You are on page 1of 11

Received: 18 February 2019 

|
  Revised: 9 April 2019 
|  Accepted: 18 April 2019

DOI: 10.1111/cbdd.13536

RESEARCH ARTICLE

Structure‐based drug design and in vitro testing reveal new


inhibitors of enoyl-acyl carrier protein reductases

Mohammad A. Ghattas1   | Nermin A. Eissa1  | Francesca Tessaro2,3  | Remo Perozzo2,3  |


Leonardo Scapozza   2,3
| Dana Obaid   1
| Noor Atatreh 1

1
College of Pharmacy, Al Ain University
of Science and Technology, Abu Dhabi,
Abstract
United Arab Emirates The need for new antibacterial agents is increasingly becoming of great importance
2
Pharmaceutical Biochemistry as bacterial resistance to current drugs is quickly spreading. Enoyl‐acyl carrier pro-
Group, University of Geneva, Geneva,
tein reductases (FabI) are important enzymes for fatty acid biosynthesis in bacte-
Switzerland
3 ria and other micro‐organisms. In this project, we conducted structure‐based virtual
University of Lausanne, Lausanne,
Switzerland screening against the FabI enzyme, and accordingly, 37 compounds were selected for
experimental testing. Interestingly, five compounds were able to demonstrate antimi-
Correspondence
Mohammad A. Ghattas and Noor Atatreh,
crobial effect with variable inhibition activity against various strains of bacteria and
College of Pharmacy, Al Ain University of fungi. Minimum inhibitory concentrations of the active compounds were determined
Science and Technology, P.O. Box 112612 and showed to be in low to medium micromolar range. Subsequently, enzyme inhibi-
Abu Dhabi, United Arab Emirates.
Emails: mohammad.ghattas@aau.ac.ae and tion assay was carried out for our five antimicrobial hits to confirm their biological
noor.atatreh@aau.ac.ae target and determine their IC50 values. Three of these tested compounds exhibited
inhibition activity for the FabI enzyme where our best hit MN02 had an IC50 value
Present address
Nermin A. Eissa, Department of of 7.8 μM. Furthermore, MN02 is a small bisphenolic compound that is predicted
Pharmacology & Therapeutics, College of to have all required features to firmly bind with the target enzyme. To sum up, hits
Medicine & Health Sciences, United Arab
discovered in this work can act as a good starting point for the future development of
Emirates University, Al Ain, United Arab
Emirates new and potent antimicrobial agents.

Funding information KEYWORDS


Al Ain University of Science and antimicrobial agents, bacterial resistance, bisphenol, dichlorophen, diphenylmethane, docking, drug
Technology discovery, enoyl-acyl carrier protein reductase, FabI

1  |   IN T RO D U C T ION wall production. This step is crucial for micro‐organism sur-


vival and cell homeostasis (Heath & Rock, 1995; Marrakchi,
The search for new antimicrobial agents is still of great impor- Zhang, & Rock, 2002; McInnes, 2007).
tance as bacterial resistance to current drugs is increasingly The fatty acid biosynthesis pathway has been considered
spreading (Chitsaz & Brown, 2017; Furusawa, Horinouchi, as increasingly attractive for antimicrobial drug discovery
& Maeda, 2018). Enoyl‐acyl carrier protein reductases (FabI) (Eissa, Farrag, Shawer, & Ammar, 2017; Kronenberger
are associated in non‐covalent complexes with their cofac- et al., 2017; Mehboob et al., 2015; Mistry, Truong, Ghosh,
tor NAD(P)H, and they play an important role in bacterial Johnson, & Mehboob, 2017; Wadhwa, Saha, & Sharma,
survival (Yao & Rock, 2017). FabI reduces fatty acids using 2015). This is mainly because of the major structural differ-
NAD(P)H in order to complete their biosynthesis during cell ences between mammalian enzymes and their corresponding

Ghattas and Eissa authors contributed equally to this work.

Chem Biol Drug Des. 2019;00:1–11. wileyonlinelibrary.com/journal/cbdd © 2019 John Wiley & Sons A/S     1 |
|
2       GHATTAS et al.

microbial isoforms (Massengo‐Tiasse & Cronan, 2009). molecule was observed to be essential for binding of nearly
FabI inhibitors can be classified into three groups accord- all studied FabI inhibitors (Ghattas et al., 2015). The bound
ing to their mechanism of action (Yao & Rock, 2016): The NAD(P) nicotinamide ring was found to be involved in π–π
first group is the bi‐substrate FabI inhibitors that bind co- stacking interactions with the aromatic ring of various FabI
valently to the cofactor NAD(P) to form an adduct complex inhibitors. On the other hand, NAD(P) ribose along with the
(e.g., isoniazid). The second group consists of inhibitors Tyr hydroxyl group was observed to be constantly involved in
that bind with FabI‐NAD(P) complex in a reversible manner hydrogen bonding interactions.
such as triclosan (Figure  1) and its closely related deriva- In this study, we aim to discover new antimicrobial agents
tives (Jiangwei Yao & Rock, 2016). The third group consists via conducting structure‐based virtual screening against the
of inhibitors that reversibly interact with the FabI‐NAD(P) FabI active site based on a previously validated and tested
complex in the presence of the substrate molecule such docking approach (Ghattas, Eissa, Bardaweel, Abu Mellal,
as CG400549 (Yum et  al., 2007) and AFN‐1252 (Zitko & & Atatreh, 2017; Ghattas et  al., 2015). Subsequently, top‐
Doležal, 2016). The latter compound is one of the very few ranked compounds identified by virtual screening and vi-
FabI inhibitors that has reached phase II clinical trial (Zitko sual inspection were tested for their antimicrobial activity
& Doležal, 2016). against a panel of microbial strains. The minimum inhibi-
FabI are reductase enzymes that act on short‐chain fatty tory concentrations were then identified for the most active
acids (Massengo‐Tiasse & Cronan, 2009). They have con- hits. Subsequently, these hits were biologically tested via
served triad in their catalytic pocket (Figure 1) that is com- FabI enzyme assay to confirm their binding with the target
posed of lysine, tyrosine, and phenylalanine (or tyrosine). protein.
The lysine residue has a role in stabilizing the binding of the
cofactor, while Tyr is actively involved in the redox catalytic
mechanism by donating a proton to the substrate (Parikh,
2  |  EXPERIM ENTAL
Moynihan, Xiao, & Tonge, 1999). The fatty acyl U‐shaped
2.1  |  Protein and ligand library preparation
conformation is directed by the third conserved residue,
that is Phe or Tyr, which is positioned nearby the NAD(P) Escherichia coli FabI (ecFabI) crystal structure was ob-
nicotinamide (Rozwarski, Vilchèze, Sugantino, Bittman, & tained from the protein data bank (http://www.rcsb.org/),
Sacchettini, 1999). Out of these conserved residues, only the PDB: 1QSG (Stewart, Parikh, Xiao, Tonge, & Kisker, 1999).
catalytic Tyr is involved in binding the FabI ligands (Ghattas, Solvent atoms were removed; then, MOE protein prepara-
Mansour, Atatreh, & Bryce, 2015). According to previ- tion module (MOE, 2016)b was used to check out the pro-
ous research, this conserved Tyr in addition to the cofactor tein‐cofactor complex for any missing atoms or residues and
correct them accordingly (Halgren, 1996). Then, a previously
validated hydrogen bonding constraint (Ghattas et al., 2015)
was created on the NAD ribose 2‐hydroxyl group using the
protein preparation module in Maestro (Schrödinger Suite,
2015)d. Triclosan bound conformation was used to identify
the active site on the FabI‐NAD complex. Another FabI crys-
tal structure brought from staphylococcus aureus (saFabI),
PDB: 4ALI (Schiebel et  al., 2012), was used for analysis
purposes and has been prepared using the above‐mentioned
procedure.
The ligand library of National Cancer Institute/USA
(www.cancer.gov) was used for this screening and was fil-
tered based on the Veber's rules (Veber et  al., 2002) and
the Lipinski's rule of five (Lipinski, Lombardo, Dominy,
& Feeney, 1997). Hence, filtered ligand library had ligands
with the following criteria: rotatable bonds ≤10, logP ≤5,
hydrogen bond donor (HBD) ≤5, hydrogen bond acceptor
(HBA) ≤10, molecular weight ≤500 and polar surface area
(PSA) ≤140. After filtration, ligand library size was re-
F I G U R E 1   The FabI active site (yellow cartoon) along with the duced from 273,885 to 191,929 ligands. Ligands were then
catalytic triad (yellow sticks), the NAD cofactor (orange sticks), and assigned the appropriate protonation states and tautomer
the standard FabI inhibitor triclosan (blue sticks) forms via LigPrep module in Maestro (Schrödinger Release,
2015)c
GHATTAS et al.   
   3
|

Star ng Database of 273,885 Ligands


Filtraon

Filtered Database of 191,929 Ligands

Const. Docking via


Drug-like characteristics:
Rotatable bond ≤ 10
Top 20,000 ranked List GLIDE-HTVS
Molecular weight ≤ 500
LogP ≤ 5.0 Const Docking via
PSA ≤ 140 Top 2000 ranked list GLIDE-SP
HB donor ≤ 5 acceptor ≤ 10

Select Const. Docking via


37 GLIDE-XP

5 hits
In vitro Evaluaon

F I G U R E 2   Virtual screening protocol used for the discovery of new FabI inhibitors. PSA: polar surface area; HB: hydrogen bond

(Small‐Molecule Drug Discovery Suite, 2015)e. A three‐step


2.2  |  Virtual screening docking protocol
docking protocol was employed to allow gradual increase
Constrained docking into ecFabI enzyme active site for the in docking precision: High‐throughput virtual screening
prepared drug‐like ligand library was employed using GLIDE (HTVS), standard precision (SP), and then extra‐precision

T A B L E 1   The antimicrobial activity of the 37 compounds tested against various strains of bacterial and fungal organisms, using the disk
diffusion test

Zone of Inhibition (mm)

S. aureus B. subtilis E. coli C. albicans


Compounds (ATCC259230) MRSA (ATCC3359) (ATCC6633) (ATCC25922) (ATCC10231)
MN01 − NT NT − NT
MN02 +++ +++ +++ + +
MN03 – MN04 − NT NT − NT
MN05 +++ +++ +++ − +
MN06 – MN13 − NT NT − NT
MN14 +++ +++ ++ − −
MN15 – MN31 − NT NT − NT
MN32 + − + − −
MN33 – MN35 − NT NT − NT
MN36 + +++ + − +
MN37 − NT NT − NT
chloramphenicol +++ ++ +++ +++ NT
Ketoconazole NT NT NT NT +++
Note: Key to symbols: − inactive (inhibition zone < 6 mm); slightly active = + (inhibition zone 7–9 mm); moderately active = + + (inhibition zone 10–13 mm);
highly active = + + + (inhibition zone > 14 mm) (NT= not tested).
|
4       GHATTAS et al.

T A B L E 2   The MIC values (μg/ml) of


MIC (μg/ml)
our five hits along with the positive standard
S. aureus MRSA B. subtilis E. coli tested against different bacterial strains
Compounds (ATCC259230) (ATCC3359) (ATCC6633) (ATCC25922)
MN02 4.0 8.0 4.0 32.0
MN05 13.9 13.9 13.9 NT
MN14 15.9 15.9 15.9 NT
MN32 66.7 NT 66.7 NT
MN36 NT 41.2 NT NT
Chloramphenicol 4.0 74.0 4.0 4.0
a a a
Triclosan 0.125 64.0 0.4 0.5a
Abbreviation: NT, not tested.
a
MIC values obtained from literature.

(XP) docking algorithms. In GLIDE_XP, flexible ligand to dry at room temperature in order to remove any residual
sampling was enabled, taking into account nitrogen inver- solvent. A sterile cotton swab was dipped into the inoculum
sions and ring conformations. Post‐docking minimization suspension that was adjusted to 0.5 McFarland standard tur-
was performed for top‐ten poses with a threshold value of bidity, and lawns were made on nutrient agar (S.  aureus,
0.5  kcal/mol for rejecting minimized poses. Afterward, all MRSA, B. subtilis, and E. coli) and sabouraud dextrose agar
produced poses were scored and ranked via the Glide_XP (C. albicans). Disks were then placed carefully on the agar
scoring function which takes accounts of several terms (i.e., plates that had been already inoculated with micro‐organ-
van der Waals, hydrogen bond, electrostatic interactions, de- isms. The plates were incubated at 37°C for 24 hr (bacteria)
solvation penalty, and penalty for intra‐ligand contact). The and 30°C for 48 hr (fungi). After incubation, the diameter of
2000 ligands docked and ranked by GLIDE‐XP were then inhibition zones was measured to the nearest whole millim-
clustered based on MACCS algorithm (MDL, 2015)a and eter. Chloramphenicol (Sigma) and ketoconazole (Bio‐rad,
were then visually inspected. Finally, 37 ligands belonging to Marnes‐la‐Coquette, France) were used as reference antibac-
various structural scaffolds and showing convenient binding terial and antifungal agents, respectively. All experiments
modes in the ecFabI active site were selected for antibacterial were made in duplicate.
testing. Minimum inhibitory concentrations (MICs) for the most
active compounds MN02, MN05, MN14, and MN36 against
E.  coli, S.  aureus, MRSA, and B.  subtilis (where applica-
2.3  |  Antimicrobial testing
ble) were determined by using the broth dilution method
Antimicrobial susceptibility testing of 37 compounds se-
lected from virtual screening was performed. Screening
was initially conducted by a disk diffusion test against the 150
Gram‐positive micro‐organism Staphylococcus aureus and
Remaining activity %

the Gram‐negative bacteria Escherichia coli. Compounds 97.4% 100.0%


showed growth inhibition were further evaluated for antibac- 100
terial activity against a selected range of bacterial strains and
fungi, including methicillin‐resistant Staphylococcus aureus
(MRSA, ATCC33591), Bacillus subtilis (ATCC6633), and 45.7%
50
Candida albicans (ATCC10231). All strains were purchased
from Remel Europe (Datford, UK). Preliminary screening
was carried out by using disk diffusion method as per Clinical 0
Laboratory Standards Institute (CLSI, 2006, 2012) and litera-
µM
µM

µM

µM

µM

ture (Reller, Weinstein, Jorgensen, & Ferraro, 2009) where


0
0

10
10

10

10

10

10

nutrient agar medium (LABM, Lancashire, UK) was used


L
02

05

14

32

36

TC
N

for bacteria and sabouraud dextrose agar medium (LABM,


M

Lancashire, UK) was used for fungi. Sterile filter paper disks F I G U R E 3   Inhibition activity of the tested compounds shown as
(6 mm, sigma, st. Louis, MO or Whatman Number 1) were remaining activity expressed in percentage of MN02, MN05, MN14,
soaked in the test compounds dissolved in dimethyl sulfoxide MN36, and triclosan. The standard deviation bars for MN0, MN32,
(20 mM, Sisco Research Lab., Mumbai, India) and were left and triclosan were lower than 0.1% thus not visible in the graph
GHATTAS et al.   
|
   5

150
IC50 = 27.4 ± 1.3 µM 150
IC50 = 7.8 ± 1.2 µM 150 IC50 = 2.7 ± 0.8 µM

100 100 100

Activity
Activity
Activity

50 50
50

0 0
0 –2 –1 0 1 2 3
–4 –2 0 2 4
–4 –2 0 2 4
Conc [Log]
Conc [Log] –50
Conc [Log] –50
–50

MN14 MN02 TCL

F I G U R E 4   Dose–response curve of molecules MN02 and MN14 as well as the positive control triclosan. The values reported in the graph
represent the mean values of a triplicate

(Wiegand, Hilpert, & Hancock, 2008). Tested compounds each strain. That was detected by unaided eye after overnight
were dissolved in dimethyl sulfoxide where twofold dilutions incubation.
of solution were prepared and then were diluted in culture
medium (Nutrient broth; LABM, Lancashire, UK) to obtain
2.4  |  Enzymatic assay against saFabI
a final concentration range of 0.56–256 μg/ml. The dimethyl
sulfoxide final concentration never exceeded 1% v/v. Each Enzyme activity at a fixed concentration of compound was
test‐tube was inoculated with micro‐organism suspensions assessed using spectrophotometric method at 25°C. That was
at 1 × 106 CFU/ml (colony forming unit/ml) to obtain final done by monitoring the decrease in NADPH absorbance at
desired inoculum of 5 × 105 CFU/ml. MICs were read after 340 nm in the presence of trans‐2‐decenoyl‐n‐acetylcysteam-
incubation at 37°C for 24  hr (bacteria). Growth controls ine substrate during 2 min (Slepikas et al., 2016). The standard
consisting of media with 1% v/v dimethyl sulfoxide were experimental conditions in a total volume of 1 ml consisted of
employed. Chloramphenicol was used as a reference for an- assay buffer (20 mM Hepes, 150 mM NaCl, pH 7.4), 9.6 μg
tibacterial activity. All experiments were repeated two times. of purified recombinant SaFabI, 200 μM of NADPH, 100 μM
MIC values were determined as the lowest concentrations of of substrate, and 100 μM of compound (20 mM stock solu-
compound that completely inhibited the visible growth of tion in DMSO). IC50 values were determined using the same

T A B L E 3   The GLIDE‐XP docking scores of our hits obtained from docking into two ENR enzymes, ecFabI and saFabI, along with their
ligand efficiency

Docking into ecFabI (kcal/mol) Docking into saFabI (kcal/mol)

Compound code Structure GLIDE‐XP score Ligand efficiencya GLIDE‐XP score Ligand efficiencya
MN02 −9.24 −0.54 −9.14 −0.52

MN05 −8.07 −0.47 −6.87 −0.40

MN14 −7.88 −0.39 −8.56 −0.43

MN32 −7.82 −0.37 −6.96 −0.32

MN36 −7.97 −0.35 −8.46 −0.37

triclosan −8.15 −0.47 −9.20 −0.51

a
Ligand efficiency is calculated via dividing docking score over molecular weight.
|
6       GHATTAS et al.

F I G U R E 5   The best predicted pose of


MN02 (pink sticks) docked into the active
site of saFabI (a) and ecFabI (b) where
the surrounding residues (including the
cofactor) are shown as green or blue sticks,
respectively. Interactions are shown as blue
dotted lines

method but by varying the compound concentration. DMSO algorithm (GLIDE‐XP) in the Schrodinger software package.
was kept constant at 0.5%. Data were analyzed by non‐lin- The resultant docked poses of these 2000 ligand were visu-
ear fit to the sigmoidal dose–response curve using GraphPad ally inspected inside the ecFabI binding pocket. Further, 37
Prism software. All measurements were carried out in tripli- compounds from various chemical clusters were selected for
cate from the same batch of protein. antibacterial testing based on their ligand–protein interaction
profiles and according to their fitting into the catalytic pocket
of the target enzyme.
3  |  R ES U LTS A N D D IS C U S S ION The best hits selected from virtual screening were screened
initially against the E. coli and S. aureus strains using disk
Constrained docking has reportedly been showing en- diffusion test. Amongst the 37 tested compounds, five hits
hanced success rates of virtual screening in many protein (i.e., MN02, MN05, MN14, MN32, and MN36) showed clear
families, such as protein tyrosine phosphatases and kinases inhibition zones in at least one of the cultured bacterial plates,
(Ghattas, Atatreh, Bichenkova, & Bryce, 2014; Perola, whereas the rest of compounds revealed no significant inhibi-
2006). Similarly, a previous study showed that constrain- tion. Those five hits were further screened against two more
ing interaction, namely to the NAD(P) ribose 2‐hydroxyl bacterial strains (MRSA and B. subtilis) and one fungal strain
group, can significantly increase hit rates and improve (C. albicans) to check their spectrum of activity. As shown in
docking accuracy for FabI virtual screens (Ghattas et  al., Table 1, MN02 and MN05 showed high inhibition activity in
2015). Accordingly, hydrogen bonding constraint was also particular against S. aureus, MRSA, and B. subtilis along with
applied in this work to raise our chance for discovering weak inhibition activity against C.  albicans. However, the
new FabI inhibitors. growth of the Gram‐negative bacteria (i.e., E. coli) was only
The virtual screening protocol employed here is schemat- inhibited by the former compound (MN02) which showed the
ically shown in Figure  2. The NCI ligand library was first broadest spectrum of activity amongst all tested compounds.
filtered for druglikeness before docking into the ecFabI ac- MN14 showed similar bacterial inhibition as MN02
tive site, reducing the library size from 273,885 ligands to and MN05 against S.  aureus and MRSA with lesser activ-
191,929 ligands. The drug‐like library was initially screened ity against B. subtilis and none against C. albicans. On the
using the GLIDE‐HTVS docking algorithm. The top 20,000 other hand, MN36 exhibited similar inhibition activity as
ligands were re‐docked in a slower but more accurate dock- MN02 against MRSA and C. albicans but was less effective
ing algorithm using GLIDE‐SP. Finally, docking results were against S. aureus and B. subtilis. Finally, MN32 exhibited the
enhanced using the extra‐precision docking and scoring weakest antimicrobial activity amongst the five compounds

F I G U R E 6   The best predicted pose of


MN05 (pink sticks) docked into the active
site of saFabI (a) and ecFabI (b) where
the surrounding residues (including the
cofactor) are shown as green or blue sticks,
respectively. Interactions are shown as blue
dotted lines
GHATTAS et al.   
   7
|
F I G U R E 7   The best predicted pose of
MN14 (pink sticks) docked into the active
site of saFabI (a) and ecFabI (b) where
the surrounding residues (including the
cofactor) are shown as green or blue sticks,
respectively. Interactions are shown as blue
dotted lines

demonstrating weak inhibition activity against S. aureus and significant inhibition for the target enzyme. Triclosan was
B. subtilis with no observed activity against E. coli or C. al- used as a positive control, and it showed full inhibition at
bicans. In conclusion, all discovered hits exhibited more pro- 100 μM.
nounced activity against Gram‐positive bacteria compared The three hits with most significant saFabI inhibition
with Gram‐negative and fungal strains. were undergone further screening to determine their IC50.
Active hits obtained from disk diffusion testing were fur- As shown in Figure 4, while MN14 exhibited an IC50 value
ther screened to determine their minimum inhibitory concen- of 27.4  μM, MN02 showed enzyme inhibition activity in
tration (MIC) using the broth dilution method. As shown in the low micromolar range with an IC50 value of 7.8  μM,
Table 2, amongst the five tested compounds, MN02 exhibited which is comparable to the standard inhibitor triclosan
the best antibacterial potency with MIC values ranging from (IC50 = 2.7 μM). The IC50 value of MN05 could not be mea-
low to intermediate micromolar range, which is compara- sured because of solubility issues related to the compound;
ble to antibacterial activity of chloramphenicol. In addition however, since MN05 was able to inhibit saFabI activity by
to showing the best inhibitory activity, MN02 demonstrated 45.7% at the 100 μM concentration, its IC50 value is estimated
the best spectrum of activity against the tested panel of to be around 100 μM. In summary, MN02, MN05, and MN14
bacterial strains, being more active though against Gram‐ have been all confirmed to exert their antimicrobial activity
positive (MIC  =  4–8  μg/ml) than Gram‐negative bacteria via inhibiting the FabI enzyme, while MN32 and MN36 seem
(MIC = 32 μg/ml). to act through other routes.
The other two hits, MN05 and MN14, showed anti- To explain and correlate in vitro results with the in silico
bacterial activity in the middle micromolar range against data, docking results of the discovered hits were studied and
all Gram‐positive bacterial strains, obtaining MIC values analyzed. Looking at Table 3, MN02 was predicted to have
of 14 and 16  μg/ml, respectively. The final two inhibi- the best GLIDE_XP docking scores for both FabI enzymes,
tors, MN32 and MN36, had weaker inhibition activities E.  coli (−9.24  kcal/mol) and S.  aureus (−9.14  kcal/mol),
than previous compounds, exhibiting MIC values in the scoring at least as good as reference ligand triclosan (−8.15
range of 40–67  μg/ml. Most interestingly, four out of to 9.20 kcal/mol). Interestingly, MN02's docking data seem
the five discovered hits (i.e., MN02, MN05, MN14, and to be consistent with the in vitro findings that recognize it
MN36) demonstrated better inhibition activities against as the most active hit. The other four hits exhibited conve-
the penicillin‐resistant strain MRSA than chloramphenicol nient docking scores with comparable or higher binding en-
(MIC = 74 μg/ml) and triclosan, MIC reported as 64 μg/ml ergies than the reference ligand, in term of how these ligands
(Regös, Zak, Solf, Vischer, & Weirich, 1979). To sum up, are scoring as per their size, as shown in Table 3. Our best
all discovered hits exhibited an interesting anti‐Gram‐posi- hit MN02 was also able to obtain the best ligand efficiency
tive activity (particularly with regard to their MRSA activ- score (−0.5  kcal/mol), and relatively similar scoring to tri-
ity), highlighting MN02 as the most potent antimicrobial closan (−0.47 to −0.51 kcal/mol). This suggests that MN02
agent with the broadest spectrum of activity. possesses lead‐like characteristics and offers possibility for
The five active compounds were further screened against structural modifications and improvement future work.
enzymatic inhibition assay (Figure  3) in order to confirm Considering the docking pose, MN02 forms all key inter-
whether or not they exert their antibacterial activity through actions with the cofactor and the catalytic Tyr residue. It had
inhibiting FabI. Amongst the five tested compounds, MN02 its hydroxyl group simultaneously interacting via hydrogen
and MN14 showed complete inhibition for saFabI enzyme at bonding with the NAD ribose and the Tyr156 side chain, in
100 μM, while MN05 exhibited a lesser inhibition percent- addition to having its phenyl group forming π–π interaction
age of 45% at the aforementioned concentration (Figure 3). with the NAD nicotinamide ring (Figure  5). Interestingly,
The other two compounds, MN32 and MN36, exhibited no MN02 bears structural similarity with triclosan which could
|
8       GHATTAS et al.

T A B L E 4   The GLIDE_XP docking


GLIDE‐XP Score (kcal/
scores of dichlorophen and other derivatives
mol)
into ecFabI and saFabI
Compound code Structure ecFabI saFabI
Resorcinol sulfide −7.52 −7.62

Dichlorophen −5.81 −6.66

Bromochlorophen −5.41 −5.66

Bithionol −5.00 −5.26

Hexachlorophene −4.43 −5.07

MN02 −9.24 −9.14

explain why both ligands possess similar binding modes and a very similar binding mode to MN02 and triclosan, form-
antimicrobial activity profiles (Stewart et al., 1999). ing the two necessary hydrogen bonding with the NAD ri-
MN05 belongs to the monohydroxydiphenylmethane class bose and Tyr156 (Figure 6). However, unlike the latter two
of compounds that are known for their bactericidal activity ligands, MN05 does not make the commonly seen π‐π inter-
(Florestano & Bahler, 1953; Klarmann, Gates, & Shternov, action with the NAD nicotinamide ring. That is probably due
1932). However, bacterial targets for such compounds have to steric hindrance caused by the meta substitution of MN05,
never been identified and it is the first time to report FabI as which is slightly pushing away the aromatic moiety from the
a drug target for them. In terms of binding, MN05 possesses deepest point in the FabI binding cavity.

(a) (b)

F I G U R E 8   (a) The best predicted pose of dichlorophen (red) along with its para derivative MN02 (pink). (b) The best predicted pose of
bithionol (brown) along with its para derivative resorcinol sulfide (green). The reference ligand triclosan is shown as blue sticks in both pictures,
and the active site of the ecFabI‐NAD complex is shown as blue ribbon
GHATTAS et al.   
   9
|
MN14 is another hit confirmed to exert its antibacterial exhibited the broadest spectrum of activity, eradicating
action via inhibiting the FabI enzyme. As shown in Figure 7, both Gram‐positive and Gram‐negative bacteria. Three
MN14 forms two hydrogen bonds with the NAD ribose via out of the discovered hits were confirmed to bind and in-
its phenol and ketone functionalities. Additionally, the phenol hibit the FabI enzyme, showing IC50 values in the low to
group was shown to make an extra hydrogen bonding with medium micromolar range. These three active compounds
the Tyr156 side chain. MN14 was also predicted to form the showed convenient docking binding modes inside the FabI
usual interaction with NAD nicotinamide ring in addition to catalytic pocket, satisfying the key interactions with the
several hydrogen–π interactions and vdW contacts with the bound cofactor as well as the catalytic Tyr residue. The
surrounding residues. It is worth mentioning that interaction obtained hits, particularly MN02, look promising leads for
between the ketone group and the NAD ribose was only ob- future development of antibacterial agents.
served in MN14 binding with ecFabI. However, this interac-
tion was compensated in saFabI by an alternative hydrogen
ACKNOWLEDGMENTS
bonding observed between Ser197 and the distant phenol
group of the compound. This work was funded by Deanship of Scientific Research
In fact, MN02 is structurally close to the known dichloro- and Graduate Studies at Al Ain University of Science and
phen with the only difference in the position of chloro substi- Technology, Al Ain, United Arab Emirates. The authors ac-
tutions. Dichlorophen is a compound used in many veterinary knowledge National Cancer Institute, USA, for providing the
preparations (Gilvydis, 1972) as anthelmintic agent and an- compounds used in this work. We would like to thank Dr
tiprotozoal agent (Kawasaki & Takeuchi, 1984; Takeuchi, Nezar Al Bataineh for proofreading the manuscript.
Kobayashi, Tanabe, & Fujiwara, 1985). As the protein tar-
get of dichlorophen and its derivatives (i.e., bithionol, bro-
mochlorophen, resorcinol sulfide, and hexachlorophene) has CONFLICT OF INTEREST
never been reported before, it would be interesting to check The authors declare that no conflict of interest is associated
the binding potential of these compounds via docking them with this work.
into the FabI active site. As shown in Table  4, all of these
compounds have similar structures to MN02; however, none
of them (except resorcinol) were able to score as low docking ORCID
score as MN02.
Mohammad A. Ghattas  https://orcid.
Figure 8 shows the binding mode of bithionol and dichlo-
org/0000-0002-2240-8037
rophen along with their para‐substituted derivatives (i.e., re-
sorcinol sulfide and MN02, respectively). It is evident that
the usual interaction with the cofactor and the catalytic Tyr ENDNOTES
residue is hindered by the meta substitution on bithionol and 1
MDL. (2015). MACCS Keys, MDL Information Systems Inc., 14600
dichlorophen. This explains why such structurally similar
Catalina Street, San Leandro, CA 94577.
compounds could not fit as good as MN02 in the FabI binding 2

MOE. (2016). manual version 2016.09, Molecular Operating
site. Hence, it can be suggested that bisphenolic compounds
Environment (MOE), Chemical Computing Group, http://www.chemc​
with para substitutions have more potential to inhibit the FabI
omp.com, Montreal, Canada.
enzyme, while those with meta substitutions are less likely 3
Schrödinger Release 2015‐2: LigPrep, version 3.4, Schrödinger, LLC,
to do so and they could rather exert their antimicrobial ef-
New York, NY, 2015.
fect via binding to other targets in the fatty acid biosynthesis 4
Schrödinger Suite 2015‐2 Protein Preparation Wizard; Epik version 3.2,
pathway, such as 3‐oxoacyl‐ACP reductase (Wickramasinghe
Schrödinger, LLC, New York, NY, 2015; Impact version 6.7, Schrödinger,
et al., 2006). LLC, New York, NY, 2015; Prime version 4.0, Schrödinger, LLC, New
York, NY, 2015.
5
Small‐Molecule Drug Discovery Suite 2015‐2: Glide, version 6.7,
4  |   CO NC LU SION S Schrödinger, LLC, New York, NY, 2015.

A validated structure‐based virtual screening has been car-


ried out against the FabI catalytic pocket. Consequently, R E F E R E NC E S
a total number of 37 hits were selected for experimental
Chitsaz, M., & Brown, M. H. (2017). The role played by drug efflux
testing, five of which showed antibacterial activity particu- pumps in bacterial multidrug resistance. Essays In Biochemistry,
larly against Gram‐positive strains. The obtained hits ex- 61(1), 127–139. https​://doi.org/10.1042/ebc20​160064
hibited MIC values in the micromolar range, showing very Clinical Laboratory Standards Institute (CLSI) (2006). Performance
interesting activity against MRSA. Our best hit, MN02, standards for antimicrobial disk susceptibility tests; Approved
|
10       GHATTAS et al.

standard‐9th ed. CLSI document M2‐A9. 26:1. Wayne, PA: Clinical Lipinski, C. A., Lombardo, F., Dominy, B. W., & Feeney, P. J. (1997).
Laboratory Standards Institute. Experimental and computational approaches to estimate solubil-
Clinical Laboratory Standards Institute (CLSI) (2012). Methods for di- ity and permeability in drug discovery and development settings.
lution antimicrobial susceptibility tests for bacteria that grow aer- Advanced Drug Delivery Reviews, 23(1–3), 3–25.
obically; approved standard—9th edition. CLSI document M07–A9 Marrakchi, H., Zhang, Y.‐M., & Rock, C. O. (2002). Mechanistic di-
32: 2. Wayne, PA: Clinical Laboratory Standards Institute. versity and regulation of Type II fatty acid synthesis. Biochemical
Eissa, S. I., Farrag, A. M., Shawer, T. Z., & Ammar, Y. A. (2017). Society Transactions, 30, 1050–1055.
Design, synthesis, 3D pharmacophore, QSAR, and docking stud- Massengo‐Tiasse, R. P., & Cronan, J. E. (2009). Diversity in enoyl‐acyl
ies of some new (6‐methoxy‐2‐naphthyl) propanamide derivatives carrier protein reductases. Cellular and Molecular Life Sciences,
with expected anti‐bacterial activity as FABI inhibitor. Medicinal 66(9), 1507–1517. https​://doi.org/10.1007/s00018-009-8704-7
Chemistry Research, 26(10), 2375–2398. https​://doi.org/10.1007/ McInnes, C. (2007). Virtual screening strategies in drug discovery.
s00044-017-1939-1 Current Opinion in Chemical Biology, 11(5), 494–502.
Florestano, H. J., & Bahler, M. E. (1953). Antibacterial activity of a se- Mehboob, S., Song, J., Hevener, K. E., Su, P.‐C., Boci, T., Brubaker, L.,
ries of diphenylmethanes. Journal of the American Pharmaceutical … Johnson, M. E. (2015). Structural and biological evaluation of
Association, 42(9), 576–578. https​://doi.org/doi:10.1002/jps.30304​ a novel series of benzimidazole inhibitors of Francisella tularensis
20916​ enoyl‐ACP reductase (FabI). Bioorganic and Medicinal Chemistry
Furusawa, C., Horinouchi, T., & Maeda, T. (2018). Toward pre- Letters, 25(6), 1292–1296.
diction and control of antibiotic‐resistance evolution. Current Mistry, T. L., Truong, L., Ghosh, A. K., Johnson, M. E., & Mehboob,
Opinion in Biotechnology, 54, 45–49. https​ ://doi.org/10.1016/j. S. (2017). Benzimidazole‐based FabI inhibitors: A promising
copbio.2018.01.026 novel scaffold for anti‐staphylococcal drug development. ACS
Ghattas, M. A., Atatreh, N., Bichenkova, E. V., & Bryce, R. A. (2014). Infectious Diseases, 3(1), 54–61. https​ ://doi.org/10.1021/acsin​
Protein tyrosine phosphatases: Ligand interaction analysis and op- fecdis.6b00123
timisation of virtual screening. Journal of Molecular Graphics and Parikh, S., Moynihan, D. P., Xiao, G., & Tonge, P. J. (1999). Roles of
Modelling, 52, 114–123. tyrosine 158 and lysine 165 in the catalytic mechanism of InhA,
Ghattas, M. A., Eissa, N. A., Bardaweel, S. K., Abu Mellal, A., & the Enoyl‐ACP reductase from Mycobacterium tuberculosis.
Atatreh, N. (2017). Computer‐aided discovery of antimicrobial Biochemistry, 38(41), 13623–13634. https​://doi.org/10.1021/bi990​
agents as potential enoyl-acyl carrier protein reductase inhibitors. 529c
Tropical Journal of Pharmaceutical Research, 16(2), 397–405. Perola, E. (2006). Minimizing false positives in kinase virtual screens.
Ghattas, M. A., Mansour, R. A., Atatreh, N., & Bryce, R. A. (2015). Proteins: Structure, Function, and Bioinformatics, 64(2), 422–435.
Analysis of Enoyl‐Acyl carrier protein reductase structure and inter- https​://doi.org/10.1002/prot.21002​
actions yields an efficient virtual screening approach and suggests Regös, J., Zak, O., Solf, R., Vischer, W. A., & Weirich, E. G. (1979).
a potential allosteric site. Chemical Biology & Drug Design, 87(1), Antimicrobial spectrum of triclosan, a broad‐spectrum antimicro-
131–142. https​://doi.org/10.1111/cbdd.12635​ bial agent for topical application. Dermatologica, 158, 72–79.
Gilvydis, D. M. (1972). Collaborative study of the determination of Reller, L. B., Weinstein, M., Jorgensen, J. H., & Ferraro, M. J. (2009).
dichlorophene in veterinary preparations by ultraviolet spectropho- Antimicrobial susceptibility testing: A review of general principles
tometry. Journal of the Association of Official Analytical Chemists, and contemporary practices. Clinical Infectious Diseases, 49(11),
55(1), 163–165. 1749–1755.
Halgren, T. A. (1996). Merck molecular force field. V. Extension of Rozwarski, D. A., Vilchèze, C., Sugantino, M., Bittman, R., &
MMFF94 using experimental data, additional computational data, Sacchettini, J. C. (1999). Crystal structure of the Mycobacterium
and empirical rules. Journal of Computational Chemistry, 17(5–6), tuberculosis Enoyl‐ACP reductase, InhA, in complex with NAD+
616–641. https​://doi.org/10.1002/(SICI)1096-987X(19960​4)17:5/6 and a C16 Fatty Acyl substrate. Journal of Biological Chemistry,
<616: AID-JCC5>3.0.CO;2-X 274(22), 15582–15589. https​://doi.org/10.1074/jbc.274.22.15582​
Heath, R. J., & Rock, C. O. (1995). Enoyl‐Acyl carrier protein reductase Schiebel, J., Chang, A., Lu, H., Baxter Michael, V., Tonge Peter, J., &
(fabI) plays a determinant role in completing cycles of fatty acid Kisker, C. (2012). Staphylococcus aureus Fab I: Inhibition, sub-
elongation in Escherichia coli. Journal of Biological Chemistry, strate recognition, and potential implications for in vivo essentiality.
270(44), 26538–26542. https​://doi.org/10.1074/jbc.270.44.26538​ Structure, 20(5), 802–813. https​://doi.org/10.1016/j.str.2012.03.013
Kawasaki, H., & Takeuchi, T. (1984). Entamoeba histolytica: Effects of Slepikas, L., Chiriano, G., Perozzo, R., Tardy, S., Kranjc, A.,
dichlorophene and hexachlorophene on respiratory activities, growth Patthey‐Vuadens, O., … Scapozza, L. (2016). In silico driven
in vitro and ultrastructure. Japanese Journal of Parasitology, 33(6), design and synthesis of rhodanine derivatives as novel antibac-
545–554. terials targeting the enoyl reductase InhA. Journal of Medicinal
Klarmann, E., Gates, L. W., & Shternov, V. A. (1932). Halogen deriv- Chemistry, 59(24), 10917–10928. https​ ://doi.org/10.1021/acs.
atives of monohydroxydiphenylmethane and their antibacterial ac- jmedc​hem.5b01620
tion. Journal of the American Chemical Society, 54(8), 3315–3328. Stewart, M. J., Parikh, S., Xiao, G., Tonge, P. J., & Kisker, C. (1999).
https​://doi.org/10.1021/ja013​47a045 Structural basis and mechanism of enoyl reductase inhibition by tri-
Kronenberger, T., Asse, L. R., Wrenger, C., Trossini, G. H. G., Honorio, closan. Journal of Molecular Biology, 290(4), 859–865. https​://doi.
K. M., & Maltarollo, V. G. A. (2017). Studies of Staphylococcus org/10.1006/jmbi.1999.2907
aureus FabI inhibitors: Fragment‐based approach based on holo- Takeuchi, T., Kobayashi, S., Tanabe, M., & Fujiwara, T. (1985). In vitro
graphic structure‐activity relationship analyses. Future Medicinal inhibition of Giardia lamblia and Trichomonas vaginalis growth
Chemistry, 9(2), 135–151. https​://doi.org/10.4155/fmc-2016-0179 by bithionol, dichlorophene, and hexachlorophene. Antimicrobial
GHATTAS et al.   
   11
|
Agents and Chemotherapy, 27(1), 65–70. https​://doi.org/10.1128/ Yao, J., & Rock, C. O. (2017). Bacterial fatty acid metabolism in mod-
aac.27.1.65 ern antibiotic discovery. Biochimica et Biophysica Acta ‐ Molecular
Veber, D. F., Johnson, S. R., Cheng, H.‐Y., Smith, B. R., Ward, K. W., & and Cell Biology of Lipids, 1862(11), 1300–1309.
Kopple, K. D. (2002). Molecular properties that influence the oral Yum, J. H., Kim, C. K., Yong, D., Lee, K., Chong, Y., Kim, C. M., …
bioavailability of drug candidates. Journal of Medicinal Chemistry, Cho, J. M. (2007). In vitro activities of CG400549, a novel FabI
45(12), 2615–2623. https​://doi.org/doi:10.1021/jm020​017n inhibitor, against recently isolated clinical staphylococcal strains in
Wadhwa, P., Saha, D., & Sharma, A. (2015). Combined 3D‐QSAR and Korea. Antimicrobial Agents and Chemotherapy, 51(7), 2591–2593.
molecular docking study for identification of diverse natural prod- https​://doi.org/10.1128/aac.01562-06
ucts as potent Pf ENR inhibitors. Current Computer‐Aided Drug Zitko, J., & Doležal, M. (2016). Enoyl-acyl carrier protein reduc-
Design, 11(3), 245–257. tase inhibitors: An updated patent review (2011–2015). Expert
Wickramasinghe, S. R., Inglis, K. A., Urch, J. E., Müller, S., Aalten, D. Opinion on Therapeutic Patents, 26(9), 1079–1094. https​ ://doi.
M. F. V., & Fairlamb, A. H. (2006). Kinetic, inhibition and structural org/10.1080/13543​776.2016.1211112
studies on 3‐oxoacyl‐ACP reductase from Plasmodium falciparum,
a key enzyme in fatty acid biosynthesis. Biochemical Journal,
393(2), 447–457. https​://doi.org/10.1042/bj200​50832​ How to cite this article: Ghattas MA, Eissa NA,
Wiegand, I., Hilpert, K., & Hancock, R. E. (2008). Agar and broth di- Tessaro F, et al. Structure‐based drug design and in
lution methods to determine the minimal inhibitory concentration vitro testing reveal new inhibitors of enoyl-acyl carrier
(MIC) of antimicrobial substances. Nature Protocols, 3(2), 163–
protein reductases. Chem Biol Drug Des. 2019;00:
175. https​://doi.org/10.1038/nprot.2007.521
Yao, J., & Rock, C. O. (2016). Resistance mechanisms and the future
1–11. https​://doi.org/10.1111/cbdd.13536​
of bacterial Enoyl‐Acyl carrier protein reductase (FabI) antibiot-
ics. Cold Spring Harbor Perspectives in Medicine, 6(3), https​://doi.
org/10.1101/cshpe​rspect.a027045

You might also like