You are on page 1of 11

pubs.acs.

org/biochemistry Article

Double Mutant of Chymotrypsin Inhibitor 2 Stabilized through


Increased Conformational Entropy
Yulian Gavrilov, Felix Kümmerer, Simone Orioli, Andreas Prestel, Kresten Lindorff-Larsen,*
and Kaare Teilum*
Cite This: https://doi.org/10.1021/acs.biochem.1c00749 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.

ABSTRACT: The conformational heterogeneity of a folded protein can affect not


Downloaded via UNIV ESTADUAL PAULISTA on January 22, 2022 at 02:24:00 (UTC).

only its function but also stability and folding. We recently discovered and
characterized a stabilized double mutant (L49I/I57V) of the protein CI2 and
showed that state-of-the-art prediction methods could not predict the increased
stability relative to the wild-type protein. Here, we have examined whether
changed native-state dynamics, and resulting entropy changes, can explain the
stability changes in the double mutant protein, as well as the two single mutant
forms. We have combined NMR relaxation measurements of the ps−ns dynamics
of amide groups in the backbone and the methyl groups in the side chains with
molecular dynamics simulations to quantify the native-state dynamics. The NMR
experiments reveal that the mutations have different effects on the conformational
flexibility of CI2: a reduction in conformational dynamics (and entropy estimated from this) of the native state of the L49I variant
correlates with its decreased stability, while increased dynamics of the I57V and L49I/I57V variants correlates with their increased
stability. These findings suggest that explicitly accounting for changes in native-state entropy might be needed to improve the
predictions of the effect of mutations on protein stability.

■ INTRODUCTION
Controlling protein stability is an important challenge in
the effect of mutations on conformational dynamics of various
protein systems.8,12−17 Here, we apply these complementary
biotechnology and biopharmaceutical applications. The ability techniques to study the small 7.3 kDa protein, chymotrypsin
to predict how protein mutations decrease stability can be used inhibitor 2 (CI2) from barley seeds,18 which has been
in clinical genetics to interpret gene or genome sequencing extensively used in both experimental and computational
data.1 Similarly, improving the stability of enzymes2 or protein studies.19−24 The conformational dynamics of the polypeptide
pharmaceuticals3 by mutations may improve shelf life and backbone and the side chains of CI2 on the ps−ns timescales
tolerance to heat. The stability of a protein depends on have been studied by NMR spectroscopy.25 Previously, it was
interactions and dynamics of both the native and denatured demonstrated that five single-point mutations in the hydro-
phobic core of CI2 (L49V, L49A, I57V, V9A, and V47A) all
states of the protein.4
The conformational dynamics of the native state of a protein lead to a similar yet small increase in backbone and side-chain
contributes to its conformational stability. If the dynamics of flexibility throughout the structure. These (and other) core
mutations only have a minor effect on the inhibitory function
the native state is increased without significantly affecting the
of CI2.26
intramolecular interactions, it will lead to a more stable
According to experimental19 and computational studies,27,28
protein.5,6 Typically, however, the increased dynamics is
residues L49 and I57 are both important for formation of the
counteracted by fewer stabilizing interactions through
CI2 folding nucleus and for the stability of the folded state as
entropy-enthalpy compensation in the more dynamic mole-
they participate in a central network of interactions.25−31 L49
cule.7,8 In some cases, evolution has circumvented this
(together with R48) is a key energetic linchpin connecting
problem. Thus, in hyperthermophilic proteins, Lys is preferred
CI2’s hydrophobic core to its inhibitory loop.27,30 Further-
over Arg compared to the mesophilic homologs. This has been
attributed to high residual entropy of the Lys side chain
relative to Arg in folded protein structures.9 However, even Received: November 18, 2021
small local changes in a protein structure induced by point Revised: December 21, 2021
mutations can affect protein conformational equilibria,
dynamics, and its interactions.10,11 NMR spectroscopy, in
combination with computational methods such as molecular
dynamics simulations, has successfully been applied to analyze

© XXXX American Chemical Society https://doi.org/10.1021/acs.biochem.1c00749


A Biochemistry XXXX, XXX, XXX−XXX
Biochemistry pubs.acs.org/biochemistry Article

more, L49 and I57 are involved in a network of interactions source and 15NH4Cl as the sole nitrogen source, (iii) 13C6-Glc
that dissipate the structural strain occurring when CI2 binds to as the sole carbon source and 15NH4Cl as the sole nitrogen
its substrate.31 This network includes residue I37 in the source in 50% D2O, and (iv) 10% 13C6-Glc/90% unlabeled Glc
inhibitory loop. The neighboring residues, T36 and V38, point as the carbon source and 15NH4Cl as the sole nitrogen source.
their side chains toward the highly conserved residues of the The growth was continued overnight at 20 °C. Harvested cells
inhibitor loop, R48 and F50. Between these residues, L49 were resuspended in 25 mM Tris HCl, pH 8, with 1 mM of
forms part of the hydrophobic core together with A16 and EDTA. Cell lysis was performed with two freeze−thaw cycles.
I57.31 The insoluble material was removed by centrifugation at
We recently showed that L49I destabilizes the folded state of 30,000g for 20 min at 5 °C.
CI2 (ΔΔGf = 3.4 ± 0.1 kJ/mol), while I57V stabilizes it Nucleic acids were precipitated with 2% polyethylene imine
slightly (ΔΔGf = −2.2 ± 0.1 kJ/mol).32 Notably, the double and removed by centrifugation at 20,000g for 15 min at 5 °C.
mutant (L49I/I57V) is significantly more stable than either of CI2 was precipitated by adding ammonium sulfate to 70%
the single mutants (ΔΔGf = −3.8 ± 0.1 kJ/mol; Figure 1). saturation and centrifuging at 20,000g for 15 min at 5 °C. The
ammonium sulfate precipitate was resuspended in 10 mM
ammonium bicarbonate, and pure CI2 was obtained by SEC
on a Superdex 75 26/600 in 10 mM ammonium bicarbonate.
NMR Experiments. All NMR experiments were performed
at constant temperature (25 °C) and buffer conditions (50
mM MES, pH 6.26, 5% D2O, and 0.02% NaN3). The 15N-
HSQC, HNCO, HN(CA)CO, HNCA, HNCACB, and
CBCA(CO)NH spectra were acquired and used for backbone
chemical shift (CS) assignments. The methyl chemical shifts
were assigned using the 13C-HSQC, C(CO)NH, and HC-
Figure 1. (A) Overview of the positions of the mutated sites in CI2 (CO)NH spectra. These spectra were acquired for all studied
(space-filling). (B) Mutation cycle labeled with experimental ΔΔGf CI2 variants on a Varian DD2 800 MHz spectrometer
values at 25 °C. equipped with a room temperature probe. The triple resonance
spectra were recorded with non-uniform sampling at 25% and
This synergistic stabilizing effect of the double mutant is reconstructed by the compressed sensing method iterative soft
unexpected and could not be predicted using commonly thresholding using the software qMDD.35 We stereospecifically
applied methods such as FoldX33 and Rosetta.34 Failure to assigned the methyl groups of the valine and leucine side
properly capture long-range structural changes and changes in chains from the constant-time 13C-HSQC spectra recorded on
the dynamic properties of the protein may contribute to the the sample expressed in the presence of 10% 13C6-Glc.36
unsuccessful attempt to predict the effect of the double Backbone dynamics was probed with a series of 15N R1, R2
mutation. and heteronuclear 1H-15N NOE relaxation experiments using
Here, we asked if the change in stability of the L49I/I57V the pulse sequences by Farrow et al.37 The data were recorded
double mutant can, at least in part, be explained by changes in on Bruker Avance III HD 600 and 750 MHz spectrometers
the dynamics of CI2. We used NMR spectroscopy to equipped with cryoprobes using ∼1.5 mM 15N-labeled protein
experimentally assess the changes in fast dynamics of the samples. Relaxation delays of 10, 200, 400, 600, 800, and 1000
three CI2 variants L49I, I57V, and L49I/I57V relative to the ms were chosen for R1 and 16.96, 84.8, 169.6, 254.4, 339.2, and
wild type. To complement the experimental data, we also 424.8 ms for R2, respectively. The 1H-15N NOE experiments
performed molecular dynamics (MD) simulations and were recorded with a 5 s interscan delay. All experiments were
reweighted the obtained computational dynamic ensembles repeated twice, processed with nmrPipe38 and analyzed in
using experimental data. We found that I57V and L49I/I57V CcpNmr.39 We performed model-free analysis40,41 for the
increase the dynamics of CI2, whereas L49I decreases the calculations of backbone order parameters S2NH using the
dynamics. The corresponding change in conformational software relax.42 Values of R1, R2, and 1H-15N NOE for each
entropy scales linearly with the change in conformational amide group can be fitted using the model-free formalism with
stability. Our findings suggest that explicitly accounting for up to six parameters.43 Our initial analysis showed that fitting
changes in native-state entropy might be needed to improve to equations with more than three parameters resulted in
the predictions of the effect of mutations on protein stability, unrealistic values of the order parameters. In the final analysis
which may have implications for the interpretation of protein of the backbone NMR data, we therefore excluded equations
engineering and sequencing data.


with more than three parameters.
Side-chain 2H relaxation experiments were performed on
MATERIALS AND METHODS ∼0.7 mM 15N/13C/2H50%-labeled protein samples using pulse
Protein Production and Sample Preparation. We sequences described elsewhere.44−46 The relaxation data were
expressed wild-type CI2 (UniProt: P01053, residues 22−84 acquired on Bruker Avance III HD 600 and 750 MHz
with an additional N-terminal Met) and the three variants spectrometers using the following delays for RD1(Dz) and
L49I, I57V, and L49I/I57V using previously described DNA RD3(3Dz2−2): 1.4, 3.6, 7.6, 16.6, 33.8, and 50 ms; for RD2(D+):
constructs. 32 Escherichia coli BL21(DE3) carrying the 0.2, 4.7, 9.7, 12.5, 16.5, and 20 ms; and for RD4(D+2): 3, 6, 12,
expression plasmid was grown in an M9 minimal medium at 22, 34, and 50 ms. All side-chain relaxation experiments were
37 °C to OD600 = 0.6 before induction with 0.4 mM isopropyl done in duplicates. The relaxation data were processed using
β-D-1-thiogalactopyranoside. Four different isotope composi- python scripts, developed based on the scripts used by
tions of the M9 media were used. (i) 15NH4Cl as the sole Hoffmann et al.47 Similar to that study, only the original
nitrogen source, (ii) 13C6-glucose (Glc) as the sole carbon model-free equation (LS2)40,41 and LS3 equation (with the
B https://doi.org/10.1021/acs.biochem.1c00749
Biochemistry XXXX, XXX, XXX−XXX
Biochemistry pubs.acs.org/biochemistry Article

rate of global tumbling treated as a fitting parameter) were analyses were performed for the last 800 ns of each trajectory,
used to fit methyl relaxation data. The global rotational discarding the first 200 ns.
diffusion times of the studied proteins were initially estimated Calculations of Order Parameters from MD Trajecto-
based on backbone R2/R1 ratio with the programs R2R1_TM ries. The calculations of the order parameters S2NH and S2axis
and QUADRIC (www.palmer.hs.columbia.edu).48 Global were done following the procedures described by Hoffmann et
tumbling of all CI2 variants was analyzed with an isotropic al.,47 Briefly, we calculated the backbone time correlation
diffusion tensor since D∥/D⊥ < 1.3 for all the variants. function (TCF) of each NH bond vector up to a maximum
Chemical Shift Analysis. We calculated secondary lag-time of 400 ns. The resulting TCF was then fitted to two
chemical shifts for Cα relative to random coil chemical shifts different Lipari−Szabo models:
using the method of Kjaergaard and Poulsen, which corrects 1 2 −t / τR
for the effect of the neighboring residues (i − 2, i − 1, i + 1, C(t ) = Se
5 (2)
and i + 2, where i is the residue of interest) and temperature.49
Entropy Calculations Using Experimental Order with S used as fitting parameter and
2

Parameters. ΔS2NH and ΔS2axis values obtained in this study 1 2 −t / τR 1


were used to estimate the change in conformational entropy C(t ) = Se + (1 − S2)e−t / τ
5 5 (3)
upon the mutations by applying the “entropy meter”
approach:50 where τ −1 = τR −1 + τe−1 and with S2 and τe as fitting
parameters. C(t) was further multiplied by ξ = (1.02/
ΔS /R = 0.88Nres⟨δ ln(1 − S NH 2)⟩ − 0.56Nχ ⟨δSaxis 2⟩ (1) 1.041),6 to include the effect of vibrational averaging on
dipolar couplings.54 An isotropic diffusion model was applied
where ΔS is the change in conformational entropy, R is the gas using the experimentally derived global rotational diffusion
constant, Nres is the total number of residues, Nχ is the total times τR (see NMR methods). The obtained order parameter
number of side-chain χ angles, ⟨δSaxis2⟩ is the measured average S 2 NH was then averaged over the three independent
change in methyl axis order parameters, and finally ⟨δln(1 − trajectories. Model selection for each residue was done using
SNH2)⟩ is the corresponding quantity for measured changes in the Akaike Information Criterion (AIC):
the amide order parameters. The factors of 0.88 and 0.56
account for the reduction in entropy due to correlations of AIC = χ 2 + 2Nparams (4)
side-chain and backbone torsional motions and for the
dependence of entropy on order parameters. where χ estimates the difference between the fitted and
2

MD Simulations. The equations of motion were integrated experimental values of the relaxation rates and Nparams is the
using the leap-frog integrator implemented in GROMACS number of the fitting parameters in the model. The model with
v2018.1, using a time step of 2 fs. We performed the the lowest AIC was chosen.
simulations using the Amber a99SB-disp force field with the For the side-chain order parameter S2axis, each trajectory was
TIP4P-D-based a99SB-disp water model.51 Crystal structures first cut into 10 ns-long blocks. After that, internal TCFs
of the CI2 WT (PDB: 7A1H) and its mutated variants L49I (without global tumbling) were calculated for each Cmethyl−
(PDB: 7AOK), I57V (PDB: 7A3M), and L49I/I57V (PDB: Himethyl (i = 1, 2, and 3) bond vector of methyl-bearing side
7AON) were used as the starting conformations for the chains up to a maximum lag-time of 5 ns and averaged over the
simulations.32 The proteins were placed in a cubic box with a three Cmethyl−Himethyl (i = 1, 2, and 3) bonds. Next, the internal
volume of 30.66 nm3 (resulting in a 0.5 nm distance from the TCFs were fitted to six exponentials plus an offset;
protein to all box edges) and solvated with 6728 to 6836 water 6
molecules. The net charge of CI2 is zero, so no ions were Cint(t ) = ∑ A i e −t / τi
+ S long 2
needed to neutralize the system. The system was minimized i=1 (5)
with 50,000 steps of steepest descent and equilibrated for 150
where ∑6i=1Ai + Slong2 = 1, 0 ≤ Ai ≤ 1, 0 ≤ Slong2 ≤ 1, 0 ≤ τ1 ≤
ps in the NPT and NVT ensembles with harmonic position
restraints (force constants of 1000 kJ mol−1 nm−2) on the 10 ps, 0 ≤ τ2 ≤ 50 ps, 0 ≤ τ3 ≤ 200 ps, 0 ≤ τ4 ≤ 500 ps, 0 ≤ τ5
heavy atoms of the protein. Equilibration was performed using ≤ 1000 ps, τ6 ≥ 0 ps, and Slong2 is the long-time limit order
a velocity rescaling thermostat52 with a temperature of 298 K parameter. The internal TCF was then multiplied by a single
and a thermostat timescale of 0.1 ps, and the Parrinello− exponent using the experimentally derived global rotational
Rahman barostat (NPT) with a reference pressure of 1 bar, a diffusion times τR:
barostat time constant of 2 ps, and an isothermal C(t ) = Cint(t )e−t / τR (6)
compressibility of 4.5 × 10−5 bar−1. A cutoff range of 1.0 nm
was chosen for the van der Waals and Coulomb interactions, The resulting total TCF was then fitted to
and the Particle Mesh Ewald method53 was applied to treat
long-range electrostatics (fourth order cubic interpolation, 0.16 C(t ) = S2e−t / τR + (1 − S2)e−t / τred (7)
nm grid spacing). The production simulations were run in the τRτf
with S2 = 1/9Saxis2, τred = , and S2, τf as fitting parameters.
NVT ensemble, using a velocity rescaling thermostat52 with a τR + τf
time constant of 1 ps and a reference temperature of 298 K. Finally, all S2 were averaged over the 3 × 80 blocks.
We applied harmonic constrains to all bonds involving In order to assess the convergence of the calculations, we
hydrogens using the LINCS algorithm. All production runs focused on the side-chain S2axis,sim values since they, due to the
were 1 μs long and the protein coordinates were saved every 1 higher flexibility of the side chains, may converge more slowly
ps and three replicates were used for each simulation, summing than the backbone order parameters.55−57 We averaged the
to a total of 3 μs per structure. Since the initial structural 240 original blocks of 10 ns into 10 blocks of 240 ns and
fluctuations may be related to the equilibration process, all performed a leave-N-out cross-validation with N = 1,..., 5 to
C https://doi.org/10.1021/acs.biochem.1c00749
Biochemistry XXXX, XXX, XXX−XXX
Biochemistry pubs.acs.org/biochemistry Article

Figure 2. Chemical shift analysis. (A) Secondary Cα chemical shifts. (B) Absolute difference in methyl 13C CSs of the wild type (WT) and mutated

ÅÄÅ É2
variants of CI2. Mutation sites are highlighted with the asterisks. Secondary structure elements are depicted at the top of each panel.

ÅÅ Sexp, n 2 − ∑N wS 2Ñ ÑÑ
ÅÅ i=1 i r ,n Ñ ÑÑ ij w yz
;(w) = ∑ ÅÅ ÑÑ + θ∑ wi logjjj 0i zzz
evaluate the convergence of the side-chain order parameters.58

Å Ñ jw z
N N

n=1 Å ÅÅÇ (σnexp)2 + (σncalc)2 ÑÑÑÖ k i {


Specifically, for each N, we split the 10 blocks in a reference set
of 10-N and a test set of N blocks (the left-out blocks), i=1
averaged the S2axis,sim values within the two sets, and computed
= χ 2 (w) − θS(w) (8)
the RMSD between the resulting average values of the order
parameters. For the same N, this procedure was repeated for all where n ranges over the overall number of experimentally
the possible combinations of the left-out blocks, and the defined order parameters and N represents the number of
corresponding RMSDs are averaged to obtain ⟨RMSD⟩. As blocks that the trajectory has been cut into, σexp n is the
shown in Figure S1, the ⟨RMSD⟩ curves for N = 5 converge to experimental error of the nth order parameter, and σcalcn is the
values around 0.04, which is smaller than the typical difference standard error of the simulated order parameters, estimated
we computed between the S2axis,sim of the different protein over the different blocks. The optimal set of weights is
obtained as the corresponding minimum in weight space, w* =
variants. For this reason, we can consider our estimations of minw S2(w). Each weight is associated with a block from the
S2axis,sim converged. trajectory, and it encodes the relevance of each block to the
Reweighting Simulations Using S2NH and S2axis. The dynamical ensemble. As reference weights, we use wi0 = 1/N to
simulated trajectories were analyzed in a block-wise fashion represent the initial, uniform weights provided by the prior
using 100 non-overlapping blocks of 10 ns each. This resulted (original simulated ensemble). The parameter θ is used to set
in 300 blocks for the set of three 1 μs-long simulations of each the balance between our trust in the prior and in the
CI2 variant. The backbone S2NH and side-chain S2axis order experimental data. K-fold cross-validation was performed to
select a suitable range of θ and to prevent over fitting (Figures
parameters were calculated for each 10 ns block in a similar
S2 and S3). For this, the S2axis and S2NH data were split into
way as described above but with a maximum lag-time of 5 ns. training (80%) and test (20%) sets. Next, the training set was
The resulting TCF was then fitted only to the classical Lipari− reweighted against experimentally derived order parameters,
Szabo formalism (eq 5). Further, we performed the weighted and the obtained weights were then applied to the test set for
average of the order parameters over the blocks, where the cross-validation. The cross-validation was repeated k = 20
weights, w, are obtained through the optimization of a function times and for a range of values of θ. The parameter ϕeff(w) =
that determines the agreement between simulated and
experimental order parameters. The approach is similar to
N
exp(S(w)), with S(w) = −∑i = 1 wi log ( ), provides a meas-
wi
wi0
ure of the effective fraction of blocks retained for a given set of
previously described methods for fitting ensembles using order weights and can thus be used to guide the selection of a
parameters55,59 but using a reweighting approach similar to our suitable θ. To allow easy comparison between different CI2
recently described ABSURDer method.58 The functional form variants, θ = 200 was selected for all data sets. See the papers
for the target function that we here used is by Kümmerer et al.58 and Orioli et al.60 for further details.
D https://doi.org/10.1021/acs.biochem.1c00749
Biochemistry XXXX, XXX, XXX−XXX
Biochemistry pubs.acs.org/biochemistry Article

Figure 3. Difference in NMR-derived order parameters (for backbone N−H groups) between one of the CI2 variants and the average value of
three other CI2 variants: ΔS2NH = S2NH (variant) − S2NH (average-of-three-other-variants). Secondary structure elements are shown at the bottom
of each panel. Mutation sites are highlighted by asterisks. (A) ΔS2NH plots; (B) the difference in order parameters is shown as a color scale on the
structure of CI2.

■ RESULTS
Initially, we assessed the overall structural differences in the
NMR relaxation experiments and interpreted it with the
model-free (MF) formalism. We analyzed both the dynamics
four variants of CI2 by comparing the chemical shift of the backbone amide groups and of the side-chain methyl
differences of Cα atoms, in the form of secondary chemical groups for Ala, Ile, Leu, Thr, and Val. The side-chain methyls
shifts, and of the side-chain methyl groups (Figure 2). In of the two Met residues were excluded from the analysis due to
general, the differences in secondary chemical shifts are small, low data quality.
except for residues 49, 55, and 57 that show differences For the backbone 15N amides, we acquired R1, R2, and
between 1 and 2 ppm, which suggests that the backbone { H}-15N NOE relaxation parameters at two magnetic fields
1

conformation of these residues in solution could differ between (600 and 750 MHz) for each residue. The six experimental
the variants or engage in slow conformational exchange. There data points allowed us to sample the spectral densities, J(ω), at
is, however, no indication of slow conformational exchange in five different frequencies for each residue. We analyzed the
this part of the protein from our relaxation data (vide infra). In data with the model-free formalism, and to avoid overfitting,
the crystal structures, the backbone conformations of the four we excluded equations containing more than three parameters.
variants also do not differ in this region of the protein.32 The Since the initial analysis in relax showed that the ratio between
largest differences in the chemical shifts of the side-chain parallel and perpendicular components (D∥/D⊥) of the
methyl groups are observed for residues Val-13, Ala-16, Ile-20, rotational diffusion tensor was less than 1.3 for all four CI2
and Ile-29 that all pack against residue 49 and/or 57. Again, variants, we used an isotropic model for the global molecular
there is no significant change in the conformation of these four tumbling. For each CI2 variant, we were able to analyze the
amino acid residues in the crystal structures, and the chemical data from all non-proline backbone amides, except for K2. The
shift differences thus primarily reflect the change in the consistency of the relaxation data for the remaining residues
substitutions at positions 49 and 57. were assessed by comparing the values of the spectral densities
Backbone Dynamics. We assessed the conformational at zero frequency, J(0), calculated from data collected at either
dynamics on the ps−ns timescale of the four CI2 variants using of the two different fields. In the absence of μs−ms motions,
E https://doi.org/10.1021/acs.biochem.1c00749
Biochemistry XXXX, XXX, XXX−XXX
Biochemistry pubs.acs.org/biochemistry Article

Table 1. Total Average Difference in Order Parameters for Each CI2 Variant
ΔS2NH,total (backbone)a ΔS2axis,total (side chains)a
simulated simulated
variant experimental original reweighted experimental original reweighted
WT 0.1 ± 0.1 0.2 ± 0.1 −0.03 ± 0.05 0.7 ± 0.4 −0.7 ± 0.1 0.3 ± 0.2
L49I 0.8 ± 0.1 −0.1 ± 0.1 0.4 ± 0.1 −0.3 ± 0.2 −0.3 ± 0.1 −0.3 ± 0.2
I57V −0.11 ± 0.04 0.1 ± 0.1 0.14 ± 0.04 −0.6 ± 0.3 0.8 ± 0.2 0.4 ± 0.2
L49I/I57V −0.8 ± 0.1 −0.03 ± 0.09 −0.61 ± 0.05 0.1 ± 0.2 0.2 ± 0.1 −0.7 ± 0.2
a
The values are calculated as a difference between a variant and the average of three other variants; simulated data are presented before and after
the reweighting procedure; the uncertainty is presented as the standard error of the mean.

Figure 4. Difference in NMR-derived squared generalized order parameters (for side-chain methyl groups) between one of the CI2 variants and the
average value of three other CI2 variants: ΔS2axis = S2axis (variant) − S2axis (average-of-three-other-variants). The borders of the structural elements
are depicted at the bottom of each panel, and mutation sites are highlighted by asterisks.

the J(0) values calculated from the two data sets are expected We compared the S2NH values between the variants of CI2 to
to be the same.61 For all four CI2 variants, J(0) values from the assess the effects of the mutations on the dynamics. To
two fields are highly correlated (Figure S4). Most of the minimize the effects of outliers, we calculated the ΔS2NH values
deviating amide groups are also characterized by elevated R1R2 as the difference between the S2NH values for each variant and
values. Unlike the R2/R1 ratio, the R1R2 product includes the the average values of the other three variants, as recently
effects of chemical exchange but not the effects of motional suggested.63 For example, in the case of WT, ΔS2WT = S2WT −
anisotropy (Figure S5).62 We note that a high correlation ⟨S2⟩L49I,I57V,L49I/I57V. The analysis demonstrates that the differ-
between the R2/R1 ratio and the R1R2 product suggests little ences in ps−ns conformational dynamics overall are small;
motional anisotropy and thus justifies an isotropic model for however, we do see some sequence-dependent variation
(Figure 3). ΔS2WT is positive for the first 11 residues, residues
the global molecular tumbling of CI2 variants.
26E and 27A, and in the region close to the mutation site 57
Overall, the order parameters, S2NH, for the backbone amides
(53K-56N and 59Q-60V), suggesting smaller than average
are quite similar for all variants with a value around 0.8 in the
amplitude of the backbone dynamics of WT for these residues.
structured regions of the protein, which also include the C- For most of the other residues, ΔS2WT are around 0.01−0.02
terminal residue as the C-terminal carboxylate engage in a salt- units lower than the average values of the other three variants.
bridge with Arg-46. S2NH drops to ∼0.7 in the more flexible The destabilizing L49I mutation results in a decrease in the
inhibitory loop around residues I37−I44 (Figure S6A). The dynamics of almost all amide groups in CI2: most of ΔS2L49I
lowest value, indicating greatest flexibility, is found at residue values are in between 0.01−0.2 units. The stabilizing I57V
M40 that plays an important functional role in chymotrypsin mutation results in a distribution of ΔS2I57V values that almost
inhibition. We note that several residues around this region, mirrors that for WT, namely, positive ΔS2I57V for the first 11
according to their elevated R1R2 values, experience slower residues and for residues 26E, 27A, 53K-56N, and 59Q-60V.
dynamics. The I57V/L49I variant is characterized by the increased
F https://doi.org/10.1021/acs.biochem.1c00749
Biochemistry XXXX, XXX, XXX−XXX
Biochemistry pubs.acs.org/biochemistry Article

dynamics for almost all backbone amide groups. The summed


(excluding the mutation sites) differences in the backbone
N
order parameters, ΔS2total = ∑n = 1 ΔSi2 (where i is an amide
group and N is the total number of amide groups), are listed in
Table 1.
Side-Chain Dynamics. We assigned the chemical shifts for
all side-chain methyl groups of WT CI2. However, some were
excluded from the relaxation analysis. For L8 2H−Cδ2, V31
2
H−Cγ1, and V63 2H−Cγ1, it was not possible to estimate the
peak volume or the maximum peak height due to significant
overlap. Moreover, the chemical shifts for L8-Cδ1, L32-Cδ1, and
L49-Cδ1 (in cases of WT and I57V) were greater than 24 ppm.
This may result in strong 13C−13C coupling between Cγ and Figure 5. Correlation between changes in thermodynamic stability of
Cδ in leucine residues and the 2H relaxation data can be CI2 variant values from unfolding experiments (ΔΔGf) and changes
compromised.47 In total, we included 42 peaks in the analysis in the entropy contribution from ps−ns dynamics to free energy
of side-chain ps−ns dynamics. (TΔΔSbb+sch).
Fast dynamics of side-chain CH2D isotopomers can be
tracked based on two to five relaxation rates. In this study, we
disordered elements without sacrificing accuracy for structured
acquired four side-chain relaxation rates at two magnetic fields
regions. This force field was chosen due to the possible higher
(600 and 750 MHz):45,46 RD1(Dz), longitudinal magnetization;
flexibility of several residues in the inhibitory CI2 loop. For all
RD2(D+), in-phase transverse magnetization; RD3(3Dz2−2),
CI2 variants, we conducted three simulations of 1 μs each,
quadrupolar order magnetization; and RD4(D+2), double
obtaining a total of 3 μs of sampling per variant.
quantum magnetization. Collection of these four relaxation
As two rotamers of I49 are observed in the crystal structure
rates results in an overdetermination since they depend on the of the L49I/I57V variant (PDB: 7AON), we ran two
spectral densities at only three different frequencies (zero independent simulations starting from each of the rotamers.
frequency J(0), the deuterium frequency at the given field In the first rotamer structure, l49I/I57V(A), the I49 χ1 angle
strength J(ωD), and double frequency of deuterium J(2ωD)).45 (N−Cα−Cβ−Cγ) is in a gauche− conformation, and in the
Therefore, according to the relationship RD4 = 1/2RD1 + 1/ second structure, l49I/I57V(B), the I49 χ1 angle is in a gauche
6RD3, we found that the consistency of the data sets at both + conformation. For L49I/I57V, the I49 χ1 angle populates the
magnetic fields is good (R2 ≥ 0.97, Figure S7). All acquired gauche− rotamer for 95% of the time and the gauche+ for the
side-chain deuterium relaxation rates were fitted to the LS2 remaining 5%. This behavior is independent of the initial
and LS3 model-free formalisms. rotameric state. I49 χ1 flips from gauche+ to gauche− within
As in the case of backbone amide order parameters, side- the first 15 ns of the simulations, which indicated that the force
chain methyl S2axis values are quite similar for all the variants field prefers the gauche− conformation of I49 χ1. We therefore
(Figure S6B). Similar to S2NH, we calculated ΔS2axis values as only consider l49I/I57V(A) in the following. Initial analysis of
the difference between S2axis values of each of the variants and the simulated ensembles revealed no changes in the overall
the average values of other three variants (Figure 4). A direct structure of either of the variants: the average all-atom RMSD
comparison suggests that there may be subtle changes in the is in the range of 0.26−0.29 nm and per-residue root-mean-
ps−ns dynamics of side-chain methyl groups. However, the square fluctuation (RMSF) does not exceed 0.6 nm for the
errors are close to the value of ΔS2axis. In general, according to most flexible residue (M40) in all five systems.
the total summed ΔS2axis, side-chain methyl groups of the three Simulation-Derived Order Parameters. For a more
mutants are more flexible than the wild type (Table 1). detailed comparison of NMR and MD data, we analyzed the
Entropy Calculations. To obtain a quantitative measure of acquired trajectories in terms of backbone amide and side-
the effects of the mutations, we used the ΔS2NH and ΔS2axis chain methyl order parameters. As for the experimental data,
values to estimate the change in the conformational entropy by we used the model-free formalism. For consistency with
applying the “entropy meter” approach.64,50 The entropy experimental data, the extended model-free formalisms were
contribution to the free energy calculated in this way correlates omitted. The impact of the amount of sampling on the values
well with experimental ΔΔGf values (Pearson’s r = −0.96; of the order parameters was checked by employing a leave-N-
Figure 5). The individual contributions from the backbone and out cross-validation. The analysis suggests that our estimates of
the side chains are of the same order of magnitude, although the order parameters are converged (Figure S1).
most of the error in the total value originate from the side- S2NH,sim values are quite similar for all CI2 variants and
chain data (Table S1). We note that the variation in the reflect the absence of any major effect of the mutations.
calculated TΔS is larger than the variation in ΔΔGf that could However, the correlations between S2NH,sim and S2NH for the
either result from entropy-enthalpy compensation or because four variants are not strong (Pearson’s correlation coefficient, r
the “entropy meter” overestimates the entropy in our case. = 0.55−0.62; Figure S8A). This may be related to several
Still, the conclusion from the strong correlation is that changes issues. First, the simulations may not be converged, though our
in conformational entropy are important for the changes in analysis described above suggests this to be a minor effect.
conformational stability we observe for CI2. Second, the 15N transverse relaxation rates can be significantly
Molecular Dynamics Simulations. We performed MD affected by μs−ms timescale dynamics itself (Rex contribution,
simulations of the CI2 variants using a99SB-disp with the Figure S5), and it is not always easy to accurately correct for
modified TIP4PD water model, 51 a force field that these effects.65,66 Third, despite the application of a state-of-
substantially improves the accuracy for simulations of the the-art force field, imperfections can still reduce the match
G https://doi.org/10.1021/acs.biochem.1c00749
Biochemistry XXXX, XXX, XXX−XXX
Biochemistry pubs.acs.org/biochemistry Article

Figure 6. (A) Experimental backbone S2NH and (B) side-chain S2axis order parameters of CI2 compared to their reweighted computational
counterparts extracted from MD simulations. Pearson correlations (r) and RMSDs were calculated for each variant separately.

between simulations and experiments. These limitations of the Similar to the experimental data, we compared ΔS2NH,sim values
computational approach may significantly affect not only the calculated as a difference between S2NH,sim values of each of the
absolute values of the order parameters but also the difference variants and the average values of other three variants.63 The
in the dynamics of CI2 variants. As in case of the backbone, ΔS2NH,sim,total values get closer to the experimental ΔS2NH,total
limited agreement between NMR and MD side-chain order values and better reflect experimentally observed changes in
parameters (S2axis,sim) may be related to the force field the order parameters (with an exception of I57V; Table 1). In
imperfections and was observed in other studies (Figure the case of the side-chain methyl order parameters (S2axis,sim)
S8B).47,67 after reweighting, r and RMSD also improve (from r = 0.55−
Reweighting MD Trajectories. It was shown previously 0.69, RMSD = 0.27−0.30 to r = 0.62−0.87, RMSD = 0.12−
that the direct integration of experimental data with molecular 0.27; Figure 6B).
simulations can significantly improve a detailed description of The trajectory reweighting results in a significant improve-
the structural and dynamic properties of biomolecules. Such ment of the correlation between experimentally measured and
integration can be performed a priori using MD simulations MD-derived backbone and/or side-chain order parameters for
biased with experimental data or by reweighting unbiased all CI2 variants, except I57V. However, even after the
simulations using experimental data a posteriori.68 The second reweighting, the MD simulations do not capture all small
approach allows us to reweight the simulated trajectories also variations in the dynamics of the four CI2 variants.
using our NMR relaxation data.60 The relaxation rates are used Consequently, we omitted calculations of conformational
in the fitting procedure to obtain S2NH and S2axis and can be entropy from the simulations. Similar to the analysis of the
considered as a source of experimental data.47,67 Due to slow experimental data (Figures 3 and 4), the reweighted computa-
μs−ms timescale motions for some residues in CI2, the R2 rate tional ensembles did not reveal any clear changes in the
values can be affected by exchange (Figure S5). This structural and dynamic properties of CI2. This suggests that
contribution is not captured in MD simulations, which can the changes in stability and conformational entropy are the
result in overfitting during the reweighting procedure if we combined effect of several subtle contributions. In accordance
directly target the relaxation rates. Consequently, instead, we with the previous study of CI2 dynamics,26 the mutations thus
reweighted our simulated trajectories using S2NH and S2axis affect the conformational dynamics of all 64 amino acid
residues as a whole, probably by introducing many small
values. We used an approach inspired by our recently
changes in the contact network of the protein.
developed ABSURDer method58 but targeting the order
parameters as also done previously.55,59 For each CI2 variant,
S2NH,sim and S2axis,sim values were calculated separately for each
of the 300 blocks of 10 ns and further reweighted against the
■ DISCUSSION
We have demonstrated that ps−ns conformational dynamics of
experimental S2NH and S2axis values. CI2 can be affected by mutations in the hydrophobic core and
The reweighting procedure resulted in a noticeable found that the level of the conformational entropy correlates
improvement of the agreement between the experimental with ΔΔGf. Based on the NMR measurements, we suggest that
and computational data (Figure 6). Pearson’s correlation the entropy and, in particular, the backbone entropy provides a
coefficient (r) and RMSD for the backbone S2sim and S2NH major contribution to the stability of the CI2 variants analyzed
improve for all CI2 variants (from r = 0.45−0.62, RMSD = in this work. From the entropy meter estimates, the increase in
0.05−0.08 to r = 0.59−0.71, RMSD = 0.04−0.05; Figure 6A). entropy is even larger than the gain in conformational stability,
H https://doi.org/10.1021/acs.biochem.1c00749
Biochemistry XXXX, XXX, XXX−XXX
Biochemistry pubs.acs.org/biochemistry Article

suggesting that the mutations could be associated with an correlations of relaxation rates. (Figure S7), correlation
unfavorable loss of intramolecular interactions. We note, between experimental order parameters and order
however, complications arising from entropic contributions parameters from non-reweighted MD (Figure S8), and
arising from dynamics on timescales not probed by the entropy changes from entropy-meter calculations (Table
relaxation measurements69 and which may not be accurately S1) (PDF)
captured by the entropy meter approach. Our analysis of the
crystal structures of the four CI2 variants showed that the Deuterium relaxation rates for CH2D groups (XLSX)
structures are very similar and only subtle structural changes
are observed throughout the protein.32 This picture is mirrored Accession Codes
by our current NMR analysis. The very small changes in CI2: P01053
dynamics between the variants are thus scattered throughout
the protein. The largest difference is near the N-terminal for
residues 5−9 that form a helical turn. This region appears the
■ AUTHOR INFORMATION
Corresponding Authors
most rigid in the wild type and most dynamic in the I57V
Kresten Lindorff-Larsen − Structural Biology and NMR
variant. We speculate that the packing between the residue at
Laboratory and the Linderstrøm-Lang Centre for Protein
positions 57 and V9 is slightly weakened by the presence of a
Science, Department of Biology, University of Copenhagen,
Val at position 59 and that this effect is alleviated by having an
2200 Copenhagen N, Denmark; orcid.org/0000-0002-
Ile at position 49. However, we do not see any structural
4750-6039; Phone: +45 35 32 20 27; Email: lindorff@
rearrangements around V9 or any other residues in the helical
bio.ku.dk
turn.
Kaare Teilum − Structural Biology and NMR Laboratory and
Our MD simulations of the CI2 variants support our
the Linderstrøm-Lang Centre for Protein Science, Department
findings from the experimental data, i.e., differences in the
of Biology, University of Copenhagen, 2200 Copenhagen N,
dynamics are subtle and distributed throughout the protein,
Denmark; orcid.org/0000-0001-6919-1982;
although the simulations cannot fully reproduce our
Phone: +45 35 32 20 29; Email: kaare.teilum@bio.ku.dk
experimental data. The difference between NMR and MD
derived dynamic properties may be related to the presence of Authors
fluctuations in the native state of CI2 and its variants, which Yulian Gavrilov − Structural Biology and NMR Laboratory
can affect fast protein dynamics but cannot be intensively and the Linderstrøm-Lang Centre for Protein Science,
sampled by relatively short simulations. This assumption may Department of Biology, University of Copenhagen, 2200
be supported by a previous computational study of CI2.70 In Copenhagen N, Denmark
that work, MD simulations, restrained with protection factors Felix Kümmerer − Structural Biology and NMR Laboratory
for local H-exchange, revealed a native-state ensemble of CI2 and the Linderstrøm-Lang Centre for Protein Science,
including rare, large fluctuations from the native state, which Department of Biology, University of Copenhagen, 2200
may take place on a timescale of milliseconds or more and Copenhagen N, Denmark
cannot be related to the folding pathway of CI2.70 This is Simone Orioli − Structural Biology and NMR Laboratory
consistent with our experimental R1R2 values suggesting that and the Linderstrøm-Lang Centre for Protein Science,
mutations of CI2 may also affect the dynamics on slower (μs− Department of Biology, University of Copenhagen, 2200
ms) timescales. Copenhagen N, Denmark; Structural Biophysics, Niels Bohr
Our best explanation for increased stability is thus that the Institute, University of Copenhagen, 2100 Copenhagen Ø,
mutations release the previously proposed unfavorable strain in Denmark
the hydrophobic mini-core of A16, L49, and I57,71 and that Andreas Prestel − Structural Biology and NMR Laboratory
this occurs without any actual rearrangement of the structure and the Linderstrøm-Lang Centre for Protein Science,
and thus of the enthalpic interactions that stabilize the folded Department of Biology, University of Copenhagen, 2200
state. This is much in line with the conclusions based on a Copenhagen N, Denmark; orcid.org/0000-0002-5459-
structural analysis of the I57V mutant by Fersht and co- 9608
workers.72
In summary, we have demonstrated that in the case of CI2 Complete contact information is available at:
and its mutated variants, changes in fast ps−ns dynamics may https://pubs.acs.org/10.1021/acs.biochem.1c00749
significantly contribute to changes in protein stability. Our
analysis suggests that also slower μs−ms timescale dynamics of Funding
CI2 may be affected by these mutations. This work was supported by the following grants: Carlsberg


Foundation (postdoc grant no. CF17-0491 to Y.G.), the
ASSOCIATED CONTENT Lundbeck Foundation to the BRAINSTRUC structural
biology initiative (155-2015-2666, to K.L.-L.), the NordForsk
*
sı Supporting Information
Nordic Neutron Science Programme (to K.L.-L.), the Novo
The Supporting Information is available free of charge at Nordisk Foundation to the NMR infrastructure facility,
https://pubs.acs.org/doi/10.1021/acs.biochem.1c00749. cOpenNMR, (grant no. NNF18OC0032996) and to the
Convergence of MD-derived side-chain order parame- ROBUST Resource for Biomolecular Simulations (grant no.
ters (Figure S1), cross validation of reweighting of MD- NNF18OC0032608). We also acknowledge access to the
derived order parameters (Figures S2 and S3), Biocomputing Core Facility at the Department of Biology,
correlations of spectral densities (Figure S4), R2/R1 vs University of Copenhagen.
R1R2 and exchange outliers (Figure S5), overview of Notes
order parameters from NMR experiments (Figure S6), The authors declare no competing financial interest.
I https://doi.org/10.1021/acs.biochem.1c00749
Biochemistry XXXX, XXX, XXX−XXX
Biochemistry pubs.acs.org/biochemistry Article

The assigned chemical shifts and backbone relaxation data (17) Kovermann, M.; Rogne, P.; Wolf-Watz, M. Protein dynamics
have been deposited at the BioMagResBank (http://bmrb.io) and function from solution state NMR spectroscopy. Q. Rev. Biophys.
with accession numbers 51230, 51234, 51235, and 51236. 2016, 49, No. e6.


(18) McPhalen, C. A.; James, M. N. G. Crystal and molecular
structure of the serine proteinase inhibitor CI-2 from barley seeds.
ACKNOWLEDGMENTS Biochemistry 1987, 26, 261−269.
The authors would like to thank Falk Hoffmann for the help (19) Itzhaki, L. S.; Otzen, D. E.; Fersht, A. R. The Structure of the
Transition State for Folding of Chymotrypsin Inhibitor 2 Analysed by
with the calculations of order parameters from MD simulations
Protein Engineering Methods: Evidence for a Nucleation-condensa-
and Yong Wang for the assistance with the setup of MD tion Mechanism for Protein Folding. J. Mol. Biol. 1995, 254, 260−288.
simulations. (20) Ladurner, A. G.; Itzhaki, L. S.; Daggett, V.; Fersht, A. R.

■ ABBREVIATIONS
CI2, chymotrypsin inhibitor 2; MD, molecular dynamics; MF,
Synergy between simulation and experiment in describing the energy
landscape of protein folding. Proc. Natl. Acad. Sci. U. S. A. 1998, 95,
8473−8478.
(21) Elcock, A. E. Molecular Simulations of Cotranslational Protein
model-free; TCF, time correlation function Folding: Fragment Stabilities, Folding Cooperativity, and Trapping in


the Ribosome. PLoS Comput. Biol. 2006, 2, No. e98.
REFERENCES (22) Matouschek, A.; Otzen, D. E.; Itzhaki, L. S.; Jackson, S. E.;
Fersht, A. R. Movement of the position of the transition state in
(1) Stein, A.; Fowler, D. M.; Hartmann-Petersen, R.; Lindorff- protein folding. Biochemistry 1995, 34, 13656−13662.
Larsen, K. Biophysical and Mechanistic Models for Disease-Causing (23) Neira, J. L.; Itzhaki, L. S.; Otzen, D. E.; Davis, B.; Fersht, A. R.
Protein Variants. Trends Biochem. Sci. 2019, 44, 575−588. Hydrogen exchange in chymotrypsin inhibitor 2 probed by muta-
(2) Morley, K. L.; Kazlauskas, R. J. Improving enzyme properties: genesis. J. Mol. Biol. 1997, 270, 99−110.
when are closer mutations better? Trends Biotechnol. 2005, 23, 231− (24) Fersht, A. R. Transition-state structure as a unifying basis in
237. protein-folding mechanisms: Contact order, chain topology, stability,
(3) Manning, M. C.; Chou, D. K.; Murphy, B. M.; Payne, R. W.; and the extended nucleus mechanism. Proc. Natl. Acad. Sci. U. S. A.
Katayama, D. S. Stability of Protein Pharmaceuticals: An Update. 2000, 97, 1525−1529.
Pharm. Res. 2010, 27, 544−575. (25) Whitley, M. J.; Zhang, J.; Lee, A. L. Hydrophobic Core
(4) Karshikoff, A.; Nilsson, L.; Ladenstein, R. Rigidity versus Mutations in CI2 Globally Perturb Fast Side-Chain Dynamics
flexibility: the dilemma of understanding protein thermal stability. Similarly without Regard to Position. Biochemistry 2008, 47, 8566−
FEBS J. 2015, 282, 3899−3917. 8576.
(5) Gavrilov, Y.; Dagan, S.; Reich, Z.; Scherf, T.; Levy, Y. An NMR (26) Whitley, M. J.; Lee, A. L. Exploring the role of structure and
Confirmation for Increased Folded State Entropy Following Loop dynamics in the function of chymotrypsin inhibitor 2. Proteins 2011,
Truncation. J. Phys. Chem. B 2018, 122, 10855−10860. 79, 916−924.
(6) Bigman, L. S.; Levy, Y. Entropic Contributions to Protein (27) Shakhnovich, E.; Abkevich, V.; Ptitsyn, O. Conserved residues
Stability. Isr. J. Chem. 2020, 60, 705−712. and the mechanism of protein folding. Nature 1996, 379, 96−98.
(7) Teilum, K.; Olsen, J. G.; Kragelund, B. B. Protein stability, (28) Dokholyan, N. V.; Li, L.; Ding, F.; Shakhnovich, E. I.
flexibility and function. Biochim. Biophys. Acta 2011, 1814, 969−976. Topological determinants of protein folding. Proc. Natl. Acad. Sci. U.
(8) Mittermaier, A.; Kay, L. E. The response of internal dynamics to S. A. 2002, 99, 8637−8641.
hydrophobic core mutations in the SH3 domain from the Fyn (29) Jiao, X.; Chang, S.; Li, C.; Chen, W.; Wang, C. Construction
tyrosine kinase. Protein Sci. 2004, 13, 1088−1099. and application of the weighted amino acid network based on energy.
(9) Berezovsky, I. N.; Chen, W. W.; Choi, P. J.; Shakhnovich, E. I. Phys. Rev. E 2007, 75, No. 051903.
Entropic Stabilization of Proteins and Its Proteomic Consequences. (30) Hilser, V. J.; Dowdy, D.; Oas, T. G.; Freire, E. The structural
PLoS Comput. Biol. 2005, 1, 11. distribution of cooperative interactions in proteins: Analysis of the
(10) Solomentsev, G.; Diehl, C.; Akke, M. Conformational Entropy native state ensemble. Proc. Natl. Acad. Sci. U. S. A. 1998, 95, 9903−
of FK506 Binding to FKBP12 Determined by Nuclear Magnetic 9908.
Resonance Relaxation and Molecular Dynamics Simulations. Bio- (31) Atilgan, A. R.; Akan, P.; Baysal, C. Small-World Communica-
chemistry 2018, 57, 1451−1461. tion of Residues and Significance for Protein Dynamics. Biophys. J.
(11) Verteramo, M. L.; Stenström, O.; Ignjatović, M. M.; Caldararu, 2004, 86, 85−91.
O.; Olsson, M. A.; Manzoni, F.; Leffler, H.; Oksanen, E.; Logan, D. (32) Hamborg, L.; Granata, D.; Olsen, J. G.; Roche, J. V.; Pedersen,
T.; Nilsson, U. J.; Ryde, U.; Akke, M. Interplay between Conforma- L. E.; Nielsen, A. T.; Lindorff-Larsen, K.; Teilum, K. Synergistic
tional Entropy and Solvation Entropy in Protein−Ligand Binding. J. stabilization of a double mutant in chymotrypsin inhibitor 2 from a
Am. Chem. Soc. 2019, 141, 2012−2026. library screen in E. coli. Commun. Biol. 2021, 4, 980.
(12) Evenäs, J.; Malmendal, A.; Akke, M. Dynamics of the (33) Schymkowitz, J.; Borg, J.; Stricher, F.; Nys, R.; Rousseau, F.;
Transition between Open and Closed Conformations in a Calm- Serrano, L. The FoldX web server: an online force field. Nucleic Acids
odulin C-Terminal Domain Mutant. Structure 2001, 9, 185−195. Res. 2005, 33, W382−W388.
(13) Millet, O.; Mittermaier, A.; Baker, D.; Kay, L. E. The Effects of (34) Das, R.; Baker, D. Macromolecular Modeling with Rosetta.
Mutations on Motions of Side-chains in Protein L Studied by 2H Annu. Rev. Biochem. 2008, 77, 363−382.
NMR Dynamics and Scalar Couplings. J. Mol. Biol. 2003, 329, 551− (35) Kazimierczuk, K.; Orekhov, V. Y. Accelerated NMR Spectros-
563. copy by Using Compressed Sensing. Angew. Chem., Int. Ed. 2011, 50,
(14) Siggers, K.; Soto, C.; Palmer, A. G., III Conformational 5556−5559.
Dynamics in Loop Swap Mutants of Homologous Fibronectin Type (36) Neri, D.; Szyperski, T.; Otting, G.; Senn, H.; Wuethrich, K.
III Domains. Biophys. J. 2007, 93, 2447−2456. Stereospecific nuclear magnetic resonance assignments of the methyl
(15) Kleckner, I. R.; Foster, M. P. An introduction to NMR-based groups of valine and leucine in the DNA-binding domain of the 434
approaches for measuring protein dynamics. Biochim. Biophys. Acta repressor by biosynthetically directed fractional carbon-13 labeling.
2011, 1814, 942−968. Biochemistry 1989, 28, 7510−7516.
(16) Stafford, K. A.; Robustelli, P.; Palmer, A. G., III Thermal (37) Farrow, N. A.; Muhandiram, R.; Singer, A. U.; Pascal, S. M.;
Adaptation of Conformational Dynamics in Ribonuclease H. PLoS Kay, C. M.; Gish, G.; Shoelson, S. E.; Pawson, T.; Forman-Kay, J. D.;
Comput. Biol. 2013, 9, No. e1003218. Kay, L. E. Backbone Dynamics of a Free and a Phosphopeptide-

J https://doi.org/10.1021/acs.biochem.1c00749
Biochemistry XXXX, XXX, XXX−XXX
Biochemistry pubs.acs.org/biochemistry Article

Complexed Src Homology 2 Domain Studied by 15N NMR (56) Genheden, S.; Akke, M.; Ryde, U. Conformational Entropies
Relaxation. Biochemistry 1994, 33, 5984−6003. and Order Parameters: Convergence, Reproducibility, and Trans-
(38) Delaglio, F.; Grzesiek, S.; Vuister, G. W.; Zhu, G.; Pfeifer, J.; ferability. J. Chem. Theory Comput. 2014, 10, 432−438.
Bax, A. NMRPipe: A multidimensional spectral processing system (57) Allnér, O.; Foloppe, N.; Nilsson, L. Motions and Entropies in
based on UNIX pipes. J. Biomol. NMR 1995, 6, 277−293. Proteins as Seen in NMR Relaxation Experiments and Molecular
(39) Skinner, S. P.; Fogh, R. H.; Boucher, W.; Ragan, T. J.; Dynamics Simulations. J. Phys. Chem. B 2015, 119, 1114−1128.
Mureddu, L. G.; Vuister, G. W. CcpNmr AnalysisAssign: a flexible (58) Kümmerer, F.; Orioli, S.; Harding-Larsen, D.; Hoffmann, F.;
platform for integrated NMR analysis. J. Biomol. NMR 2016, 66, 111− Gavrilov, Y.; Teilum, K.; Lindorff-Larsen, K. Fitting Side-Chain NMR
124. Relaxation Data Using Molecular Simulations. J. Chem. Theory
(40) Lipari, G.; Szabo, A. Model-free approach to the interpretation Comput. 2021, 17, 5262−5275.
of nuclear magnetic resonance relaxation in macromolecules. 1. (59) Lindorff-Larsen, K.; Best, R. B.; Depristo, M. A.; Dobson, C.
Theory and range of validity. J. Am. Chem. Soc. 1982, 104, 4546− M.; Vendruscolo, M. Simultaneous determination of protein structure
and dynamics. Nature 2005, 433, 128−132.
4559.
(60) Orioli, S.; Larsen, A. H.; Bottaro, S.; Lindorff-Larsen, K. How
(41) Lipari, G.; Szabo, A. Model-free approach to the interpretation
to learn from inconsistencies: Integrating molecular simulations with
of nuclear magnetic resonance relaxation in macromolecules. 2.
experimental data. In Progress in Molecular Biology and Translational
Analysis of experimental results. J. Am. Chem. Soc. 1982, 104, 4559− Science; Academic Press: Strodel, B., Barz, B., Eds., 2020, pp 123−
4570. 176..
(42) d’Auvergne, E. J.; Gooley, P. R. Optimisation of NMR dynamic (61) Morin, S.; Gagné, S. M. Simple tests for the validation of
models I. Minimisation algorithms and their performance within the multiple field spin relaxation data. J. Biomol. NMR 2009, 45, 361−372.
model-free and Brownian rotational diffusion spaces. J. Biomol. NMR (62) Kneller, J. M.; Lu, M.; Bracken, C. An Effective Method for the
2008, 40, 107−119. Discrimination of Motional Anisotropy and Chemical Exchange. J.
(43) Clore, G. M.; Szabo, A.; Bax, A.; Kay, L. E.; Driscoll, P. C.; Am. Chem. Soc. 2002, 124, 1852−1853.
Gronenborn, A. M. Deviations from the simple two-parameter model- (63) Wallerstein, J. Molecular recognition and dynamics in proteins
free approach to the interpretation of nitrogen-15 nuclear magnetic studied by NMR. Ph.D. Dissertation, Lund University, Lund. 2019.
relaxation of proteins. J. Am. Chem. Soc. 1990, 112, 4989−4991. (64) Kasinath, V.; Sharp, K. A.; Wand, A. J. Microscopic Insights
(44) Muhandiram, D. R.; Yamazaki, T.; Sykes, B. D.; Kay, L. E. into the NMR Relaxation-Based Protein Conformational Entropy
Measurement of 2H T1 and T1. rho. Relaxation Times in Uniformly Meter. J. Am. Chem. Soc. 2013, 135, 15092−15100.
13C-Labeled and Fractionally 2H-Labeled Proteins in Solution. J. Am. (65) Palmer, A. G. NMR Characterization of the Dynamics of
Chem. Soc. 1995, 117, 11536−11544. Biomacromolecules. Chem. Rev. 2004, 104, 3623−3640.
(45) Millet, O.; Muhandiram, D. R.; Skrynnikov, N. R.; Kay, L. E. (66) Jarymowycz, V. A.; Stone, M. J. Fast Time Scale Dynamics of
Deuterium Spin Probes of Side-Chain Dynamics in Proteins. 1. Protein Backbones: NMR Relaxation Methods, Applications, and
Measurement of Five Relaxation Rates per Deuteron in 13C-Labeled Functional Consequences. Chem. Rev. 2006, 106, 1624−1671.
and Fractionally 2H-Enriched Proteins in Solution. J. Am. Chem. Soc. (67) Hoffmann, F.; Mulder, F. A. A.; Schäfer, L. V. Predicting NMR
2002, 124, 6439−6448. relaxation of proteins from molecular dynamics simulations with
(46) Skrynnikov, N. R.; Millet, O.; Kay, L. E. Deuterium Spin Probes accurate methyl rotation barriers. J. Chem. Phys. 2020, 152,
of Side-Chain Dynamics in Proteins. 2. Spectral Density Mapping and No. 084102.
(68) Bottaro, S.; Lindorff-Larsen, K. Biophysical experiments and
Identification of Nanosecond Time-Scale Side-Chain Motions. J. Am.
biomolecular simulations: A perfect match? Science 2018, 361, 355−
Chem. Soc. 2002, 124, 6449−6460.
360.
(47) Hoffmann, F.; Xue, M.; Schäfer, L. V.; Mulder, F. A. A.
(69) Hoffmann, F.; Mulder, F. A. A.; Schäfer, L. V. How much
Narrowing the gap between experimental and computational entropy is contained in NMR relaxation parameters? preprint.
determination of methyl group dynamics in proteins. Phys. Chem. ChemRxiv 2021, DOI: 10.33774/chemrxiv-2021-l686w.
Chem. Phys. 2018, 20, 24577−24590. (70) Best, R. B.; Vendruscolo, M. Structural Interpretation of
(48) Lee, L. K.; Rance, M.; Chazin, W. J.; Palmer, A. G., III Hydrogen Exchange Protection Factors in Proteins: Characterization
Rotational diffusion anisotropy of proteins from simultaneous analysis of the Native State Fluctuations of CI2. Structure 2006, 14, 97−106.
of 15N and 13Cα nuclear spin relaxation 12. J. Biomol. NMR 1997, (71) Ladurner, A. G.; Itzhaki, L. S.; Fersht, A. R. Strain in the folding
287−298. nucleus of chymotrypsin inhibitor 2. Folding Des. 1997, 2, 363−368.
(49) Kjaergaard, M.; Poulsen, F. M. Sequence correction of random (72) Jackson, S. E.; Moracci, M.; elMasry, N.; Johnson, C. M.;
coil chemical shifts: correlation between neighbor correction factors Fersht, A. R. Effect of cavity-creating mutations in the hydrophobic
and changes in the Ramachandran distribution. J. Biomol. NMR 2011, core of chymotrypsin inhibitor 2. Biochemistry 1993, 32, 11259−
50, 157−165. 11269.
(50) Sharp, K. A.; O’Brien, E.; Kasinath, V.; Wand, A. J. On the
relationship between NMR-derived amide order parameters and
protein backbone entropy changes: Measurement of Protein Back-
bone Entropy by NMR. Proteins 2015, 83, 922−930.
(51) Robustelli, P.; Piana, S.; Shaw, D. E. Developing a molecular
dynamics force field for both folded and disordered protein states.
Proc. Natl. Acad. Sci. U. S. A. 2018, 115, E4758−E4766.
(52) Bussi, G.; Donadio, D.; Parrinello, M. Canonical sampling
through velocity rescaling. J. Chem. Phys. 2007, 126, No. 014101.
(53) Essmann, U.; Perera, L.; Berkowitz, M. L.; Darden, T.; Lee, H.;
Pedersen, L. G. A smooth particle mesh Ewald method. J. Chem. Phys.
1995, 103, 8577−8593.
(54) Case, D. A. Calculations of NMR dipolar coupling strengths in
model peptides. J. Biomol. NMR 1999, 15, 95−102.
(55) Best, R. B.; Vendruscolo, M. Determination of Protein
Structures Consistent with NMR Order Parameters. J. Am. Chem.
Soc. 2004, 126, 8090−8091.

K https://doi.org/10.1021/acs.biochem.1c00749
Biochemistry XXXX, XXX, XXX−XXX

You might also like