (LCPO) method.43 ! is estimated by the molecular

210

mechanics energy of the molecule consisting of internal energy terms: bond (Ebond), angle

211

(Eangle), and torsion energies (Etorsion) as well as van der Waals (EvdW), and electrostatic

212

interactions (Eel) (eq. 4).

213 ! = % !&' + % !' &'#

' #



+

(#

(#

*+

*+

% !(#' + % !$) + % !(#((

(#'#

214

(4)

ACS Paragon Plus Environment

10

Page 11 of 29 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Chemical Information and Modeling

215 216

Internal energy terms were omitted in this study, since single-trajectory MD simulations were

217

used in the calculation.44

218

MD trajectory analysis

219

We utilised cpptraj program in AmberTools16 to analyse MD trajectories. The analysis

220

includes computation of H-bond conservation, dihedral angles in the azepane ring, Solvent-

221

Accessible Surface Area (SASA), and Root-Mean-Square Fluctuation (RMSF).

222 223

Images and graphs generation

224

Conformational ensembles of balanoids were generated in VMD 1.9.2.45 For analysis

225

purposes, balanoid conformations were collated at intervals of 1250 frames, from the MD

226

trajectories for 50-100 ns (5000 frames). Representative non-covalent interactions of kinase-

227

balanoid, according to H-bond conservation analysis, were visualised using BIOVIA Discovery

228

Studio Visualizer 2016.29 2D structures of balanoids were sketched in BIOVIA Draw 2016.46 All

229

plots were generated using an R47 package (ggplot2)48 in RStudio 0.99.892.49 GNU Image

230

Manipulation Program (GIMP) 2.8.1450 and Inkscape 0.48.551 were utilised for image editing.

231 232

RESULTS AND DISCUSSION

233

Overall effect of balanoid binding on conformational freedom of PKA and PKCε

234

The binding of balanoids to PKA and PKCε reduces the conformational freedom of both

235

kinases, regardless of their binding affinity values (Supporting information: Figure S3). For

236

instance, the binding of 1e has a similar effect to other balanoids in decreasing conformational

237

freedom of PKA, although 1e binds relatively weakly to this kinase. Likewise, 1c and 1d, ligands

ACS Paragon Plus Environment

11

Journal of Chemical Information and Modeling 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 29

238

with the strongest and the weakest affinity to PKCε respectively,4 reduce the conformational

239

freedom of PKCε. Moreover, the Wilcoxon sum-rank test shows that the binding of 1c or 1d

240

leads to a statistically significant reduction of conformational freedom of the Gly-rich loop, the

241

ATP pocket, and even the entire kinase domain of PKCε (Supporting information: Table S2).

242

To investigate whether the strength of balanoid binding to PKA or PKCε is related to their

243

Solvent-Accessible Surface Area (SASA), we conducted accessible surface calculations. The

244

results suggest that there is no notable effect on SASA on the binding affinity of balanoids to

245

PKA or PKCε. For example, 1d- and 1c-bound PKCε have comparable SASA (Supporting

246

information: Figure S4). Likewise, 1e and other balanoid-bound PKA complexes show similar

247

SASA.

248 249 250

Comparison of conformational ensembles of balanoids in the ATP sites of PKA and PKCε

251

Sequence analysis24 shows that PKA and PKCε share 43% identity and 68% similarity in their

252

kinase domains. At the ATP sites, however, their residue identity and similarity level are higher

253

(68% and 79%, respectively). Of the nine ATP-binding residues (indicated by green bars in

254

Supporting Information: Figure S5) six are conservatively substituted and two are non-

255

conservatively substituted. Visual inspection of overlaid PKA and PKCε homology models (data

256

not shown) reveals that two (occurring in the Gly-rich loop of the kinases, in Supporting

257

Information: Figure S5, underlined by orange bars) of these nine non-identical residues

258

participate in the ATP site only through their backbones. Other non-identical residues between

259

PKA and PKCε mainly reside outside of the ATP site.

260

The natural product 1 binds stronger to PKCε than PKA (0.73 nM vs. 5.9 nM,4 respectively).

261

Superimposition of conformational ensembles (Figure 2.A) revealed that 1 binds differently in

ACS Paragon Plus Environment

12

Page 13 of 29 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Chemical Information and Modeling

262

the ATP sites of the two kinases. Overall, 1 in PKA is more flexible than that in PKCε

263

(Supporting information: Note 2). The flexibility of each moiety of the ligand increases in the

264

following order: benzamide moiety < azepane ring < benzophenone moiety. The benzamide

265

moiety shows comparable conformation and binding mode in both kinases, with limited

266

flexibility, whereby the azepane ring in PKA is more mobile than in PKCε. The largest

267

conformational variations occur on the benzophenone moiety in the ATP site of PKA and PKCε.

268

In PKA, the benzophenone moiety of balanol (1) is closer to the ATP site entrance, whereas in

269

PKCε, this moiety moves deeper inside the triphosphate subsite.

ACS Paragon Plus Environment

13

Journal of Chemical Information and Modeling 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 14 of 29

270 271 272

Figure 2. Superimposition of conformational ensemble of balanol (1) and its fluorinated

273

analogues bound to PKCε and PKA. (A) depicts the superimposition of PKCε- (grey) and

274

PKA-bound 1 (white) where the right side figure displays a 90°-rotated view. Superimposition of

275

1 (shown in black) and its fluorinated analogues (1a-1e in green, pick, blue and orange,

276

respectively) bound to PKA is shown in (B) and to PKCε in (C). 1a.1 is shown, as a

277

representative for 1a. Each ATP subsite occupied by the respective balanol moiety is represented

ACS Paragon Plus Environment

14

Page 15 of 29 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Chemical Information and Modeling

278

in dashed lines: blue for the adenine subsite, yellow for the ribose subsite, and orange for the

279

triphosphate subsite.

280

Stereospecific fluorination of 1 on its azepane ring results in large changes in binding affinity

281

to PKCε (Table 1). In one instance, C5(S)-fluorinated analogue (1c) shows improved binding

282

affinity, compared to 1. By contrast, such binding affinity variation is not observed in PKA

283

(Table 1). PKA is less sensitive to the effect of fluorination than PKCε,4 which is consistent with

284

superimposition of conformational ensembles of balanoids (Figure 2.B and C). During binding,

285

all fluorinated balanoids in the ATP site of PKA exhibit conformational changes on their azepane

286

moiety relative to 1, but these changes do not alter the binding conformations of the other

287

moieties (Figure 2.B). Slight alterations occur only on the benzophenone moiety. Balanoid 1e,

288

which is the ligand with the weakest affinity to PKA, appears to have its benzophenone moiety

289

closer to the entrance of the ATP site. In the ATP site of PKCε, conformational changes on the

290

azepane ring are accompanied by changing positions of both the benzamide and benzophenone

291

moieties (Figure 2.C). As the ligand with the strongest affinity to PKCε, 1c shows enhanced

292

interaction to the protein whereby its benzophenone moiety is deeper inside the triphosphate

293

subsite. This is contrasted by 1d, the weakest binding ligand to PKCε, in which case the

294

benzophenone moiety is found closer to the ATP site entrance. The residential position of the

295

benzophenone moiety seems to correlate with the binding affinity of balanoids, whereby the

296

deeper this moiety is buried in the triphosphate subsite, the stronger the binding affinity.

297 298 299

A single residue important for the local and remote effects of fluorine substitution in PKA and PKCε

ACS Paragon Plus Environment

15

Journal of Chemical Information and Modeling 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 16 of 29

300

Thr184 in the ATP site of PKA and its corresponding residue in PKCε, Ala549, may be a

301

critical amino acid residue for the varied response of both kinases to balanoids due to local and

302

remote effects of fluorine substitution. As previously reported,52 the mutation T184A expands

303

the ATP site of PKA and improves the binding of several kinase inhibitors, particularly JNJ-

304

7706621, VX-680, and H-89. However, no mutation study of Ala549 in PKCε has yet been

305

reported. The mutation study in PKCε addressed only the phosphorylation sites,53 rather than the

306

ATP-binding site as a potential target for kinase inhibition.

307 308

Figure 3. Interactions between representative balanoids and the invariant Lys and the

309

azepane ring residues of PKA and PKCε. 1c and 1d are respectively the best and the worst

310

ligands for PKCε (Table1). (A) and (B) respectively depict interactions of 1c in PKA and PKCε

311

while (C) and (D) show interactions of 1d in PKA as well as PKCε, respectively. The invariant

312

Lys and the representative azepane ring are highlighted in cyan and orange, respectively. The

ACS Paragon Plus Environment

16

Page 17 of 29 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Chemical Information and Modeling

313

key determinant residue (Thr184 in PKA and Ala549 in PKCε) is highlighted in yellow. Water-

314

mediated intramolecular interactions are highlighted in green. H-bonds are indicated by green

315

dashed lines, whereas electrostatic or ion pair interactions are depicted by orange dashed lines.

316

As 1c and 1d are respectively the best and the worst ligands for PKCε (Table1), the following

317

sections will focus on these two balanoids. The interactions of the other two balanoids, 1a and

318

1e, are discussed in Supporting Information: Note 3. In PKA, the side chain of Thr184 is a key

319

relay station for a hydrogen bond network between the C1'=O of the amide linkage, the azepane,

320

and the ribose subsite of the protein (Figure 3.A and C, highlighted in yellow). Such a hydrogen

321

bond network helps to anchor the benzamide moiety in the adenine subsite and dampens the

322

transmission of the fluorination effect in the azepane ring. As a result, the cascade effect of

323

fluorination to the benzamide and benzophenone moiety is attenuated (Figure 2.B). On the

324

contrary, the replacement of Thr184 to Ala549 in PKCε (Figure 3.B and D, highlighted in

325

yellow) abrogates the essential H-bonding that holds and minimises the fluorine effect in the

326

azepane ring. As a consequence, perturbations caused by fluorination not only significantly

327

affect the azepane ring locally but also remotely influence interactions of the other moieties,

328

particularly in the benzophenone moiety, with the ATP site residues of PKCε (Figure 2.C).

329

Balanoid 1c (C5(S)-fluorinated balanol analogue) is the best available example that the

330

fluorine effect improves affinity to PKCε but not PKA (Table 1). In PKCε, the presence of a

331

fluorine atom at C5(S) in the azepane ring of 1c allows the intramolecular interaction with NH of

332

the amide linkage that is facilitated by a water molecule (Figure 3.B, highlighted in green).

333

Furthermore, the non-bonding orbital of the fluorine atom (σ*C5(S)-F) appears to be in

334

hyperconjugation with σC6-H, a fluorine effect suggested by Gillis et al.15 and Hunter.17 This

335

may serve to stabilise the azepane ring conformation (Supporting Information: Figure S6, 1c in

ACS Paragon Plus Environment

17

Journal of Chemical Information and Modeling 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 18 of 29

336

PKCε). Additionally, the ammonium group of 1c form an H-bond with the backbone of Asn537

337

and an ion pair interaction with the side chain of Asp550 (Figure 3.B; orange dashed line).

338

Altogether, these interactions shape the conformation ensemble and bound position of the

339

azepane ring and further position the benzophenone moiety to interact better with residues in the

340

triphosphate subsite. In PKA, the azepane ring of 1c possesses a similar conformation

341

(Supporting Information: Figures S6, S7 and S8) to that of PKCε-bound 1c. Moreover, the same

342

water-mediated intramolecular interaction also occurs (Figure 3.A). Also, the ammonium group

343

of 1c form an H-bond with the side chain of Glu171 and an ion pair interaction with the side

344

chain of Asp185 (Figure 3.A; orange dashed line). Nevertheless, Thr184 anchors the azepane

345

ring and deters the remote effect of C5(S)-fluorination to the benzophenone moiety.

346

The fluorine effect on balanol may lead to a dramatic fall in binding affinity to PKCε as shown

347

by 1d (doubly fluorinated at C6 of the azepane ring, which leads to a zero formal charge on

348

N1)20 (Table 1). Without Thr184 that anchors C1'=O, the uncharged N1 group20 cannot maintain

349

H-bonds with Asp550 (Supporting information: Table S3). Meanwhile, one fluorine atom in 1d

350

interacts with the backbone of Lys416 and slightly moves the azepane moiety (Figure 3.D). Such

351

a change in the position of the azepane ring remotely interferes with the interactions between the

352

benzophenone moiety and residues in the triphosphate subsite. As seen in Table S3 (Supporting

353

information), the benzophenone moiety is missing H-bonds with Phe419 and Lys437, residues at

354

the triphosphate subsite. On the other hand in PKA, the interaction of Thr52 (which corresponds

355

to Lys416 in PKCε) with the azepane ring is counterbalanced by Thr184. The residue Thr184

356

also allows water-mediated intramolecular interactions that involve fluorine substituents and the

357

N1 group (Figure 3.C). These interactions with the azepane ring modify and restrain its

ACS Paragon Plus Environment

18

Page 19 of 29 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Chemical Information and Modeling

358

conformation. Therefore, unlike the case with PKCε-bound 1d, PKA minimises the remote effect

359

of fluorine substitution on the binding of the benzophenone moiety.

360

Overall, balanoids 1c and 1d together demonstrate the malleability of PKCε to cooperatively

361

accommodate perturbations due to fluorine substitutions, that is absent in PKA. The structurally

362

equivalent residue in each kinase, Thr184 in PKA and Ala549 in PKCε, seems to be crucial for

363

the different responses of PKCε and PKA to the fluorine effect. Details of this important residue

364

and the malleability of PKCε are valuable for designing the next generation of balanoids for

365

improved affinity and specificity to PKCε and not to PKA.

366 367

Invariant Lys as the major contributor residue in balanoid binding

368

The invariant Lys in the triphosphate subsite is the most conserved residue of protein kinases.

369

It participates in the phosphotransfer reaction and orients ATP for catalysis by interacting with

370

the α- and β-phosphate groups of the nucleotide. Furthermore, this residue is also involved in

371

stabilising the conformation of other catalytically active kinases.54 Interestingly, our results

372

suggest that the invariant Lys (Figure 3, highlighted in blue) also frequently acts as a major

373

contributor to balanoid binding (Supporting information: Figure S9). This may indicate that

374

future inhibitor design may need to preserve and optimise interactions to the invariant lysine

375

residue.

376

In PKCε, this invariant residue is Lys437, which contributes higher binding energy (mean = -

377

16.33 ± 3.62 kcal.mol-1, Supporting information: Figure S9) than the corresponding Lys73 in

378

PKA (mean = -11.16 ± 2.06 kcal.mol-1). This higher binding energy could be attributed to

379

Lys437 providing more conserved H-bonding networks than Lys73 (Supporting information:

380

Table S3). Nevertheless, Lys73 in PKA remains the most prominent contributor to binding than

ACS Paragon Plus Environment

19

Journal of Chemical Information and Modeling 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 20 of 29

381

the other residues in PKA to balanoids, except to 1a. We found that the invariant Lys can non-

382

covalently bind balanoids in four ways: H-bonding, charge-charge, alkyl-π hydrophobic, and

383

cation-π interactions (Supporting information: Figure S10). The last two interactions are

384

conserved in binding of all balanoids (1 to 1e) to PKA as well as PKCε. Such interactions may

385

be responsible for the stronger binding contribution of this invariant Lys (Supporting

386

information: Figure S9) more so than other kinase residues, even with transient H-bonding

387

interaction observed (Supporting information: Table S3).

388

The natural balanol 1 builds an H-bonding network with Lys437 in PKCε via its C1'=O of the

389

amide linkage. The hydrogen bonding network may persist or dissipate due to the effect of

390

fluorine substitutions that occur in the azepane ring of balanoids (Supporting information: Note

391

4). C5(S)-Fluorine substitution in 1c, however, eliminates the H-bonding between the C1'=O and

392

Lys437 and directs its phenolate C6''-O- group to interact intensively with the invariant Lys. The

393

interaction between the phenolate group C6''-O- and the positively charged side chain of Lys437

394

provides not only H-bonding but also a charge-charge interaction. Moreover, Lys437 forms

395

alkyl-π hydrophobic and cation-π interactions with the ring C of 1c which make this balanoid

396

exhibit all possible noncovalent binding interactions with the invariant Lys. As a result, the

397

binding of 1c to Lys437 in PKCε is the strongest among the balanoids studied here.

398

In PKA, the presence of Thr184 restrains the azepane rings of all balanoids. These restrained

399

azepane rings limit the chance of the invariant Lys73 to form H-bonds with balanoids, as shown

400

in Table S3 (Supporting information). Hence, the binding of balanoids to PKA is unable to

401

completely interact with Lys73, particularly in terms of H-bonding. As a consequence, the

402

binding energy contribution of the invariant Lys73 is within a narrow range for all balanoids (-

403

8.54 to -13.76 kcal.mol-1; Supporting information: Figure S9) and weaker than the contributions

ACS Paragon Plus Environment

20

Page 21 of 29 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Chemical Information and Modeling

404

of Lys437 to PKCε-balanoid interaction (-12.86 to -21.39 kcal.mol-1). In particular, the

405

restraining effect of Thr184 on the triply fluorinated balanoid, 1e, diminishes its binding affinity

406

to both PKA and PKCε (Supporting Information Note 5 and Figure S11).

407

As we discussed above, the invariant Lys residues in both PKA and PKCε provide the main

408

contribution to balanoid binding over the other ATP site residues. While in PKA, the invariant

409

Lys73 cannot optimise the interactions with balanoids due to the restraining effect from Thr184,

410

Lys437 in PKCε is not under any restraint in its binding to 1c and, thus, provides a substantial

411

binding energy contribution to this balanoid. This finding suggests that optimising interaction

412

with Lys437 is an important requirement for designing PKCε inhibitors with greater affinity than

413

1c.

414

Interactions with the invariant Lys may also explain the correlation of the residential position

415

of the benzophenone moiety, where the deeper this moiety is buried in the triphosphate subsite,

416

the stronger the binding affinity (Figure 2). The invariant Lys is located inside the triphosphate

417

subsite. Thus, intensive interactions of a balanoid to this residue lead the benzophenone moiety

418

residing deeper inside the subsite, as observed in 1c bound to PKCε (Figure 2.C).

419

Unfavourable binding interaction

420

The effect of fluorine substitution in the azepane ring may generate unfavourable binding

421

interactions in PKCε as well as PKA. For instance, in PKCε, the weakest binding ligand 1d

422

involves unfavourable binding contributions from Asp449 (0.02 kcal.mol-1) and Asn537 (0.28

423

kcal.mol-1). The unfavourable binding interaction from Asn537 is the result of water infiltration

424

that prevents the azepane moiety from interacting with the ribose subsite, whereas the one from

425

Asp449 is due to a remote fluorine effect in the azepane ring which interferes with the binding of

426

benzophenone moiety to the triphosphate subsite. Such unfavourable binding interactions explain

ACS Paragon Plus Environment

21

Journal of Chemical Information and Modeling 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 22 of 29

427

why 1d is a poor PKCε ligand. Moreover, a relatively weaker binding energy contribution (-

428

13.82 kcal.mol-1) provided by invariant Lys (Lys437) may also contribute to the very low

429

binding affinity of 1d to PKCε (110 nM). On the contrary, fluorine substitution in 1c does not

430

result in any unfavourable binding interaction with the residues of the ATP site of PKCε

431

(Supporting information: Figure S9).

432

In PKA, an unfavorable binding interaction (+0.04 kcal.mol-1) is only found in the ATP site

433

with 1e, the weakest PKA ligand (Kd = 43 nM) among balanoids. The presence of such an

434

unfavorable binding energy contribution reveals that the fluorine substitution in 1e interferes

435

with the benzophenone moiety interacting optimally with residues of the ATP site of PKA.

436

In summary, we show that fluorine substitution can generate unfavourable binding interaction

437

between balanoids and the ATP sites. Nonetheless, judicious stereospecific fluorine

438

incorporation, such as in 1c, can produce a constructive effect which is free from any

439

unfavourable binding interaction. Thus, the design of the next balanol-based inhibitors needs to

440

consider substituents to avoid unfavourable binding interactions in the ATP site.

441 442

CONCLUSIONS

443

In this current study, using a molecular dynamics simulation approach, we unravelled the

444

perturbation effects of stereospecific fluorination on balanoid binding affinity in the ATP sites of

445

PKCε as well as PKA. Stereospecific fluorination affects balanoid binding in the ATP sites of

446

both kinases, by modifying the kinase conformation and ligand flexibility, especially the binding

447

mode of the azepane ring in the ribose subsite. The fluorination effects further remotely

448

influence the benzophenone moiety interactions in the triphosphate subsite. The effects are

449

receptor-dependent, which is also reflected in the experimental different binding affinity values

ACS Paragon Plus Environment

22

Page 23 of 29 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Chemical Information and Modeling

450

of balanoids to PKA and PKCε. For instance, in PKCε-bound 1c, the azepane ring conformation

451

cooperatively set the benzophenone moiety to optimise interactions with the ATP site residues,

452

especially with the invariant Lys437. From our study, the invariant Lys, which has an essential

453

role in catalytic activity, is the major contributor for binding affinity of balanoids to PKA and

454

PKCε. On the other hand, in PKA, the azepane ring of 1c leads to an interaction decrease with

455

the invariant Lys73.

456

Here, for the first time to the best of our knowledge, we also found an interesting factor that

457

causes susceptibility of PKCε but insensitivity of PKA, to fluorination substitution. The

458

insensitivity of PKA is due to the presence of Thr184, which is situated in the ribose subsite,

459

close to the adenine subsite. Thr184 firmly holds the amide linkage through an H-bond and, thus,

460

dampens fluorine effect in the azepane ring. In PKCε, Ala549 replaces Thr184, preventing the

461

formation of an H-bond that protects the azepane ring from fluorine perturbation. As a result,

462

PKCε is sensitive to fluorine substitution, and our study showed that 1c, fluorinated at C5(S),

463

provides a fluorine effect which improves binding affinity. Overall, this study provides valuable

464

information which facilitates the further rational design of balanol-based inhibitor targeting

465

PKCε for cancer therapy.

466 467

ASSOCIATED CONTENT

468

Notes 1-5, Table S1-S3 and Figures S1-S11(PDF)

469

AUTHOR INFORMATION

470

Corresponding Author

471

SR: [email protected]

ACS Paragon Plus Environment

23

Journal of Chemical Information and Modeling 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 24 of 29

472

Author Contributions

473

AH carried out the computational simulation study and drafted the manuscript. AH, FL, and SR

474

participated in the design of the study and interpretation of data. SR, FL and AH finalised the

475

manuscript. All authors have read and approved the final manuscript.

476

Notes

477

The authors declare that they have no competing interests.

478

ACKNOWLEDGMENT

479

We acknowledge the Indonesia Endowment Fund for Education scholarship to AH. This project

480

was partially supported by resources and services from the National Computational

481

Infrastructure (NCI), funded by the Australian Government.

482

ABBREVIATIONS

483

ATP: adenosine triphosphate; AM1-BCC: Austin Model 1 - Bond Charge Corrections; Balanol:

484

(-)-balanol; DOPE: Discrete Optimized Protein Energy; GAFF: General Amber Force-fields;

485

LCPO: Linear Combination of Pairwise Overlaps; MD: Molecular Dynamics; MMGBSA:

486

Molecular Mechanics Generalized Born Surface Area; PKA: cAMP-dependent protein kinase;

487

PKC: protein kinase C; PMEMD: Particle-Mesh Ewald Molecular Dynamics; RMSF: Root-

488

Mean-Square Fluctuation; SASA: Solvent-Accessible Surface Area.

489 490

REFERENCES

491 492 493

(1) Kulanthaivel, P.; Hallock, Y. F.; Boros, C.; Hamilton, S. M.; Janzen, W. P.; Ballas, L. M.; Loomis, C. R.; Jiang, J. B.; Katz, B., Balanol: a Novel and Potent Inhibitor of Protein Kinase C from the fungus Verticillium balanoides. J. Am. Chem. Soc. 1993, 115, 6452-6453.

ACS Paragon Plus Environment

24

Page 25 of 29 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

494 495 496 497 498 499 500 501 502 503 504 505 506 507 508 509 510 511 512 513 514 515 516 517 518 519 520 521 522 523 524 525 526 527 528 529 530 531 532 533 534 535 536 537 538

Journal of Chemical Information and Modeling

(2) Taylor, S. S.; Yang, J.; Wu, J.; Haste, N. M.; Radzio-Andzelm, E.; Anand, G., PKA: A Portrait of Protein Kinase Dynamics. Biochim. Biophys. Acta 2004, 1697, 259-269. (3) Narayana, N.; Diller, T. C.; Koide, K.; Bunnage, M. E.; Nicolaou, K. C.; Brunton, L. L.; Xuong, N.-H.; Ten Eyck, L. F.; Taylor, S. S., Crystal Structure of the Potent Natural Product Inhibitor Balanol in Complex with the Catalytic Subunit of cAMP-Dependent Protein Kinase. Biochemistry 1999, 38, 2367-2376. (4) Patel, A. R.; Hardianto, A.; Ranganathan, S.; Liu, F., Divergent Response of Homologous ATP Sites to Stereospecific Ligand Fluorination for Selectivity Enhancement. Org. Biomol. Chem. 2017, 15, 1570-1574. (5) Koide, K.; Bunnage, M. E.; Gomez Paloma, L.; Kanter, J. R.; Taylor, S. S.; Brunton, L. L.; Nicolaou, K. C., Molecular Design and Biological Activity of Potent and Selective Protein Kinase Inhibitors Related to Balanol. Chem. Biol. 1995, 2, 601-608. (6) Cho, Y. S.; Lee, Y. N.; Cho-Chung, Y. S., Biochemical Characterization of Extracellular cAMP-Dependent Protein Kinase as a Tumor Marker. Biochem. Biophys. Res. Commun. 2000, 278, 679-684. (7) Garg, R.; Benedetti, L. G.; Abera, M. B.; Wang, H.; Abba, M.; Kazanietz, M. G., Protein Kinase C and Cancer: What We Know and What We Do Not. Oncogene 2014, 33, 5225-37. (8) Mochly-Rosen, D.; Das, K.; Grimes, K. V., Protein Kinase C, an Elusive Therapeutic Target? Nat. Rev. Drug Discov. 2012, 11, 937-957. (9) Hu, H.; Mendoza, J. S.; Lowden, C. T.; Ballas, L. M.; Janzen, W. P., Synthesis and Protein Kinase C Inhibitory Activities of Balanol Analogues with Modification of 4Hydroxybenzamido Moiety. Bioorg. Med. Chem. 1997, 5, 1873-1882. (10) Crane, H. M.; Menaldino, D. S.; Erik Jagdmann, G.; Darges, J. W.; Buben, J. A., Increasing the Cellular PKC Inhibitory Activity of Balanol: A Study of Ester Analogs. Bioorg. Med. Chem. Lett. 1995, 5, 2133-2138. (11) Heerding, J. M.; Lampe, J. W.; Darges, J. W.; Stamper, M. L., Protein Kinase C Inhibitory Activities of Balanol Analogs Bearing Carboxylic Acid Replacements. Bioorg. Med. Chem. Lett. 1995, 5, 1839-1842. (12) Lampe, J. W.; Biggers, C. K.; Defauw, J. M.; Foglesong, R. J.; Hall, S. E.; Heerding, J. M.; Hollinshead, S. P.; Hu, H.; Hughes, P. F.; Jagdmann, G. E.; Johnson, M. G.; Lai, Y.-S.; Lowden, C. T.; Lynch, M. P.; Mendoza, J. S.; Murphy, M. M.; Wilson, J. W.; Ballas, L. M.; Carter, K.; Darges, J. W.; Davis, J. E.; Hubbard, F. R.; Stamper, M. L., Synthesis and Protein Kinase Inhibitory Activity of Balanol Analogues with Modified Benzophenone Subunits. J. Med. Chem. 2002, 45, 2624-2643. (13) Nicolaou, K. C.; Koide, K.; Bunnage, M. E., Total Synthesis of Balanol and Designed Analogues. Chem. Eur. J. 1995, 1, 454-466. (14) Hu, X.-G.; Hunter, L., Stereoselectively Fluorinated N-Heterocycles: A Brief Survey. Beilstein J. Org. Chem. 2013, 9, 2696–2708. (15) Gillis, E. P.; Eastman, K. J.; Hill, M. D.; Donnelly, D. J.; Meanwell, N. A., Applications of Fluorine in Medicinal Chemistry. J. Med. Chem. 2015, 58, 8315-59. (16) Khakshoor, O.; Wheeler, S. E.; Houk, K. N.; Kool, E. T., Measurement and Theory of Hydrogen Bonding Contribution to Isosteric DNA Base Pairs. J. Am. Chem. Soc. 2012, 134, 3154-3163. (17) Hunter, L., The C-F Bond as a Conformational Tool in Organic and Biological Chemistry. Beilstein J. Org. Chem. 2010, 6, 38.

ACS Paragon Plus Environment

25

Journal of Chemical Information and Modeling 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

539 540 541 542 543 544 545 546 547 548 549 550 551 552 553 554 555 556 557 558 559 560 561 562 563 564 565 566 567 568 569 570 571 572 573 574 575 576 577 578 579 580 581 582 583 584

Page 26 of 29

(18) Mortier, J.; Rakers, C.; Bermudez, M.; Murgueitio, M. S.; Riniker, S.; Wolber, G., The Impact of Molecular Dynamics on Drug Design: Applications for the Characterization of Ligand-Macromolecule Complexes. Drug Discov. Today 2015, 20, 686-702. (19) Hospital, A.; Goni, J. R.; Orozco, M.; Gelpi, J. L., Molecular Dynamics Simulations: Advances and Applications. Adv. Appl. Bioinform. Chem. 2015, 8, 37-47. (20) Hardianto, A.; Yusuf, M.; Liu, F.; Ranganathan, S., Exploration of Charge States of Balanol Analogues Acting as ATP-Competitive Inhibitors in Kinases. BMC Bioinformatics 2017, 18 (Suppl 16), 572. (21) Prlic, A.; Bliven, S.; Rose, P. W.; Bluhm, W. F.; Bizon, C.; Godzik, A.; Bourne, P. E., Pre-Calculated Protein Structure Alignments at the RCSB PDB Website. Bioinformatics 2010, 26, 2983-2985. (22) McWilliam, H.; Li, W.; Uludag, M.; Squizzato, S.; Park, Y. M.; Buso, N.; Cowley, A. P.; Lopez, R., Analysis Tool Web Services from the EMBL-EBI. Nucleic Acids Res. 2013, 41, W597-W600. (23) Taylor, S. S.; Kornev, A. P., Protein Kinases: Evolution of Dynamic Regulatory Proteins. Trends Biochem. Sci. 2011, 36, 65-77. (24) Campanella, J. J.; Bitincka, L.; Smalley, J., MatGAT: An Application That Generates Similarity/identity Matrices Using Protein or DNA Sequences. BMC Bioinformatics 2003, 4, 29. (25) Eswar, N.; Eramian, D.; Webb, B.; Shen, M.-Y.; Sali, A. Protein structure modeling with MODELLER. In Structural Proteomics High-Throughput Methods, Kobe, B.; Guss, M.; Huber, T., Eds.; Humana Press: 2008; Chapter 8. pp. 145–159. (26) Shen, M. Y.; Sali, A., Statistical Potential for Assessment and Prediction of Protein Structures. Protein Sci. 2006, 15, 2507-2524. (27) Laskowski, R. A.; MacArthur, M. W.; Thornton, J. M. PROCHECK: Validation of protein structure coordinates. In International Tables of Crystallography, Volume F. Crystallography of Biological Macromolecules, Rossmann, M. G.; Arnold, E. D., Eds.; Kluwer Academic Publishers: 2001; Vol. F, Chapter 21.4, pp 722-725. (28) UniProt, C., UniProt: a hub for protein information. Nucleic Acids Res. 2015, 43, D204D212. (29) Systèmes, D. Discovery Studio Visualizer, v16.1.0.15350; Dassault Systèmes BIOVIA: San Diego. http://accelrys.com/products/collaborative-science/biovia-discoverystudio/visualization-download.php (accessed Feb 2, 2017). (30) Jakalian, A.; Bush, B. L.; Jack, D. B.; Bayly, C. I., Fast, Efficient Generation of HighQuality Atomic Charges. AM1-BCC Model: I. Method. J. Comput. Chem. 2000, 21, 132-146. (31) Case, D. A.; Betz, R. M.; Botello-Smith, W.; Cerutti, D. S.; Cheatham III, T. E.; Darden, T. A.; Duke, R. E.; Giese, T. J.; Gohlke, H.; Goetz, A. W.; Homeyer, N.; Izadi, S.; Janowski, P.; Kaus, J.; Kovalenko, A.; Lee, T. S.; LeGrand, S.; Li, P.; Lin, C.; Luchko, T.; Luo, R.; Madej, B.; Mermelstein, D.; Merz, K. M.; Monard, G.; Nguyen, H.; Nguyen, H. T.; Omelyan, I.; Onufriev, A.; Roe, D. R.; Roitberg, A.; Sagui, C.; Simmerling, C. L.; Swails, J.; Walker, R. C.; Wang, J.; Wolf, R. M.; Wu, X.; Xiao, L.; York, D. M.; Kollman, P. A. Amber 2016, University of California: San Francisco, 2016. (32) Wang, J. M.; Wolf, R. M.; Caldwell, J. W.; Kollman, P. A.; Case, D. A., Development and Testing of a General Amber Force Field. J. Comput. Chem. 2004, 25, 1157-1174. (33) Maier, J. A.; Martinez, C.; Kasavajhala, K.; Wickstrom, L.; Hauser, K. E.; Simmerling, C., ff14SB: Improving the Accuracy of Protein Side Chain and Backbone Parameters from ff99SB. J. Chem. Inf. Model. 2015, 11, 3696-3713.

ACS Paragon Plus Environment

26

Page 27 of 29 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

585 586 587 588 589 590 591 592 593 594 595 596 597 598 599 600 601 602 603 604 605 606 607 608 609 610 611 612 613 614 615 616 617 618 619 620 621 622 623 624 625 626 627 628 629

Journal of Chemical Information and Modeling

(34) Homeyer, N.; Horn, A. H.; Lanig, H.; Sticht, H., AMBER force-field parameters for phosphorylated amino acids in different protonation states: phosphoserine, phosphothreonine, phosphotyrosine, and phosphohistidine. J. Chem. Theory Comput. 2006, 12, 281-289. (35) Jorgensen, W. L.; Tirado-Rives, J., Potential energy functions for atomic-level simulations of water and organic and biomolecular systems. Proc. Natl. Acad. Sci. U. S. A. 2005, 102, 6665-6670. (36) Fisette, O.; Wingbermuhle, S.; Tampe, R.; Schafer, L. V., Molecular Mechanism of Peptide Editing in the Tapasin-MHC I Complex. Sci. Rep. 2016, 6, 19085. (37) Kumar, A.; Cocco, E.; Atzori, L.; Marrosu, M. G.; Pieroni, E., Structural and Dynamical Insights on HLA-DR2 Complexes That Confer Susceptibility to Multiple Sclerosis in Sardinia: A Molecular Dynamics Simulation Study. PLoS One 2013, 8, e59711. (38) Ryckaert, J.-P.; Ciccotti, G.; Berendsen, H. J. C., Numerical Integration of the Cartesian Equations of Motion of a System with Constraints: Molecular Dynamics of N-Alkanes. J. Comput. Phys. 1977, 23, 327-341. (39) Berendsen, H. J. C.; Postma, J. P. M.; van Gunsteren, W. F.; DiNola, A.; Haak, J. R., Molecular Dynamics with Coupling to an External Bath. J. Chem. Phys. 1984, 81, 3684-3690. (40) Larini, L.; Mannella, R.; Leporini, D., Langevin stabilization of molecular-dynamics simulations of polymers by means of quasisymplectic algorithms. .J. Chem. Phys. 2007, 126, 104101-104101. (41) Gohlke, H.; Case, D. a.; Biology, M.; Scripps, T.; Rd, N. T. P., Converging Free Energy Estimates: MM-PB(GB)SA Studies on the Protein–protein Complex Ras–Raf. J. Comput. Chem. 2004, 25, 238-250. (42) Miller, B. R.; McGee, T. D.; Swails, J. M.; Homeyer, N.; Gohlke, H.; Roitberg, A. E., MMPBSA.py: An efficient program for end-state free energy calculations. J. Chem. Theory Comput. 2012, 8, 3314-3321. (43) Weiser, J.; Shenkin, P. S.; Still, W. C., Approximate Atomic Surfaces from Linear Combinations of Pairwise Overlaps (LCPO). J. Comput. Chem. 1999, 20, 217–230. (44) Poongavanam, V.; Olsen, J. M.; Kongsted, J., Binding free energy based structural dynamics analysis of HIV-1 RT RNase H-inhibitor complexes. Integr. Biol. 2014, 6, 1010-22. (45) Humphrey, W.; Dalke, A.; Schulten, K., VMD: Visual Molecular Dynamics. J. Mol. Graph. 1996, 14, 27-28,33-38. (46) Systèmes, D. BIOVIA Draw 2016, Dassault Systèmes: San Diego http://accelrys.com/products/collaborative-science/biovia-draw/draw-no-fee.php (accessed Feb 2, 2017). (47) R Core Team. R: A language and environment for statistical computing; R Foundation for Statistical Computing. Vienna, Austria, 2015. (48) Wickham, H. ggplot2: Elegant graphics for data analysis. In; Springer New York: 2009, pp 21-54. (49) Rstudio Team. RStudio: Integrated Development for R, 0.99.892; RStudio, Inc.: Boston, MA. http://www.rstudio.com/ (accessed Feb 2, 2017). (50) GIMP. GNU Image Manipulation Program 2.8.14 GPLv3. http://www.gimp.org (accessed Feb 2, 2017). (51) Inkscape. Inkscape 0.48.5 GPLv2. http://www.inkscape.org (accessed Feb 2, 2017). (52) Pflug, A.; de Oliveira, T. M.; Bossemeyer, D.; Engh, R. A., Mutants of Protein Kinase A That Mimic the ATP-Binding Site of Aurora Kinase. Biochem. J. 2011, 440, 85 LP-93.

ACS Paragon Plus Environment

27

Journal of Chemical Information and Modeling 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

630 631 632 633 634

Page 28 of 29

(53) Cenni, V.; Döppler, H.; Sonnenburg, E. D.; Maraldi, N.; Newton, A. C.; Toker, A., Regulation of Novel Protein Kinase C Epsilon by Phosphorylation. Biochem. J. 2002, 363, 537545. (54) Steinberg, S. F., Structural Basis of Protein Kinase C Isoform Function. Physiol. Rev. 2008, 88, 1341-1378.

635 636

ACS Paragon Plus Environment

28

Page 29 of 29 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Chemical Information and Modeling

637 638

Table of Content Graphic

639 640

ACS Paragon Plus Environment

29