Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

pubs.acs.

org/jcim Article

Simulation Reveals the Chameleonic Behavior of Macrocycles


Daniel Sethio,† Vasanthanathan Poongavanam,† Ruisheng Xiong, Mohit Tyagi, Duc Duy Vo,
Roland Lindh, and Jan Kihlberg*
Cite This: J. Chem. Inf. Model. 2023, 63, 138−146 Read Online

ACCESS Metrics & More Article Recommendations *


sı Supporting Information
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.
Downloaded via UPPSALA UNIV on January 9, 2023 at 10:28:12 (UTC).

ABSTRACT: Conformational analysis is central to the design of bioactive molecules. It is particularly challenging for macrocycles
due to noncovalent transannular interactions, steric interactions, and ring strain that are often coupled. Herein, we simulated the
conformations of five macrocycles designed to express a progression of increasing complexity in environment-dependent
intramolecular interactions and verified the results against NMR measurements in chloroform and dimethyl sulfoxide. Molecular
dynamics using an explicit solvent model, but not the Monte Carlo method with implicit solvation, handled both solvents correctly.
Refinement of conformations at the ab initio level was fundamental to reproducing the experimental observations�standard state-of-
the-art molecular mechanics force fields were insufficient. Our simulations correctly predicted the intramolecular interactions
between side chains and the macrocycle and revealed an unprecedented solvent-induced conformational switch of the macrocyclic
ring. Our results provide a platform for the rational, prospective design of molecular chameleons that adapt to the properties of the
environment.

■ INTRODUCTION
Macrocycles are prominent among natural products and
the design of macrocyclic molecular chameleons remains an
important yet unresolved problem.
synthetic bioactive molecules, with vital roles in fields such In general, the conformational space populated by macro-
cycles is sampled well by different algorithms,14,15 but the
as supramolecular chemistry,1 artificial molecular machines,2
identification of the biologically relevant conformations, i.e.,
and drug discovery.3−5 Their unique properties originate from
solution- and/or target-bound conformations, remains a
the molecular-level preorganization of structural elements
formidable challenge.16,17 Recently, progress has been made
induced by macrocyclization. Predicting the conformational
for macrocyclic peptides which may form transannular
landscape of molecules is essential as it determines their
IMHBs.18−20 However, for nonpeptidic and semipeptidic
physical and biomolecular properties, but conformational
macrocycles, minimum energy conformations (MECs) identi-
analysis of macrocycles is complex and challenging.6,7 Reasons
fied by molecular mechanics using implicit solvation models
for this include that macrocycles often have a large number of
show large differences from the biologically relevant con-
degrees of freedom, form noncovalent transannular inter-
formations, while conformations resembling the biologically
actions such as intramolecular hydrogen bonds (IMHBs), and
relevant ones usually reside at ≥10−15 kcal mol−1 above the
that steric interactions and ring strain also influence macro-
cycle conformations. Large and flexible side chains further
increase the complexity of macrocycle conformational analysis, Received: August 29, 2022
in particular, as the flexibility of the ring may amplify the Published: December 23, 2022
overall flexibility.8,9 Nonbonded intramolecular interactions
between side chains and the macrocyclic ring allow macro-
cycles to behave as molecular chameleons8−13 a feature which
may provide major advantages in drug discovery.12 However,
© 2022 The Authors. Published by
American Chemical Society https://doi.org/10.1021/acs.jcim.2c01093
138 J. Chem. Inf. Model. 2023, 63, 138−146
Journal of Chemical Information and Modeling pubs.acs.org/jcim Article

Figure 1. (A) Structures of the macrocycles investigated in this study. (B) 1H NMR spectra of 1−5 in CDCl3 and DMSO-d6. Chemical shifts of
each of the three amide protons NH-I (red), NH-II (green), and NH-III (blue) are indicated.

MECs.16,17,21 Hence, this calls for the development of formation of an intramolecular NH···π interaction (2−4) or a
computational protocols which correctly model the behavior hydrogen bond (5) with either of the adjacent amide bonds.
of macrocycles in a variety of environments such that the The chemical shifts of the three amide protons of 1−5 indicate
predicted properties can guide the design of molecules for major structure- and environment-dependent conformational
specific purposes. The need for the prospective, property-based differences between 1−5 (Figure 1B, Supplementary Figures
design of macrocycles3,22 is further underlined by the S1−S12, Tables S1 and S2). For instance, the chemical shift of
complexity of their synthesis.23 NH-I shows a large variation between 1−5 in a nonpolar
Inspired by the natural products hymenocardine,24 K-13,25 environment (CDCl3, ε = 4.8) but not in the polar one
and OF4949-III,26 we designed macrocycles 1−5 that contain (DMSO-d6, ε = 46.7). As chemical shifts are sensitive to
an 18-membered macrocyclic ring with two side-chains (R and molecular structure and dynamics,28−32 macrocycles 1−5
Boc, Figure 1A).27 The R side-chain was chosen to allow the constitute an ideal system for the development and assessment
139 https://doi.org/10.1021/acs.jcim.2c01093
J. Chem. Inf. Model. 2023, 63, 138−146
Journal of Chemical Information and Modeling pubs.acs.org/jcim Article

of computational protocols to predict macrocycle conforma- bar) using the Berendsen coupling algorithm43 for another 500
tions. Herein, we report our successful efforts to this end. ps MD simulation. Finally, the MD production run was
Molecular dynamics (MD) simulations, followed by DFT performed for 50 ns for each compound in the two solvents.
optimization of the generated conformational ensembles, The SHAKE algorithm44 was used to constrain all bonds
allowed the correct prediction of solvent-dependent dynamic involving hydrogen atoms. Snapshots were collected every 10
intramolecular interactions and of a conformational switch of ps from the 50 ns simulation, resulting in a total of 5000
the ring in the series of macrocycles. snapshots for further analysis for each compound and solvent.

■ METHODS
Monte Carlo Conformational Sampling (MCMM). The
A nonbiased selection of 10 of the 5000 snapshots provides a
representative subset with minimal likelihood of not including
the major conformation and several high-energy conforma-
2D structures of the designed macrocycles 1−5 were built with tions. Every 500th conformation was therefore chosen, and the
correct stereochemistry using Maestro from the Schrödinger energy was calculated at the B3LYP/6-311++G(2d,p) level of
suite.33 The resulting initial structures were directly used as theory. The MD protocol used the MEC out of these 10
input for the Monte Carlo Multiple Minimum (MCMM) conformations for each compound in each of the two solvents
Conformational Sampling using implicit solvation models. to calculate the chemical shifts of NH-I, NH-II, and NH-III.
The MCMM conformational samplings were conducted Subsequently, all conformations (10 per compound and
using the Macromodel software as implemented in the solvent) were subjected to DFT-based geometry optimization
Schrödinger Suite version 2020-4.34 The MCMM sampling35 as described below.
was done in both polar and nonpolar environments as Trajectory Analysis. The conformational change of
represented by dimethyl sulfoxide (DMSO, ε = 46.7) and macrocycles was followed through RMSD, inter- and intra-
chloroform (CHCl3, ε = 4.8) using the OPLS3e force field.36 molecular hydrogen bonding, and NH···π interactions utilizing
Briefly, the conformational space of each macrocycle in each of the cpptraj tool.45 The RMSD values were calculated with
the two solvents was stochastically explored by extended respect to the initial geometry, while the NH···π interactions
sampling of the torsional angles. The following settings were were evaluated by measuring the distance between the N atom
used for the study: energy window (10 kcal/mol), elimination to the center of the phenyl ring. The flexibility of the dipeptide
of duplicate conformer threshold (RMSD, 0.75 Å), the total of the macrocyclic ring was monitored through the change of
number of iterations (10,000 steps), and force field [Polak- the two dihedral angles ϕ and ψ (see the Supporting
Ribière Conjugate Gradient (PRCG) in combination with the Information).
OPLS3e force field36]. For each kept unique conformation Geometry Optimization (DFT). The 10 conformations
(within the 10 kcal/mol window, RMSD >0.75 Å relative to obtained from the MCMM calculations and the MD
conformations identified earlier in the sampling), a degeneracy simulations for each compound in each of the two solvents
index was assigned, indicating how many times identical were further optimized at the DFT level using the Gaussian 16
conformations were found in the iterative sampling. Finally, the Rev. C.01 program.41 To find a reliable and practical method
10 conformations that had the highest degeneracy index were for describing nonbonded interactions, several DFT func-
selected. The MC protocol used the MEC from the MCMM tionals and basis sets have been tested. The dispersion-
calculations for each compound in each of the two solvents to corrected functionals such as B3LYP46,47 augmented with the
calculate the chemical shifts of NH-I, NH-II, and NH-III. semiempirical Grimme’s D3 dispersion corrected,48 M06-2X,49
Subsequently, all conformations (10 per compound and and wB97X-D50 functionals have been tested together with
solvent) were subjected to DFT-based geometry optimization Pople’s 6-31+G(d,p)51 and 6−31++G(d,p)52 as well as
as described below. Dunning’s cc-pVDZ and cc-pVTZ basis sets.53,54 The M06-
MD Simulations. All MD simulations for compounds 1−5 2X functional in combination with Pople’s 6-31 + G(d,p) was
were performed using the Amber18 package37 utilizing the chosen as the best compromise to accurately describe
General Amber Force Field (GAFF).38 The initial geometry nonbonded interactions as suggested by Willoughby and co-
for 1−5 was taken from the lowest energy conformer as workers.55 The implicit polarizable continuum model of
obtained from the MCMM calculation. Charges for atoms Tomasi and co-workers was used to account for CHCl3 and
were assigned based on the electrostatic potential (ESP) fitting DMSO solvation effects.56 To ensure the optimized geometry
using the RESP procedure39 as implemented in Antechamber corresponding to a minimum on the potential energy surface,
software.40 The ESP calculation was carried out at the B3LYP/ vibrational frequency calculations were conducted at the same
cc-pVTZ level of theory, using the Gaussian 16 Rev. C.01 level of theory. The MECs for each compound in each of the
software.41 Solvent molecules (CHCl3 or DMSO) were added two solvents were used to calculate the chemical shifts of NH-
to the macrocycles with 30 Å buffering distance between the I, NH-II, and NH-III in the MC:DFT and MD:DFT protocols.
edges of the truncated cubic box. Subsequently, the tleap tool42 All 10 conformations were used for the MC:DFTens and
from the Amber package was used to build topology and MD:DFTens protocols.
coordinate files for carrying out the MD calculations. Initially, NMR Chemical Shift Calculations. NMR chemical
energy minimization was carried out in two steps. First, the shielding(s) calculations were calculated at the B3LYP/6-
steepest descent method was used to minimize the energy with 311++G(2d,p) level of theory utilizing the Gauge-Independent
respect to the hydrogen atom positions with a maximum of Atomic Orbital (GIAO) method.57 The triple-ξ polarized
1000 steps, whereas heavy atoms were constrained. Second, quality augmented with the diffuse functions basis set is
the conjugate gradient method with 2000 steps was used to necessary to obtain accurate NMR shielding tensor values.58
minimize the whole solute−solvent system. Subsequently, the The NMR chemical shifts are obtained by subtracting the
system was heated from 0 to 300 K at a constant volume for chemical shielding tensor value of the reference compound
200 ps. After the thermalization process, the system was with the chemical shielding tensor value of the molecule of
equilibrated at constant temperature (300 K) and pressure (1 interest (δ = σref − σ). Tetramethyl silane (TMS) was used as a
140 https://doi.org/10.1021/acs.jcim.2c01093
J. Chem. Inf. Model. 2023, 63, 138−146
Journal of Chemical Information and Modeling pubs.acs.org/jcim Article

reference compound for calculating the GIAO tensor of 1H (DMSO) were developed. Second, the robustness of the
NMR. protocols was evaluated by a comparison between theoretical
Model Evaluation. Prediction of relevant conformations and experimental results, i.e., the chemical shifts of protons
was evaluated in the following scenario, if the minimum energy NH-I, NH-II, and NH-III (Figure 1),273JH,H scalar coupling
conformer (MEC) is representative (or not) of the constants, and nuclear Overhauser effects (NOEs) of protons
conformations populated by the compound in different within the macrocyclic ring of 1−5. For the first part, six
environments. different protocols were explored to generate representative
Interproton Distances (NOEs). To obtain quantitative conformations of 1−5 (Table 1). Initial conformational
inter-proton distances, seven 1H,1H-NOESY experiments were sampling was performed either by the MCMM method,
evaluated with increasing mixing times ranging from 50 to 350 which incorporated an implicit solvation model, or by the more
ms. Experiments were recorded in random order to counteract computationally intensive unrestrained-MD simulations using
systematic errors. Spectra were acquired with 1024 × 2048 an explicit solvation model. These sampling methods are based
complex points (F1 × F2) and a spectral width of 6410 Hz. on state-of-the-art force fields and are denoted MC and MD,
The recycle relaxation delay (d1) was set to 3 s. Deuterium respectively, when their MECs are used as representative
signals of CDCl3 and DMSO-d6 were used for the lock signal. conformations. An improved level of conformational sampling
Diagonal and cross peaks were integrated for all mixing times was established by ab initio DFT molecular geometry
and individual NOE intensities taken as the absolute integral optimization of ten conformations selected from the
were normalized according to established MC and MD ensembles (cf. Methods). We denote
the use of the MECs from these optimized ensembles
|cross peakij × cross peakij| protocols MC:DFT and MD:DFT. Finally, the full ensembles
=2
norm, ij
|diagonal peaki × diagonal peakj| consisting of ten conformations for each macrocycle constitute
(1) the MC:DFTens and MD:DFTens protocols, respectively. We
per mixing time, respectively. A minimum of four normalized note that the MECs of the MC:DFT protocol were used as a
intensities, ηnorm, for consecutive mixing times describing a seed for subsequent MD conformational exploration.
linear initial build-up with R2 ≥ 0.95 were used to determine Protocol for Finding Relevant Conformations of
the build-up rate σij. Build-up rates of locked-in-distance Macrocycles. As expected from previous reports,16,17 the
protons were used as distance reference (e.g., geminal protons: MECs predicted by MC sampling were not representative of
1.78 Å). Distances were calculated according to the solution conformations of 1−5 (Table 1, protocol MC,
Supplementary Figure S13). The MC:DFT protocol provided
1/6
ij yz a significant improvement in the correlation between predicted
rij = rref jjjj ref zz
zz and observed chemical shifts for the MECs in chloroform, but
j z
k ij { (2) not in DMSO (Table 1, Figure 2A). Inspection of the MECs
from this protocol indicated that the shielding and deshielding
where rij is the resulting distance between protons i and j, rref is
of NH-I observed for compounds 3 and 5 in chloroform
the fixed distance of the reference proton pair, σref is the build-
(Figure 1B) originated from the formation of an intramolecular
up rate of the used reference protons, and σij is the build-up
NH···π interaction or hydrogen bond with the adjacent R
rate of the protons of interest using normalized NOE
intensities.59 phenyl and 2-pyridyl groups, respectively. The IMHB was
maintained in implicit DMSO, providing one explanation for
■ RESULTS AND DISCUSSION
Study Design. First, protocols to identify the conforma-
the poor correlation in this solvent. Averaging of chemical
shifts calculated from the DFT-optimized conformational
ensembles within an energy window of 5, 10, 15, and 20
tional dynamics of macrocycles as a function of the polarity of kcal mol−1 did not provide any major improvement (Table 1,
the environment, i.e., apolar (chloroform) versus polar protocol MC:DFTens, Supplementary Figure S14). An
implicit solvation model thus failed to reproduce solvation
Table 1. Protocolsa Evaluated for Conformational Sampling effects in the polar solvent DMSO, while it provided a good
of Macrocycles 1−5 and the Quality of the Models for the representation of the conformations adopted in an apolar
Prediction of the Chemical Shifts of NH-I, NH-II, and NH- environment.60,61 We, therefore, turned to unrestrained MD
IIIb simulations in explicit solvents to capture the solvation effects
CHCl3 (ε 4.8) DMSO (ε 46.7) on the conformations of 1−5.
a 2 The MECs of protocol MC:DFT, all had the R side-chain
protocol solvation model R RMSE R2 RMSE
oriented equatorially on the macrocyclic ring (Supplementary
MC implicit 0.13 1.16 −0.32 2.72 Figure S15), but the ensembles also contained conformations
MC:DFT implicit 0.81 0.57 0.0003 2.41 in which both amide bonds within the macrocycle had been
MC:DFTens implicit 0.91 0.68 −0.002 2.17 rotated in the plane, thereby placing R axially (Figure 2B). In
MD explicit 0.29 0.93 0.02 1.82
chloroform, conformations that had R axially had energies
MD:DFT explicit 0.88 0.66 0.63 1.26
higher than that of the MEC (>3−22 kcal mol −1 )
MD:DFTens explicit 0.97 0.45 0.59 0.95
(Supplementary Table S3), suggesting the equatorial con-
a formations as starting points for the MD simulations. To
MC: Monte Carlo Multiple Minimum; MD: Molecular Dynamics;
DFT: Density Functional Theory. The subscript “ens” indicates that understand whether these two minima are connected in the
the model is based on the average of the chemical shifts of the potential energy surfaces, we performed MD simulations at
conformational ensemble within 10 kcal mol−1 of the MEC. bThe increasing temperatures for compound 5 in DMSO, since 5
quality of the protocols are described by the Pearson’s correlation may be particularly prone to remain in a local minimum due to
coefficient (R2) and the root mean square error (RMSE, δ, ppm). the formation of an IMHB between the pyridyl group and NH-
141 https://doi.org/10.1021/acs.jcim.2c01093
J. Chem. Inf. Model. 2023, 63, 138−146
Journal of Chemical Information and Modeling pubs.acs.org/jcim Article

Figure 2. (A) Correlation between predicted and experimental 1H chemical shifts in chloroform and DMSO for compounds 1−5. The top two
panels show the correlations obtained for the MECs from protocol MC:DFT, the middle panels show the correlations obtained for the MECs from
protocol MD:DFT, while the two panels at the bottom show the correlations based on the MD:DFTens ensemble within 10 kcal mol−1 of the MEC.
R2: the Pearson correlation coefficient; RMSE: root-mean-square-error (ppm). (B) Schematic description of the conformational interconversion of
the macrocyclic ring that results in the reorientation of the R side-chain from and equatorial to an axial position.

R equatorial as suitable starting points for the MD simulations


also in DMSO.
The MECs from the MD protocol provided poor
correlations to the experimental chemical shifts both in
chloroform and DMSO (Table 1, Supplementary Figure
S17) probably due to an inadequate description of the
solvation effects by the force field. However, the analysis of
the frequency of hydrogen bonds revealed that IMHBs found
in chloroform were broken up in DMSO and that the amide
protons of 1−5 were hydrogen-bonded to DMSO (Supple-
mentary Figures S18 and S19). DFT optimization, protocol
MD:DFT, of the conformations, including the hydrogen-
bonded DMSO molecules, led to a major improvement of the
chemical shift correlations for the MECs both in chloroform
and in DMSO (Table 1, Figure 2A). Lastly, the use of the <10
kcal mol−1 ensembles of protocol MD:DFTens provided even
better models in both solvents (Table 1, Figure 2A,
Supplementary Figure S20) than the MD:DFT protocol.
The 3JH,H coupling constants for the protons in the macrocyclic
Figure 3. Distance between (A) the nitrogen atom of NH-I and the ring of 1−5 were used to further validate the protocols.
center of the R phenyl ring and (B) the nitrogen atom of NH-I and Encouragingly, the simulated coupling constants for the MECs
the nitrogen atom of the pyridyl moiety in the MD trajectories of of protocol MD:DFT agreed well with the experimentally
compounds 3, 4, and 5, respectively, in DMSO and CHCl3. Distance
range for the formation of a NH···π shielding (2.9−3.6 Å) and non-
determined 3JH,H values both in chloroform and in DMSO
negligible IMHB (1.5−2.5 Å) are highlighted in green. (Supplementary Tables S4 and S5). This was also the case for
the optimized MECs of the MC:DFT sampling protocol in
chloroform, but not in DMSO.
Validation of the Ability of the MD:DFTens Protocol
I. These simulations showed a transition from the axial to the To Predict Molecular Chameleons. The MD simulations
equatorial orientation of the R side-chain at 500 K suggested that the phenyl ring of phenylalanine 3 formed a
(Supplementary Figure S16), revealing conformations having significantly larger proportion of NH···π interactions with NH-
142 https://doi.org/10.1021/acs.jcim.2c01093
J. Chem. Inf. Model. 2023, 63, 138−146
Journal of Chemical Information and Modeling pubs.acs.org/jcim Article

Figure 4. (A) RMSD of the heavy atoms of the macrocyclic ring of compounds 1, 3, and 5 calculated from the MD trajectories in explicit
chloroform and DMSO. Representative conformations having the ring in states 1 and 2 have been indicated for each compound. (B) Structures of
the state 1 and 2 conformations adopted in DMSO. The 3JH,H coupling constants and NOEs for protons NH-I and Ha−Hc provide key information
that differentiates the two states. Coupling constants calculated for the state 1 and 2 conformations of the MEC from the MD:DFT protocol of
compound 3 in DMSO have been included in the figure. Coupling constants determined by NMR spectroscopy in DMSO-d6 for 3 have been
inserted in the box for comparison.

I in chloroform than for homophenylalanine 4 (Figure 3A). major conformations with 1 showing this behavior both in
Experimental support is provided by the large shielding of NH- chloroform and DMSO, while 3 and 5 displayed it only in
I of 3 in chloroform as compared to 4 (Figure 1). For DMSO (Figure 4A). The conformations from protocol
compound 5, NH-I forms an IMHB to the pyridine moiety MD:DFTens of 1, 3, and 5 revealed that the two conforma-
with high frequency in chloroform (Figure 3B), as expected tional states originated from an in-plane rotation of the NH-I
from the large deshielding of NH-I observed in the NMR amide bond (Figure 4B). This places NH-I and NH-II on
spectrum of 5 (Figure 1). These nonbonded interactions are opposite faces of the macrocycle in state 1 and on the same
disrupted in the MD simulations of 3 and 5 in explicit DMSO face in state 2. The DFT-optimized MECs for the three
(Figure 3), in excellent agreement with that shielding/ macrocycles belong to state 1 both in chloroform and in
deshielding of NH-I are not observed in DMSO (Figure 1). DMSO. In chloroform, the second state of 1 was found to
The MD trajectories of compounds 1−5 in the two reside >20 kcal mol−1 above the MEC, i.e. the population of
environments also suggested that the macrocyclic ring state 2 is negligible for 1 just as suggested by the MD
populates a few major conformations (Supplementary Figure simulations for 3 and 5 (Supplementary Table S6). However,
S21). For 1, 3, and 5, the ring appeared to switch between two in DMSO, the relative population of state 2 was indicated to be
143 https://doi.org/10.1021/acs.jcim.2c01093
J. Chem. Inf. Model. 2023, 63, 138−146
Journal of Chemical Information and Modeling pubs.acs.org/jcim Article

significant for all three macrocycles due to the coordination of Tables and figures with calculated data and NMR data;
a DMSO molecule by NH-I and NH-II (Figure 4). and experimental procedures for the synthesis of 5
The 3JH,H coupling constants and NOEs for protons NH-I, (PDF)
Ha, and Hb support the solvent-induced switch of the Excel showing SMILES of compounds 1−5 (XLSX)
macrocyclic ring (Figure 4). In chloroform, the experimentally Excel showing chemical shifts statistics (XLSX)
determined coupling constants between NH-I and the two RMSD analysis (XLSX)
adjacent methylene protons agree very well with the coupling Dihedral angle analysis (XLSX)
constants predicted for state 1 (Supplementary Table S7), Distance and IMHB analysis (XLSX)
providing experimental support for the population of only this
state in an apolar environment. In DMSO, the 3JNH‑I‑Hb
coupling constant determined by NMR spectroscopy lies in
between the ones calculated for the state 1 and 2
■ AUTHOR INFORMATION
Corresponding Author
conformations for the three macrocycles (exemplified for 3 Jan Kihlberg − Department of Chemistry − BMC, Uppsala
in Figure 4B, Table S8). This is also the case for 3JNH‑I‑Ha for 3, University, SE-751 23 Uppsala, Sweden; orcid.org/0000-
whereas line-broadening prevented analysis for 1 and 5. The 0002-4205-6040; Email: jan.kihlberg@kemi.uu.se
coupling constants thus support that 1, 3, and 5 interconvert
rapidly between the state 1 and 2 conformations in DMSO. Authors
NOESY confirmed the existence of the solvent-induced Daniel Sethio − Department of Chemistry − BMC, Uppsala
conformational switch. Specifically, NH-I displays an observ- University, SE-751 23 Uppsala, Sweden; orcid.org/0000-
able NOE only to one of the adjacent methylene protons (Ha) 0002-8075-1482
in chloroform, but to both methylene protons in DMSO Vasanthanathan Poongavanam − Department of Chemistry
(Figure 4B, Supplementary Table S9, Figures S22−S27). In − BMC, Uppsala University, SE-751 23 Uppsala, Sweden;
addition, the distance between NH-I and methine proton Hc orcid.org/0000-0002-8880-9247
was determined to be longer in DMSO than in chloroform for Ruisheng Xiong − Department of Chemistry − BMC,
1, 3, and 5 as would be expected from the population of both Uppsala University, SE-751 23 Uppsala, Sweden
states in DMSO (Supplementary Table S9). Mohit Tyagi − Department of Chemistry − BMC, Uppsala
University, SE-751 23 Uppsala, Sweden

■ CONCLUSIONS
Herein, we have reported the successful prediction of the
Duc Duy Vo − Department of Chemistry − BMC, Uppsala
University, SE-751 23 Uppsala, Sweden
Roland Lindh − Department of Chemistry − BMC, Uppsala
conformational ensembles of a series of macrocycles in University, SE-751 23 Uppsala, Sweden; orcid.org/0000-
environments that differ in polarity and hydrogen bonding 0001-7567-8295
capacity. Monte Carlo conformational sampling in implicit Complete contact information is available at:
solvent, followed by DFT optimization, provided an excellent https://pubs.acs.org/10.1021/acs.jcim.2c01093
description of the conformations populated in chloroform, but
not so in the polar DMSO. The lack of predictivity of implicit Author Contributions
solvent models in polar environments has been noted earlier †
D.S. and V.P. contributed equally.
for macrocycles16,17,21 and was comprehensively illustrated in a
Notes
most recent study of linear peptides.62 However, we found that
MD simulations using explicit solvation followed by DFT The authors declare no competing financial interest.
optimization described the conformations of the macrocycles MD trajectories and molecular topology parameter files from
correctly in both environments. These observations indicate the MD simulations are available free of charge at https://doi.
that hydrogen bonding dictates the use of explicit solvent org/10.5281/zenodo.7022549. Python scripts used for auto-
models for correct predictions in polar solvents. mation of the calculation of chemical shifts and optimized
Our MD simulations suggested that DMSO induced a geometries are available free of charge at https://github.com/
conformational switch of the macrocyclic ring, while chloro- danielquantum/NMR-Macrocycles.
form did not. Determination of coupling constants and NOEs
by NMR spectroscopy verified this dynamic, solvent-depend-
ent conformational flexibility. As the solvent-dependent
■ ACKNOWLEDGMENTS
This work was funded by grants from the Swedish Research
formation of intramolecular NH···π and IMHB interactions Council [Grants No. 2021-04747 (JK) and 2020-03182 (RL)].
between side-chains and the macrocyclic backbone were also Computations were performed on resources provided by the
correctly predicted, we conclude that MD simulations, Swedish National Infrastructure for Computing (SNIC) under
followed by DFT optimization, can be used for the prospective project numbers SNIC 2021/22-244, SNIC 2021/22-125,
design of molecular chameleons. Molecular chameleonicity is SNIC 2021/22-860, and SNIC 2019/3-295. This study made
of major interest in drug discovery as it allows compounds to use of the NMR Uppsala infrastructure, which is funded by the
display high aqueous solubility and high cell permeability, Department of Chemistry - BMC and the Disciplinary Domain
properties that are highly desired but often difficult to engineer of Medicine and Pharmacy. The authors are grateful to Marjan
into a drug candidate. Khamesian, N. Arul Murugan, and Máté Erdélyi for their
valuable contributions and discussions.
■ ASSOCIATED CONTENT
* Supporting Information

■ REFERENCES
(1) Liu, Z.; Nalluri, S. K. M.; Stoddart, J. F. Surveying macrocyclic
The Supporting Information is available free of charge at chemistry: from flexible crown ethers to rigid cyclophanes. Chem. Soc.
https://pubs.acs.org/doi/10.1021/acs.jcim.2c01093. Rev. 2017, 46, 2459−2478.

144 https://doi.org/10.1021/acs.jcim.2c01093
J. Chem. Inf. Model. 2023, 63, 138−146
Journal of Chemical Information and Modeling pubs.acs.org/jcim Article

(2) Kassem, S.; van Leeuwen, T.; Lubbe, A. S.; Wilson, M. R.; (20) Hosseinzadeh, P.; Bhardwaj, G.; Mulligan, V. K.; Shortridge, M.
Feringa, B. L.; Leigh, D. A. Artificial molecular motors. Chem. Soc. Rev. D.; Craven, T. W.; Pardo-Avila, F.; Rettie, S. A.; Kim, D. E.; Silva, D.-
2017, 46, 2592−2621. A.; Ibrahim, Y. M.; Webb, I. K.; Cort, J. R.; Adkins, J. N.; Varani, G.;
(3) Driggers, E. M.; Hale, S. P.; Lee, J.; Terrett, N. K. The Baker, D. Comprehensive computational design of ordered peptide
exploration of macrocycles for drug discovery–an underexploited macrocycles. Science 2017, 358, 1461−1466.
structural class. Nat. Rev. Drug Discovery 2008, 7, 608−624. (21) Bonnet, P.; Agrafiotis, D. K.; Zhu, F.; Martin, E. Conforma-
(4) Villar, E. A.; Beglov, D.; Chennamadhavuni, S.; Porco, J. A.; tional analysis of macrocycles: Finding what common search methods
Kozakov, D.; Vajda, S.; Whitty, A. How proteins bind macrocycles. miss. J. Chem. Inf. Model. 2009, 49, 2242−2259.
Nat. Chem. Biol. 2014, 10, 723−731. (22) Giordanetto, F.; Kihlberg, J. Macrocyclic drugs and clinical
(5) Over, B.; Matsson, P.; Tyrchan, C.; Artursson, P.; Doak, B. C.; candidates: What can medicinal chemists learn from their properties?
Foley, M. A.; Hilgendorf, C.; Johnston, S. E.; Lee, M. D. T.; Lewis, R. J. Med. Chem. 2014, 57, 278−295.
J.; McCarren, P.; Muncipinto, G.; Norinder, U.; Perry, M. W.; Duvall, (23) Martí-Centelles, V.; Pandey, M. D.; Burguete, M. I.; Luis, S. V.
J. R.; Kihlberg, J. Structural and conformational determinants of Macrocyclization reactions: The importance of conformational,
macrocycle cell permeability. Nat. Chem. Biol. 2016, 12, 1065−1074. configurational, and template-induced preorganization. Chem. Rev.
(6) Appavoo, S. D.; Huh, S.; Diaz, D. B.; Yudin, A. K. 2015, 115, 8736−8834.
Conformational control of macrocycles by remote structural (24) Pais, M.; Marchand, J.; Ratle, G.; Jarreau, F. X. Peptidic
modification. Chem. Rev. 2019, 119, 9724−9752. alkaloids. VI. Hymenocardine, alkaloid from Hymenocardia acida Tul.
(7) Hawkins, P. C. D. Conformation generation: The state of the art. Bull. Soc. Chim. Fr. 1968, 7, 2979−2984.
J. Chem. Inf. Model. 2017, 57, 1747−1756. (25) Yasuzawa, T.; Shirahata, K.; Sano, H. K-13, a novel inhibitor of
(8) Rossi Sebastiano, M.; Doak, B. C.; Backlund, M.; Poongavanam, angiotensin I converting enzyme produced by Micromonospora
V.; Over, B.; Ermondi, G.; Caron, G.; Matsson, P.; Kihlberg, J. Impact halophytica subsp. exilisia. II Structure determination. J. Antibiot.
of dynamically exposed polarity on permeability and solubility of 1987, 40, 455−458.
chameleonic drugs beyond the rule of 5. J. Med. Chem. 2018, 61, (26) Sano, S.; Ikai, K.; Katayama, K.; Takesako, K.; Nakamura, T.;
4189−4202. Obayashi, A.; Ezure, Y.; Enomoto, H. OF4949, new inhibitors of
(9) Danelius, E.; Poongavanam, V.; Peintner, S.; Wieske, L. H. E.; aminopeptidase B. II Elucidation of structure. J. Antibiot. 1986, 39,
Erdélyi, M.; Kihlberg, J. Solution conformations explain the 1685−1696.
chameleonic behavior of macrocyclic drugs. Chem. − Eur. J. 2020, (27) Tyagi, M.; Poongavanam, V.; Lindhagen, M.; Pettersen, A.; Sjö,
26, 5231−5244. P.; Schiesser, S.; Kihlberg, J. Toward the design of molecular
(10) Rezai, T.; Yu, B.; Millhauser, G. L.; Jacobson, M. P.; Lokey, R. chameleons: Flexible shielding of an amide bond enhances macro-
S. Testing the conformational hypothesis of passive membrane cycle cell permeability. Org. Lett. 2018, 20, 5737−5742.
permeability using synthetic cyclic peptide diastereomers. J. Am. (28) Buckingham, A. D.; Schaefer, T.; Schneider, W. G. Solvent
Chem. Soc. 2006, 128, 2510−2511. effects in nuclear magnetic resonance spectra. J. Chem. Phys. 1960, 32,
(11) Wang, C. K.; Swedberg, J. E.; Harvey, P. J.; Kaas, Q.; Craik, D. 1227−1233.
J. Conformational flexibility is a determinant of permeability for (29) Taubert, S.; Konschin, H.; Sundholm, D. Computational
cyclosporin. J. Phys. Chem. B 2018, 122, 2261−2276. studies of 13C NMR chemical shifts of saccharides. Phys. Chem. Chem.
(12) Whitty, A.; Zhong, M.; Viarengo, L.; Beglov, D.; Hall, D. R.; Phys. 2005, 7, 2561−2569.
Vajda, S. Quantifying the chameleonic properties of macrocycles and (30) Case, D. A. Chemical shifts in biomolecules. Curr. Opin. Struct.
other high-molecular-weight drugs. Drug Discovery Today 2016, 21, Biol. 2013, 23, 172−176.
712−717. (31) Robustelli, P.; Stafford, K. A.; Palmer, A. G., 3rd Interpreting
(13) Sheikh, A. Y.; Mattei, A.; Miglani Bhardwaj, R.; Hong, R. S.; protein structural dynamics from NMR chemical shifts. J. Am. Chem.
Abraham, N. S.; Schneider-Rauber, G.; Engstrom, K. M.; Diwan, M.; Soc. 2012, 134, 6365−6374.
Henry, R. F.; Gao, Y.; Juarez, V.; Jordan, E.; DeGoey, D. A.; Hutchins, (32) Balazs, A. Y. S.; Carbajo, R. J.; Davies, N. L.; Dong, Y.; Hird, A.
C. W. Implications of the conformationally flexible, macrocyclic W.; Johannes, J. W.; Lamb, M. L.; McCoull, W.; Raubo, P.; Robb, G.
structure of the first-generation, direct-acting anti-viral paritaprevir on R.; Packer, M. J.; Chiarparin, E. Free ligand 1D NMR conformational
its solid form complexity and chameleonic behavior. J. Am. Chem. Soc. signatures to enhance structure based drug design of a Mcl-1 inhibitor
2021, 143, 17479−17491. (AZD5991) and other synthetic macrocycles. J. Med. Chem. 2019, 62,
(14) Martin, S. J.; Chen, I.-J.; Chan, A. W. E.; Foloppe, N. Modelling 9418−9437.
the binding mode of macrocycles: Docking and conformational (33) Schrödinger Release 2020-4. Maestro; Schrödinger, LLC: New
sampling. Bioorg. Med. Chem. 2020, 18, No. 115143. York, NY, 2020.
(15) Alogheli, H.; Olanders, G.; Schaal, W.; Brandt, P.; Karlén, A. (34) Schrödinger Release 2020-4. MacroModel; Schrödinger, LLC:
Docking of macrocycles: Comparing rigid and flexible docking in New York, NY, 2020.
glide. J. Chem. Inf. Model. 2017, 57, 190−202. (35) Chang, G.; Guida, W. C.; Still, W. C. An internal-coordinate
(16) Poongavanam, V.; Danelius, E.; Peintner, S.; Alcaraz, L.; Caron, Monte Carlo method for searching conformational space. J. Am.
G.; Cummings, M. D.; Wlodek, S.; Erdelyi, M.; Hawkins, P. C. D.; Chem. Soc. 1989, 111, 4379−4386.
Ermondi, G.; Kihlberg, J. Conformational sampling of macrocyclic (36) Harder, E.; Damm, W.; Maple, J.; Wu, C.; Reboul, M.; Xiang, J.
drugs in different environments − Can we find the relevant Y.; Wang, L.; Lupyan, D.; Dahlgren, M. K.; Knight, J. L.; Kaus, J. W.;
conformations? ACS Omega 2018, 3, 11742−11757. Cerutti, D. S.; Krilov, G.; Jorgensen, W. L.; Abel, R.; Friesner, R. A.
(17) Poongavanam, V.; Atilaw, Y.; Ye, S.; Wieske, L. H. E.; Erdelyi, OPLS3: a force field providing broad coverage of drug-like small
M.; Ermondi, G.; Caron, G.; Kihlberg, J. Predicting the permeability molecules and proteins. J. Chem. Theory Comput. 2016, 12, 281−296.
of macrocycles from conformational sampling − Limitations of (37) D. A., Case; H. M., Aktulga; K., Belfon; I.Y., Ben-Shalom; S. R.,
Molecular Flexibility. J. Pharm. Sci. 2021, 110, 301−313. Brozell; D. S., Cerutti; T. E., Cheatham, III; G. A., Cisneros; V. W. D.,
(18) Wang, S.; Konig, G.; Roth, H. J.; Fouche, M.; Rodde, S.; Cruzeiro; T. A., Darden; R. E., Duke; G., Giambasu; M. K., Gilson;
Riniker, S. Effect of flexibility, lipophilicity, and the location of polar H., Gohlke; A. W., Goetz; R., Harris; S., Izadi; S. A., Izmailov; C., Jin;
residues on the passive membrane permeability of a series of cyclic K., Kasavajhala; M. C., Kaymak; E., King; A., Kovalenko; T.,
decapeptides. J. Med. Chem. 2021, 64, 12761−12773. Kurtzman; T. S., Lee; S., LeGrand; P., Li; C., Lin; J., Liu; T.,
(19) Witek, J.; Keller, B. G.; Blatter, M.; Meissner, A.; Wagner, T.; Luchko; R., Luo; M., Machado; V., Man; M., Manathunga; K. M.,
Riniker, S. Kinetic models of cyclosporin A in polar and apolar Merz; Y., Miao; O., Mikhailovskii; G., Monard; H., Nguyen; K. A.,
environments reveal multiple congruent conformational states. J. O’Hearn; A., Onufriev; F., Pan; S., Pantano; R., Qi; A., Rahnamoun;
Chem. Inf. Model. 2016, 56, 1547−1562. D. R., Roe; A., Roitberg; C., Sagui; S., Schott-Verdugo; J., Shen; C. L.,

145 https://doi.org/10.1021/acs.jcim.2c01093
J. Chem. Inf. Model. 2023, 63, 138−146
Journal of Chemical Information and Modeling pubs.acs.org/jcim Article

Simmerling; N. R., Skrynnikov; J., Smith; J., Swails; R. C., Walker; J., (53) Dunning, T. H. Gaussian basis sets for use in correlated
Wang; H., Wei; R. M., Wolf; X., Wu; Y., Xue; D. M., York; S., Zhao; molecular calculations. I. The atoms boron through neon and
P. A., KollmanAmber 2018; University of California: San Francisco, hydrogen. J. Chem. Phys. 1989, 90, 1007−1023.
2018. (54) Woon, D. E., Jr. Gaussian basis sets for use in correlated
(38) Wang, J.; Wolf, R. M.; Caldwell, J. W.; Kollman, P. A.; Case, D. molecular calculations. V. Core-valence basis sets for boron through
A. Development and testing of a general amber force field. J. Comput. neon. J. Chem. Phys. 1995, 103, 4572−4585.
Chem. 2004, 25, 1157−1174. (55) Willoughby, P. H.; Jansma, M. J.; Hoye, T. R. A guide to small-
(39) Bayly, C. I.; Cieplak, P.; Cornell, W. D.; Kollman, P. A. A well- molecule structure assignment through computation of (1H and 13C)
behaved electrostatic potential based method using charge restraints NMR chemical shifts. Nat. Protoc. 2014, 9, 643−660.
for deriving atomic charges - the resp model. J. Phys. Chem. 1993, 97, (56) Tomasi, J.; Mennucci, B.; Cammi, R. Quantum mechanical
10269−10280. continuum solvation models. Chem. Rev. 2005, 105, 2999−3094.
(40) Wang, J. M.; Wang, W.; Kollman, P. A. Antechamber: An (57) Wolinski, K.; Hinton, J. F.; Pulay, P. Efficient implementation
accessory software package for molecular mechanical calculations. of the gauge-independent atomic orbital method for NMR chemical
Abstr. Pap. Am. Chem. Soc. 2001, 222, U403−U403. shift calculations. J. Am. Chem. Soc. 1990, 112, 8251−8260.
(41) Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; (58) Sethio, D.; Raggi, G.; Lindh, R.; Erdélyi, M. Halogen Bond of
Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Petersson, Halonium Ions: Benchmarking DFT Methods for the Description of
G. A.; Nakatsuji, H.; Li, X.; Caricato, M.; Marenich, A. V.; Bloino, J.; NMR Chemical Shifts. J. Chem. Theory Comput. 2020, 16, 7690−
Janesko, B. G.; Gomperts, R.; Mennucci, B.; Hratchian, H. P.; Ortiz, J. 7701.
V.; Izmaylov, A. F.; Sonnenberg, J. L.; Williams; Ding, F.; Lipparini, (59) Butts, C. P.; Jones, C. R.; Towers, E. C.; Flynn, J. L.; Appleby,
F.; Egidi, F.; Goings, J.; Peng, B.; Petrone, A.; Henderson, T.; L.; Barron, N. J. Interproton distance determinations by NOE −
Ranasinghe, D.; Zakrzewski, V. G.; Gao, J.; Rega, N.; Zheng, G.; surprising accuracy and precision in a rigid organic molecule. Org.
Liang, W.; Hada, M.; Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.; Biomol. Chem. 2011, 9, 177−184.
Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H.; Vreven, T.; (60) Schattenberg, C. J.; Kaupp, M. Extended benchmark set of
Throssell, K.; Montgomery, Jr., J. A.; Peralta, J. E.; Ogliaro, F.; main-group nuclear shielding constants and NMR chemical shifts and
Bearpark, M. J.; Heyd, J. J.; Brothers, E. N.; Kudin, K. N.; Staroverov, its use to evaluate modern DFT methods. J. Chem. Theory Comput.
V. N.; Keith, T. A.; Kobayashi, R.; Normand, J.; Raghavachari, K.; 2021, 17, 7602−7621.
Rendell, A. P.; Burant, J. C.; Iyengar, S. S.; Tomasi, J.; Cossi, M.; (61) Gao, P.; Zhang, J.; Peng, Q.; Zhang, J.; Glezakou, V.-A. General
Millam, J. M.; Klene, M.; Adamo, C.; Cammi, R.; Ochterski, J. W.; protocol for the accurate prediction of molecular 13C/1H NMR
Martin, R. L.; Morokuma, K.; Farkas, O.; Foresman, J. B.; Fox, D. chemical shifts via machine learning augmented DFT. J. Chem. Inf.
J.Gaussian 16 Rev. C.01: Wallingford, CT, 2016. Model. 2020, 60, 3746−3754.
(42) Case, D. A.; Cheatham, T. E., 3rd; Darden, T.; Gohlke, H.; (62) Lang, E. J. M.; Baker, E. G.; Woolfson, D. N.; Mulholland, A. J.
Luo, R.; Merz, K. M., Jr.; Onufriev, A.; Simmerling, C.; Wang, B.; Generalized Born implicit solvent models do not reproduce secondary
Woods, R. J. The Amber biomolecular simulation programs. J. structures of de novo designed Glu/Lys peptides. J. Chem. Theory
Comput. Chem. 2005, 26, 1668−1688. Comput. 2022, 18, 4070−4076.
(43) Berendsen, H. J. C.; Postma, J. P. M.; Gunsteren, W. F. V.;
DiNola, A.; Haak, J. R. Molecular dynamics with coupling to an
external bath. J. Chem. Phys. 1984, 81, 3684−3690.
(44) 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.
(45) Roe, D. R.; Cheatham, T. E. PTRAJ and CPPTRAJ: Software
for processing and analysis of molecular dynamics trajectory data. J.
Chem. Theory Comput. 2013, 9, 3084−3095.
(46) Becke, A. D. Density-functional exchange-energy approxima-
tion with correct asymptotic behavior. Phys. Rev. A 1988, 38, 3098−
3100.
(47) Lee, C.; Yang, W.; Parr, R. G. Development of the Colle-
Salvetti correlation-energy formula into a functional of the electron
density. Phys. Rev. B 1988, 37, 785−789.
(48) Grimme, S.; Antony, J.; Ehrlich, S.; Krieg, H. A consistent and
accurate ab initio parametrization of density functional dispersion
correction (DFT-D) for the 94 elements H-Pu. J. Chem. Phys. 2010,
132, 154104.
(49) Zhao, Y.; Truhlar, D. G. The M06 suite of density functionals
for main group thermochemistry, thermochemical kinetics, non-
covalent interactions, excited states, and transition elements: two new
functionals and systematic testing of four M06-class functionals and
12 other functionals. Theor. Chem. Acc. 2008, 120, 215−241.
(50) Chai, J.-D.; Head-Gordon, M. Long-range corrected hybrid
density functionals with damped atom−atom dispersion corrections.
Phys. Chem. Chem. Phys. 2008, 10, 6615−6620.
(51) Hehre, W. J.; Ditchfield, R.; Pople, J. A. Self�consistent
molecular orbital methods. XII. Further extensions of Gaussian�type
basis sets for use in molecular orbital studies of organic molecules. J.
Chem. Phys. 1972, 56, 2257−2261.
(52) Hariharan, P. C.; Pople, J. A. The influence of polarization
functions on molecular orbital hydrogenation energies. Theor. Chim.
Acta 1973, 28, 213−222.

146 https://doi.org/10.1021/acs.jcim.2c01093
J. Chem. Inf. Model. 2023, 63, 138−146

You might also like