You are on page 1of 8

RESEARCH ARTICLE | BIOPHYSICS AND COMPUTATIONAL BIOLOGY

CHEMISTRY
OPEN ACCESS

Ligand-­induced protein transition state stabilization switches


the binding pathway from conformational selection to
induced fit
Olof Stenströma,1 , Carl Diehla,1 , Kristofer Modiga , and Mikael Akkea,2

Edited by Lewis Kay, University of Toronto, Toronto, ON, Canada; received October 22, 2023; accepted February 29, 2024

Protein–ligand complex formation is fundamental to biological function. A central


question is whether proteins spontaneously adopt binding-­competent conformations Significance
to which ligands bind conformational selection (CS) or whether ligands induce the
binding-­competent conformation induced fit (IF). Here, we resolve the CS and IF Proteins change their
binding pathways by characterizing protein conformational dynamics over a wide conformations as part of their
range of ligand concentrations using NMR relaxation dispersion. We determined biological functions, for example,
the relative flux through the two pathways using a four-­state binding model that when binding another molecule
includes both CS and IF. Experiments conducted without ligand show that galectin-­3 (ligand). Two pathways are
exchanges between the ground-­state conformation and a high-­energy conformation conceivable: Either the ligand
similar to the ligand-­bound conformation, demonstrating that CS is a plausible path- binds preferentially to a rare
way. Near-­identical crystal structures of the apo and ligand-­bound states suggest that
conformation with higher affinity
the high-­energy conformation in solution corresponds to the apo crystal structure.
(“conformational selection”) or the
Stepwise additions of the ligand lactose induce progressive changes in the relaxation
dispersions that we fit collectively to the four-­state model, yielding all microscopic ligand first binds to the dominant
rate constants and binding affinities. The ligand affinity is higher for the bound-­like conformation and induces a
conformation than for the ground state, as expected for CS. Nonetheless, the IF path- transition to the high-­affinity
way contributes greater than 70% of the total flux even at low ligand concentrations. conformation (“induced fit”). We
The higher flux through the IF pathway is explained by considerably higher rates of determined all rate constants of
exchange between the two protein conformations in the ligand-­associated state. Thus, these two pathways for lactose
Downloaded from https://www.pnas.org by 102.188.205.138 on April 3, 2024 from IP address 102.188.205.138.

the ligand acts to decrease the activation barrier between protein conformations in a binding to the protein galectin-­3.
manner reciprocal to enzymatic transition-­state stabilization of reactions involving Even though galectin-­3 exists in
ligand transformation. a preestablished equilibrium of
protein dynamics | ligand-­binding kinetics | molecular recognition
low-­and high-­affinity
conformations, lactose binds
The mechanism of protein–ligand binding has been central to life science for more than primarily via the induced-­fit
60 y, ever since the seminal papers on enzyme flexibility and activity (1, 2) and allosteric pathway because lactose lowers
regulation (3, 4). At the heart of the matter lies the question whether proteins spontaneously the free-­energy barrier of
adopt conformations complementary to their ligands, to which the ligands subsequently galectin-­3’s conformational
bind, as envisioned by Monod–Wyman–Changeux (4), or whether ligands induce those change. As a result, the reaction
conformations upon binding, as initially proposed by Koshland (1). These mechanisms of flux switches from conformational
molecular recognition are denoted “conformational selection” (CS; sometimes referred to selection to induced fit already at
as “select fit”) and “induced fit” (IF), respectively (Fig. 1A). Koshland-­Nemethy-­Filmer (5) quite low ligand concentrations.
subsequently presented a modified view that allowed for CS as a possible mechanism for
ligand binding. IF has dominated the textbook view of protein–ligand interactions for a
long time. However, research over the past decades increasingly supports CS as an alternative
to the IF paradigm (6, 7). In particular, recent developments in NMR spectroscopy and
computational chemistry have provided evidence of short-­lived, weakly populated con- Author contributions: O.S., C.D., and M.A. designed
formers of the nonbound protein that resemble the ligand-­bound state (8–18). While these research; O.S. and C.D. performed research; K.M. and
M.A. contributed new reagents/analytic tools; O.S., C.D.,
results suggest that ligand binding might proceed via the CS pathway, they do not in K.M., and M.A. analyzed data; and O.S., C.D., K.M., and
themselves provide any direct evidence that this is the case. It is the net reaction flux, rather M.A. wrote the paper.

than the rates and equilibrium populations of individual reaction steps, that determines The authors declare no competing interest.

which pathway is the dominating one (18–22). This article is a PNAS Direct Submission.

NMR relaxation dispersion experiments and line-­shape analysis can probe conforma- Copyright © 2024 the Author(s). Published by PNAS.
This open access article is distributed under Creative
tional transitions that give rise to changes in the chemical shift of individual nuclei, which Commons Attribution License 4.0 (CC BY).
provides a powerful means of monitoring ligand-­binding kinetics under equilibrium 1
Present address: SAR AB, Medicon Village, SE-­223 81
conditions (8, 24–33). The results yield both an atomic-­resolution view of the binding Lund, Sweden.
process, as sensed by the protein, and the microscopic rate constants of the individual 2
To whom correspondence may be addressed. Email:
mikael.akke@bpc.lu.se.
steps of the reaction. Recent work has presented analytical expressions for chemical
This article contains supporting information online at
exchange processes involving an arbitrary number of states (34, 35). Previous studies https://www.pnas.org/lookup/suppl/doi:10.1073/pnas.​
have derived detailed expressions for each of the IF and CS pathways that make it possible 2317747121/-­/DCSupplemental.
to distinguish between these pathways based on the kinetic rate constants (36–42). Published March 25, 2024.

PNAS 2024 Vol. 121 No. 14 e2317747121 https://doi.org/10.1073/pnas.2317747121 1 of 8


However, it is important to recognize that the two pathways are A
not mutually exclusive but can both contribute simultaneously CS
to the overall reaction flux (41, 43).
Here, we use NMR relaxation dispersion experiments to meas- k12
ure the underlying protein dynamics and determine the relative 1 2
importance the CS and IF pathways for lactose binding to the k21
carbohydrate-­binding domain of human galectin-­3 (galectin-­3C).
Galectin-­3C binds saccharides and synthetic ligands in a shallow, k13 k31 k42 k24
solvent-­exposed groove on one side of its β-­sandwich structure
(23); see Fig. 1B. We performed 15N CPMG relaxation dispersion k34
measurements at variable degrees of ligand saturation and ana- 3 4
lyzed the data using a four-­state model that includes both the CS k43
and IF pathways (Fig. 1A). We found that ligand-­free (apo)
galectin-­3C transiently adopts an “open” conformation similar
to the ligand-­bound state, indicating that the protein is capable IF
of binding ligand via the CS pathway. Indeed, the determined
rate constants reveal that the ligand has a higher affinity for the
open conformation than for the closed one. Yet, the IF pathway
dominates across the entire range of ligand concentrations used B
here, corresponding to saturation levels between 4 and 76%. This
result is explained by significantly higher exchange rates between
states 3 and 4 than between 1 and 2, see Fig. 1A, which implies
that the ligand stabilizes the transition state of the protein con- R144
formational change. H158
W181 N160
Results and Discussion
R162
We investigated the kinetic ligand-­binding mechanisms in galec- N174
tin-­3C by recording 15N CPMG relaxation dispersion experiments
on the apo protein and seven additional samples with increasing E184 E165
Q187
amounts of added ligand; the resulting relaxation rate constants have
been deposited in the Biological Magnetic Resonance Data Bank
Downloaded from https://www.pnas.org by 102.188.205.138 on April 3, 2024 from IP address 102.188.205.138.

(44) under accession numbers 52174 and 52175. By numerically C


fitting the CPMG data globally to a four-­state model, using the
Bloch–McConnell equations (45), we included both the CS and IF
pathways of ligand binding. We analyzed the data in two steps: First,
we fitted the data for the apo state separately using a two-­state model
and then proceeded to fit the concentration dependent data, includ-
ing the results from the apo fit as fixed parameters. Fitting the four-­
state model is tractable in the present case, due in part to the
distinctive variation in dispersion profiles with ligand concentration,
and because one of the four equilibria can be isolated and fully deter-
mined from a two-­state fit of the apo state dynamics, which reduces
the roughness of the target function (46).
Fig. 1.   Ligand binding by galectin-­3C. (A) Schematic outline of the four-­state
model including the conformational-­selection (CS; states 1-­2-­4) and induced-­fit
Apo Galectin-­3C Transiently Adopts a Conformation Similar (IF; states 1-­3-­4) pathways for ligand binding. Free ligand is not shown. State 1
to the Ligand-­Bound Conformation. We first investigated the is the binding-­incompetent closed state; state 2 is the binding-­competent open
conformational dynamics of galectin-­3C in the absence of ligand. state; state 3 is the intermediate state in the IF pathway; and state 4 is the
final, highest-­affinity bound state. kon = k13 = k24, koff,3 = k31, and koff,4 = k42. The
We recorded 15N CPMG relaxation dispersion experiments on apo macroscopic apo state comprises states 1 and 2, while the macroscopic ligand-­
galectin-­3C at two static magnetic field strengths, 14.1 T and 18.8 bound state comprises states 3 and 4. (B) Three-­dimensional structure of lactose-­
T. Significant exchange contributions to the transverse relaxation bound galectin-­3C. The backbone trace is shown in gray ribbon representation
with ligand-­coordinating side chains highlighted in yellow. The lactose molecule
rates were observed for 13 residues in apo galectin-­3C (Fig. 2A and is shown in stick representation with carbon atoms in green and oxygen atoms in
SI Appendix, Fig. S1A and Table S1; BMRB entry 52174 (47)). red. (C) Comparison of the binding site in the apo and lactose-­bound galectin-­3C
Residue-­specific fits reveal exchange processes on two distinct crystal structures, highlighting the positions of oxygen atoms in water molecules
(pink spheres) bound to the apo protein and in lactose (red). Water oxygens
timescales. One set of six residues (V189, V213, D215, A216, primarily match those lactose oxygens that are less exposed to solvent. The
L219, and Y221) shows fast exchange dynamics between two protein backbone and side chains are shown in stick representation: blue, apo;
states, kex = (4.8 ± 0.8) × 103 s−1 and pMajor = 0.99 ± 0.01. These green, lactose bound. PDB entities: apo, 3zsl; lactose bound, 3zsj (23). Panels (B)
residues are all located in β-­strands 8 to 9 and the connecting and (C) were prepared using PyMOL (Schrödinger, LLC).
hairpin that form the edge of the β-­sheet on the opposite side of
the protein from the ligand-­binding site (Fig. 2I). Another set
of eight residues, viz. N174, K176, L177, R183, E184, E185, β-­strands 5 to 6 that form part of the binding site; I236 is located
Q187, and I236 shows slower exchange rates, k12 = 34 ± 4 s−1 and >5 Å from the binding site in β-­strand 11.
k21 = 735 ± 80 s−1, and relative populations, p1 = 0.96 ± 0.02 and The chemical shift difference, Δδ, between exchanging states
p2 = 0.04 ± 0.02. Except for I236, these residues are all located in can be extracted from the CPMG dispersions (48, 49), see

2 of 8 https://doi.org/10.1073/pnas.2317747121 pnas.org
A Apo B 4% I
I236

L177
K176

N174
C 8% D 11% R183
E184
E185 Q187
J
E184
L177
R183
Q187 N174

K176
E 21% F 27%
I236
E185

G 30% H 76% V155


L147 R144

K176

N174

R183 E184
Downloaded from https://www.pnas.org by 102.188.205.138 on April 3, 2024 from IP address 102.188.205.138.

Q187

Fig. 2.   Conformational exchange and ligand-­binding dynamics in galectin-­3C. (A–H) Representative 15N CPMG relaxation dispersion data for residue E184.
SI Appendix, Fig. S1 shows the corresponding data for all residues included in the fits. Error bars correspond to ±1 SD. (A) Relaxation dispersions for apo galectin-­3C,
with blue (red) symbols indicating data obtained at a static magnetic field strength of 14.1 T (18.8 T). (B–H) Relaxation dispersions obtained at 14.1 T for galectin-­3C
with varying levels of lactose saturation: (B) 4%, (C) 8 %, (D) 11 %, (E) 21 %, (F) 27 %, (G) 30 %, and (H) 75 %. (I) Residues showing conformational exchange in apo
galectin-­3C mapped onto the structure. Red indicates residues in the binding site showing slower exchange dynamics between states 1 (major population) and
2 (minor population). Blue indicates residues showing fast exchange dynamics unrelated to the exchange between states 1 and 2 (all located on the backside
of the protein, away from the binding site). Side chains involved in ligand coordination are shown in stick representation. (J) Chemical shift differences between
states 1 and 2, ΔδCPMG, in apo galectin-­3C determined from 15N CPMG relaxation dispersion experiments plotted versus chemical shift differences between
the apo and lactose-­bound forms, ΔδHSQC, determined either directly from HSQC spectra of each form (blue symbols; Pearson correlation coefficient: r = 0.96;
slope = 1.82) or from the chemical shifts obtained by fitting the four-­state model (red symbols; r = 0.99; slope = 1.47). A straight line with slope = 1 is drawn to
guide the eye. (K) Residues colored red show conformational exchange in all seven samples with lactose added. Panels (I) and (K) were prepared using PyMOL
(Schrödinger, LLC) and PDB entry 3zsj (23).

SI Appendix, Table S2. Notably, for the eight residues in the bind- Assuming that lactose-­bound galectin-­3C adopts the same con-
ing site, the chemical shifts of the high-­energy conformation of formation in solution as in the crystal, the highly similar chemical
apo galectin-­3C are similar to those of lactose-­bound galectin-­3C, shifts of the apo and lactose-­bound states suggest that the
as evidenced by comparing ΔδCPMG obtained from the CPMG high-­energy conformation of apo galectin-­3C in solution corre-
relaxation dispersion data and ΔδHSQC measured from HSQC sponds to the apo crystal structure. In the apo crystal structure,
spectra of the apo and lactose-­bound states (Fig. 2J); Pearson’s several water molecules are present in the ligand binding site with
correlation coefficient r = 0.96. In contrast, none of the exchanging their oxygen atoms occupying the same positions as oxygens in
residues located on the backside of the protein experience any the bound lactose. Furthermore, the ligand-­coordinating protein
significant chemical shift perturbation upon lactose binding; side chains have the same orientation in the apo and lactose-­bound
|ΔδHSQC| ≤ 0.025 ppm, further indicating that these residues are crystal structures (Fig. 1C). These observations explain the simi-
involved in a separate exchange process unrelated to ligand bind- larity in chemical shifts (Fig. 2J).
ing. Note that we do not expect a perfect correlation in Fig. 2J We conducted additional experiments to validate this result and
because minor conformational adjustments upon binding and the refute the possibility that there might be residual lactose present
presence of the lactose molecule in the binding site are expected in the apo sample that would result in exchange between a major
to give rise to additional chemical shift changes. The good agree- population of the apo state and a minor population of the
ment between ΔδCPMG and ΔδHSQC provides strong evidence that lactose-­bound state (see SI Appendix for details). Taken together,
apo galectin-­3C transiently samples conformations resembling our results demonstrate that apo galectin-­3C exchanges between
the lactose-­bound state. As shown previously, the crystal structures a “closed” (binding-­incompetent) ground state and an “open”
of apo and lactose-­bound galectin-­3C are virtually identical (23). (binding-­competent) high-­energy state, implying that ligand

PNAS 2024 Vol. 121 No. 14 e2317747121 https://doi.org/10.1073/pnas.2317747121 3 of 8


binding can occur via conformational selection. Similar results Table 1.   Rate constants describing the four-­state model
have been reported for several other proteins (8–11). The exchange Rate constant* Fitted value Estimated error
observed for apo galectin-­3C defines the equilibrium between
k12 (s−1) 34 4
state 1 (ground state) and state 2 (high-­energy state) of the
four-­state model, K2,1 = k12/k21 = p2/p1. These parameters were
−1
k21 (s ) 735 80
subsequently fixed when fitting the full four-­state model as kon (106 M−1 s−1) 172 37
described below. koff,3 (105 s−1) 5.5 1.9
koff,4 (103 s−1) 2.1 0.5
Mapping Binding Pathways by Relaxation Dispersion Measure­
ments at Variable Ligand Concentration. To resolve the CS and IF
3
k34 (10 s ) −1 27 5
binding pathways and determine the flux through each pathway, we 3
k43 (10 s ) −1 1.93 0.05
measured 15N CPMG relaxation dispersion on galectin-­3C at seven *koff,3, koff,4, k34, and k43 were calculated from the fitted parameters ρoff and ρclose; see Mate-
rials and Methods and SI Appendix for details.
different concentrations of lactose, covering saturation levels between
4% and 76%. Five of the eight residues that exhibit exchange in the
absence of lactose also display relaxation dispersions of sufficient total 8 × 7 dispersion curves, each containing 15 to 24 data points.
quality at all seven concentrations of lactose: N174, K176, R183, The best-­fit results yielded kon = (172 ± 37) × 106 s−1M−1, ρclose=
E184, and Q187. Three additional residues, R144, L147, and 0.43 ± 0.01, and ρoff = 266 ± 48. From these fitted parameter values,
V155, show exchange for all lactose concentrations and were also we then extracted the off-­rate constants, koff,3 = (5.5 ± 1.9) ×
included in the analysis (Fig. 2K). The relaxation dispersion data 105 s−1 and koff,4 = (2.1 ± 0.5) × 103 s−1, and the rates of confor-
vary significantly with the concentration of lactose (Fig. 2 B–H and mational exchange involving the macroscopic bound state, k43 =
SI Appendix, Fig. S2; BMRB entry 52175 (50)), as a consequence (1.93 ± 0.05) × 103 s−1 and k34 = (27 ± 5) × 103 s−1. Since k12 and
of the changing populations and the progressively faster on-­rate. k21 are known from the experiments conducted on apo galectin-­3C,
Fig. 1 defines the four-­state model with the CS and IF pathways described above, the four-­state model is thus completely described
connecting states 1-­2-­4 and 1-­3-­4, respectively. In fitting the (Table 1).
four-­state model, we include three global parameters in addition For comparison with the fitted value of kon, we estimated the
to residue-­specific chemical shifts and R2,0 values. The three global diffusion-­limited on-­rate constant for lactose binding using Eq. 5,
parameters are the on-­rate constant for ligand binding, kon = k13 = which yielded konD = 10 × 109 s−1M−1. Estimates that take the rel-
k24, the ratio of the rate constants for ligand release from states 3 ative size of the binding site into account (51, 52) are reduced by
and 4, ρoff = koff,3/koff,4, and the ratio of the rate constants for roughly a factor of 10 from this value. Thus, the experimental value
protein conformational change from open to closed in each path- of kon is significantly lower than konD, which is expected because
way, ρclose = k21/k43. All rate constants (and relative populations) the former includes not only the diffusion-­limited rate of encounter
Downloaded from https://www.pnas.org by 102.188.205.138 on April 3, 2024 from IP address 102.188.205.138.

can be calculated from these parameters and the dissociation con- but also the diffusive “search” by the ligand across the protein sur-
stant, Kd, measured by ITC, together with k12 and k21, determined face from the initial point of contact to the immediate neighbor-
from CPMG dispersion experiments on the apo state (see Materials hood of the binding site. This result is also in keeping with our
and Methods and SI Appendix for details). The assumption that recent work, showing that several hundred ligand–galectin-­3C
the on-­rate constants (k13 and k24) of the two pathways are equal encounters occur for each successful binding event (53).
is expected to hold in the present case for the following reasons.
First, the NMR experiments detect ligand-­induced chemical shift Ligand Affinity Is Higher for the Open State. The fitted rate
changes of residues located in the ligand-­binding site. Thus, the constants enable us to determine the affinity of each protein
on-­rate constant measured here includes processes such as the conformation (open or closed) for lactose, which is not available
diffusion of the ligand across the protein surface following encoun- from traditional binding assays, including isothermal titration
ters where the ligand contacts the protein at locations remote from calorimetry or fluorescence (54, 55). The rate constants establish
the binding site. Second, the open and closed conformations of that the lactose affinity is considerably higher for state 2 than for
galectin-­3C are very similar, as evidenced by the fact that chemical state 1, Ka,2/Ka,1 = koff,3/koff,4 = 270, in keeping with the notion
shift changes between the apo and lactose-­bound states (ΔδHSQC) that state 2 has an open, binding-­competent conformation,
are limited and observed only for residues in the binding site. In whereas state 1 is closed and binding-­incompetent, as required
the absence of large-­scale conformational changes, it is unlikely in the CS binding model. These results might give the (incorrect)
that the on-­rate constant will differ between the two conforma- impression that the CS pathway dominates—however, at this
tions, especially since lactose binding does not involve interactions level of interpretation, we have not yet considered the critical
between charged species. The assumption of equal on-­rate con- determinant, namely the flux through each pathway (19, 20).
stants is not expected to be generally valid for protein–ligand
systems. The Induced-­Fit Pathway Dominates Despite the Presence of
To fit the four-­state model to the CPMG dispersion data, we a Bound-­Like Conformation in the Apo State. Next, we calculated
initially performed grid searches on kon, ρoff, and ρclose, while fitting the fluxes through the CS and IF pathways as a function of lactose
the chemical shifts and R20 values, so as to characterize the χ2 concentration based on the determined rate constants; see Eqs. 6
space. Projections of the χ2 space onto two dimensions (i.e., pairs and 7, which show that the flux through a given pathway is the
of kon, ρoff, and ρclose; see SI Appendix, Fig. S3) clearly show that inverse of the sum of the “waiting times” at each intermediate
χ2 has a well-­defined minimum for 0.3 < ρclose < 0.4, whereas the state. The results reveal that the IF pathway dominates under all
model is less sensitive to ρoff, although there is a preference for conditions used in the present experiments (Fig. 3A). The flux
200 < ρoff < 600. Furthermore, we find that low χ2 values are through the IF pathway varies between 73% of the total flux at
obtained with 150 × 106 s−1M−1 < kon < 250 × 106 s−1M−1. We the lowest ligand concentration, [L]tot = 29 µM (4% saturation),
then performed free fits of all parameters, starting from values of to 99% at the highest ligand concentration, [L]tot = 0.95 mM
kon, ρoff, and ρclose corresponding to low χ2. The model parameters (76% saturation). By extrapolation, we find that the binding
were fitted simultaneously to all relaxation data, comprising in pathway switches from CS to IF when the ligand concentration

4 of 8 https://doi.org/10.1073/pnas.2317747121 pnas.org
A 100 The dominance of the IF binding pathway is explained by a
significantly higher rate of transition from closed to open confor-
mation in the ligand-­bound form, k34 = 27 × 103 s−1, than in the
apo form, k12 = 34 s−1, cf. Fig. 1. Importantly, this result indicates
50 that the ligand decreases the barrier between the two protein con-
formations (Fig. 4), which provides direct validation of the mech-
anism for ligand-­induced conformational change hypothesized by
Koshland–Nemethy–Filmer (5); similar results have previously
0 been reported for the case of coupled folding and binding (41).
We recognize that this phenomenon is reciprocal to enzymatic
transition-­state stabilization in reactions that change the ligand
B 80 conformation or chemical structure. Thus, the effect is fully
expected from a physicochemical perspective, although it is argu-
60 ably a much less commonly acknowledged concept than enzymatic
transition-­state stabilization.
Fig. 3 C and D show how the relative populations of states 1 to
40 4 vary with lactose concentration and degree of saturation. State 2
dominates over state 3 in the regime below a protein saturation level
20 of 42% (Fig. 3C), but again at least 97% of the total flux goes
through the IF pathway under these conditions. The macroscopic
lactose-­bound form is dominated by state 4, but a non-­negligible
fraction of ligand-­bound protein is in state 3, p3:p4 = 0.07:0.93. It
C is conceivable that the relatively weak affinity of lactose leads to the
detectable population of closed conformations in the macroscopic
lactose-­bound form, whereas it might be expected that high-­affinity
ligands drive the system further toward state 4 (56). Given that p3:p4
is non-­negligible, we reevaluated the chemical shift comparison in
Fig. 2J by plotting ΔδCPMG (=|δ1 – δ2|) against Δδ14 = |δ1 – δ4|,
instead of ΔδHSQC = |δapo – (p3δ3 + p4δ4)/(p3 + p4)|. The corrected
chemical shift differences are represented by the red symbols in
Fig. 2J, which show improved agreement between the two datasets
Downloaded from https://www.pnas.org by 102.188.205.138 on April 3, 2024 from IP address 102.188.205.138.

(r = 0.99), further bolstering the conclusion that apo galectin-­3C


D samples bound-­like conformations.

The Four-­State Model Resolves Conformation-­Specific Chemical


Shifts. The chemical shifts determined for each residue included
in the four-­state model are listed in SI Appendix, Table S3. For
all residues but one, the chemical shifts of states 2 (δ2) and 4 (δ4)
are more similar to one another than they are to either of δ1 or
δ3. These results are perfectly in line with the conformational
changes expected for the CS and IF pathways (cf. Fig. 1) but
are not presumed when constructing the four-­state model. We
note that for all residues but E184, the conformational change
Fig. 3.   Fluxes and populations as a function of the degree of saturation
and ligand concentration. (A and B) Fluxes through the CS (blue) and IF (red)
pathways, plotted versus the degree of saturation. (B) Close-­up view of the
flux at low degrees of saturation, showing the cross-­over between the CS and G
IF pathways. (C and D) Relative populations of states 1 to 4, plotted versus the
degree of saturation (C), or the free ligand concentration (D). Blue, state 1; red,
state 2; yellow, state 3; and magenta, state 4.

becomes greater than 2 µM, where the relative population of state 2


bound states is only 0.3% (Fig. 3B). This result reinforces the state 1
notion that the existence of exchange between closed (binding-­ 0
state 3 reaction coordinate
incompetent) and open (binding-­competent) conformations state 4
of the apo protein does not in and of itself imply that CS is
the dominant binding pathway (19, 20), even though the open
conformation has a higher affinity for ligand, as shown here.
Fig. 4.   Ligand-­induced transition-­state stabilization of the closed–open
Rather, the conformational exchange in the free state simply conformational transition. Schematic, qualitative view of the free energy
reflects the fact that the two conformations differ in free energy profile along the reaction coordinate of the closed–open conformational
by a few kBT and are separated by relatively moderate free energy transitions. The curves illustrate the transition between states 1 and 2 (red,
barriers that can be readily crossed at ambient temperature, as without ligand) and between states 3 and 4 (blue, with ligand present). The free
energies of the closed states (1 and 3) have been set to the same value (ΔG
expected for a protein that changes its conformation as part of = 0) for ease of comparison. The figure is not intended to give a quantitative
its biological function. representation of the relative free energies of the different states.

PNAS 2024 Vol. 121 No. 14 e2317747121 https://doi.org/10.1073/pnas.2317747121 5 of 8


progressively with increasing lactose concentration, but the
CS A dispersion profiles maintain a noticeable upward bend to higher
R2 values at low νcp, even for the highest [L] (Fig. 5A). By contrast,

R2 (1/s)
when the IF pathway dominates, R2 starts out at low values and
first increases progressively with increasing [L], while flattening
out at high [L] (Fig. 5B). The experimental data are effectively a
mixture of the dispersion profiles expected for pure IF and CS
binding pathways, in that the profiles progressively decrease in R2
and flatten out with increasing [L], without showing any upward
bend at low νcp (Fig. 5C). Thus, the dispersion profiles describing
IF B lactose binding to galectin-­3C indicate that exchange between
all 4 states must be taken into account, even though the flux
R2 (1/s)

through the CS pathway is negligible at higher [L]. This result


is explained by the fact that the exchange between states 1 and 2
(a key feature of the CS model) exist in conjunction with the IF
pathway. That is, in the case of lactose binding to galectin-­3C,
the four-­state exchange predominantly follows a linear exchange
mechanism: 2 ⇌ 1 ⇌ 3 ⇌ 4, where state 1 is dominant at low
ligand concentrations, whereas the step 2 ⇌ 4 does not contribute
4-state C appreciably to the overall flux (cf. Fig. 1A).
R2 (1/s)

Concluding Remarks. We have shown that protein NMR relaxation


dispersion experiments can resolve the CS and IF ligand-­binding
pathways, even in those cases where both contribute to binding. By
fitting a four-­state binding model comprising both the CS and IF
pathways, we determined the individual rate constants that provide
information on conformation-­specific ligand affinities, as well as
cp (Hz)
the relative flux through the two pathways. Our results reveal that
Fig. 5.   Relaxation dispersion profiles as a function of ligand concentration for galectin-­3C binds lactose almost exclusively via the IF pathway,
different binding mechanisms. (A) CS binding pathway. (B) IF binding pathway. even though the apo protein populates bound-­like conformations
(C) The four-­state binding model comprising both CS and IF. The individual
curves show relaxation dispersions for protein samples with different extents
with higher ligand affinity. The dominance of the IF pathway arises
Downloaded from https://www.pnas.org by 102.188.205.138 on April 3, 2024 from IP address 102.188.205.138.

of ligand saturation, indicated by the color scheme going from light blue for primarily from a faster rate of protein conformational change in
0% saturation to magenta for 90% saturation, in steps of 10%; in total 10 the lactose-­associated form than in the apo form. Thus, the ligand
curves. In panel (A), with increasing ligand saturation, the dispersion curves acts to lower the transition barrier for protein conformational
reach higher R2 values at νcp = 0 and show a sharper decline at higher νcp. In
panel (B), the dispersion profile is flat at 0% saturation, then first increases dynamics, a phenomenon reciprocal to enzymatic transition-­state
with increasing saturation, followed by a decrease to again reach a flat plateau. stabilization. Our results provide an enriched view of induced fit
In panel (C), the dispersion profiles resemble those in panel (B) but start out and conformational selection that extends current understanding of
with a significant dispersion at 0% saturation [identical to the case in panel
(A)]. The figure was generated by simulating relaxation dispersions for the
protein–ligand binding processes.
different binding mechanisms using the rate constants determined herein
for lactose binding to galectin-­3C. Materials and Methods
15
Protein Expression and Sample Preparation. N-­labeled galectin-­3C sam-
from state 3 to 4 results in a greater change in chemical shift
ples were expressed, purified, and prepared as described (54, 57). Following
than does binding of lactose to the open conformation (i.e., the an initial set of CPMG experiments on the original apo galectin-­3C sample, the
change from state 2 to 4). This result might also be expected since sample was filtered using a Vivaspin 2 MWCO 5,000 centrifugal concentrator to
the chemical shift is arguably more sensitive to local and distinct dilute by a factor of 104 any lactose possibly remaining in the original sample.
conformational changes than to low-­affinity binding of lactose Lactose titrations were carried out by titrating small aliquots of lactose (8.8 mM or
onto the closed conformation. Thus, the overall picture emerging 35 mM lactose dissolved in NMR buffer) into three different galectin-­3C samples,
from the chemical shifts extracted by fitting the four-­state model resulting in seven samples covering 7 to 75% saturation of the protein with ligand
makes intuitive sense. (see SI Appendix, Table S4 for detailed information on sample concentrations).
15
N CPMG relaxation dispersion datasets were acquired at each titration point.
Relaxation Dispersion Profiles Are Diagnostic of the Binding
Relaxation Dispersion Experiments. Relaxation-­compensated CPMG 15N
Mechanism. Fig. 5 compares relaxation dispersion curves for
relaxation dispersion experiments (58) were performed at static magnetic field
the full four-­state model with those expected for purely CS or
strengths of 14.1 T (apo and lactose-­containing samples) and 18.8 T (apo sample
IF binding mechanisms, using for each pathway the relevant only) and a temperature of 301 ± 0.1 K, which was calibrated using a methanol
exchange parameters resulting from the fit of the four-­state sample before each series of experiments. An equilibration period of 5 ms was
model to our experimental data. Clearly, the dispersion curves added prior to the CPMG refocusing pulse train to ensure that the initial magnet-
describing the four-­state model, which is a representation of the izations reflect the relative populations of the exchanging states (59). The experi-
underlying experimental data, are very different from the pure ments performed at 18.8 T utilized the phase cycle proposed by Yip and Zuiderweg
CS or IF scenarios. The simulated CPMG relaxation dispersion to suppress artifacts due to off-­resonance effects (60). The effective relaxation rate,
profiles reveal diagnostic signatures for distinguishing between the R2, at a given CPMG refocusing frequency was determined from 2 data points (61).
different reaction mechanisms, as has recently been demonstrated Relaxation dispersions were sampled using 18 to 24 refocusing frequencies, νcp.
for the IF and CS pathways by the Weikl and Griesinger groups The sign of the chemical shift difference between the exchanging conformations in
(42). For example, when the CS pathway dominates, R2 decreases the apo state was determined using the HSQC/HMQC approach (62). The relaxation

6 of 8 https://doi.org/10.1073/pnas.2317747121 pnas.org
data have been deposited in the Biological Magnetic Resonance Data Bank (BMRB) ligand off-­rate constants from states 3 and 4, i.e., ρoff = koff,3/koff,4, where koff,3 and
as entries 52174 (apo) (47) and 52175 (lactose, all concentrations) (50): DOI: koff,4 relate to the IF and CS pathway, respectively. ρclose describes the ratio of the
10.13018/BMR52174; DOI: 10.13018/BMR52175. rates going from the open conformation to the closed conformation along the
CS and IF pathways, i.e., ρclose = k21/k34 (Fig. 1). Furthermore, we assumed that
Data Analysis. NMR data were processed using NMRPipe (63). The processing
the on-­rate constant of lactose is the same for the CS and IF pathways, which is
protocol involved a solvent filter, cosine-­squared apodization function, zero filling reasonable given that the NMR relaxation experiment reports on chemical shift
to twice the number of points in both dimensions, and a polynomial baseline changes between states 1 and 3, or 2 and 4, induced by ligand binding to the
correction in the 1H dimension. Peak integration and error analysis was performed binding site or its proximity.
using PINT (64, 65). Errors in the extracted R2 values were estimated from the Prior to minimization of the target function, we performed extensive grid
S/N using error propagation. searches on the parameters ρoff, ranging from 30 to 1,000; ρclose, 0.001 to
The relaxation dispersion data of apo galectin-­3C were analyzed using CPMGfit 1,000; and kon, (8 to 300) × 106 M−1 s−1, while the chemical shift differences
v2.42, an in-­house Matlab program. The exchange parameters of the Carver-­Richards and R20 were optimized. The combination of parameters yielding the lowest
two-­state exchange model were fitted to the relaxation dispersion data (24, 66): χ2 was taken as input for the minimization algorithm. At each iteration, the full
relaxation dispersion curve was simulated for all residues, using the current
R2 + R20 + Rex (1 ∕ 𝜏), [1] values of the fitted model parameters in the Bloch–McConnell equations, and
χ2 was evaluated. Errors in the fitted parameters were estimated from Monte–
in which Carlo simulations using 500 synthetic datasets normally distributed by the
experimental SDs (68). The Matlab code for fitting the relaxation dispersion data
{ }
1 1 −1
[2]
( ( ) ( ))
Rex (1 ∕ 𝜏) = kex − cosh D + cosh 𝜂 + − D _ cosh 𝜂 − , is available on GitHub (https://github.com/Akke-­group/4 _ state_fits.git) (69).
2 𝜏
We estimated the diffusion-­limited on-­rate constant for lactose binding based
on the relationship (21):
1
{ )1∕2 }
𝜓 + 2Δ𝜔2 + 𝜓 2 + 𝜁 2 [3]
( ) (
D± = ±1+ ,
2
kon,D = 4𝜋 r DNA × 103 , [5]
𝜏
�� �1∕2 ��1∕2
±𝜓 + 𝜓2 +𝜁2

𝜂± = √ , [4] where r ≈ 20 Å is the distance between the centers of ligand and protein in the
2 bound complex, NA is Avogadro’s number, and D is the sum of the diffusion
constants for lactose, Dlac = 0.57 10−9 m2s−1 (70), and galectin-­3C, DP = 0.126
and ψ = kex2 – Δω2, ζ = –2Δωkex(1–2pminor), kex = k1 + k−1 is the sum of the 10−9 m2s−1 as calculated using HydroNMR (71).
forward and reverse rate constants, Δω = Δδ γ B0 is the chemical shift difference The fluxes through the CS pathway (FCS) and the IF pathway (FIF) were calcu-
between the exchanging conformations, where γ is the gyromagnetic ratio of the lated as (20)
)−1
nuclear spin and B0 is the static magnetic field strength, R20 is the average limiting
(
1 1
value of the relaxation rate constant for processes other than chemical exchange, FCS = [ ] + [ ][ ] , [6]
Downloaded from https://www.pnas.org by 102.188.205.138 on April 3, 2024 from IP address 102.188.205.138.

pminor is the population of the minor (less populated) conformational state, which k12 P1 k24 P2 L
is related to the major conformational state by pmajor = 1−pminor, and τ = 1/(2νcp) )−1
is the spacing between refocusing pulses in the CPMG train.
(
1 1
Relaxation dispersion data for apo galectin-­3C were initially modeled by FIF = [ ][ ] + [ ] , [7]
k13 P1 L k34 P3 L
residue-­specific fits of the Carver–Richards equations. The statistical significance
of each fit was assessed by also fitting the data to a constant R20 value (i.e., mod-
eling a flat relaxation profile, indicating the absence of exchange), and the F-­test where [P1] and [P2] are the protein concentrations of states 1 and 2, respectively, [L] is
was used to discriminate between models by rejecting the simpler model at the the free ligand concentration, [P3L] is the concentration of the protein–ligand complex
level P < 0.01 (67). In a subsequent step, residues with similar residue-­specific in state 3, and kAB is the rate constant going from state A to B. Thus, the flux through a
exchange parameters were fitted simultaneously to yield a single, global set of given pathway is the inverse of the sum of the waiting times at each intermediate state.
parameters kex and pminor. Errors in the fitted parameters were estimated from Data, Materials, and Software Availability. NMR relaxation rate constants
1,000 synthetic datasets created using Monte–Carlo simulations (68). data have been deposited in BMRB (52174 (47) and 52175 (50)). The Matlab
To fit the 15N CPMG relaxation dispersions of galectin-­3C with variable code for fitting the relaxation dispersion data is available on GitHub (https://
amounts of lactose present, we constructed a four-­state model based on the github.com/Akke-­group/4_state_fits.git) (69).
Bloch–McConnell equation (46), in which state 1 is the ligand-­free, closed ground
state; state 2 is the ligand-­free, open high-­energy state resembling the ligand-­ ACKNOWLEDGMENTS. We thank Kristine Steen Jensen for helpful discussions
bound state; state 3 is the ligand-­associated, closed state; and state 4 is the and an anonymous reviewer for valuable comments. High-­field (18.8 T) NMR
ligand-­bound, open state (Fig. 1). This model includes both the CS (states 1-­2-­4) data were acquired at the Swedish NMR Center at the University of Gothenburg.
and IF (states 1-­3-­4) pathways. The exchange parameters determined from the This work was funded by the Swedish Research Council (2018-­4995), the FLÄK
experiments on apo galectin-­3C, i.e., k12 and k21, were fixed when fitting the four-­ Research School for Pharmaceutical Sciences at Lund University, the Knut and
state model. δ1 was fixed to the value measured in the 15N HSQC spectrum of apo Alice Wallenberg Foundation (2013.0022), and the European Union (ERC,
galectin-­3C (given that p2 ≪ p1, the contribution from δ2 on the apo chemical DynaPLIX, SyG-­2022 101071843). Views and opinions expressed are however
shift is insignificant). δ2 and δ3 were fitted while δ4 was calculated from p3, p4, those of the authors only and do not necessarily reflect those of the European
and the chemical shift measured in the 15N HSQC spectrum of lactose-­saturated Union or the European Research Council. Neither the European Union nor the
galectin-­3C: δlac = p3δ3 + p4δ4. R20 was fitted individually for each residue, but granting authority can be held responsible for them.
globally over all lactose concentrations.
We introduced two parameters ρoff and ρclose describing the ratios of pairs Author affiliations: aDivision of Biophysical Chemistry, Center for Molecular Protein
of corresponding rates along the two pathways. ρoff describes the ratio of the Science, Department of Chemistry, Lund University, SE-­221 00 Lund, Sweden

1. D. E. Koshland, Application of the theory of enzyme specificity to protein synthesis. Proc. Natl. Acad. 3. J. Monod, J. P. Changeux, F. Jacob, Allosteric proteins and cellular control systems. J. Mol. Biol. 6,
Sci. U.S.A. 44, 98–104 (1958). 306–329 (1963).
2. D. E. Koshland, Enzyme flexibility and enzyme action. J. Cell Comp. Physiol. 54, 245–258 4. J. Monod, J. Wyman, J.-­P. Changeux, On the nature of allosteric transitions: A plausible model. J.
(1959). Mol. Biol. 12, 88–118 (1965).

PNAS 2024 Vol. 121 No. 14 e2317747121 https://doi.org/10.1073/pnas.2317747121 7 of 8


5. D. E. Koshland, G. Nemethy, D. Filmer, Comparison of experimental binding data and theoretical 39. S. Gianni, J. Dogan, P. Jemth, Distinguishing induced fit from conformational selection. Biophys.
models in proteins containing subunits. Biochemistry 5, 365–385 (1966). Chem. 189, 33–39 (2014).
6. J. P. Changeux, S. Edelstein, Conformational selection or induced fit? 50 years of debate resolved. 40. T. R. Weikl, D. D. Boehr, Conformational selection and induced changes along the catalytic cycle of
F1000 Biol. Rep. 3, 19 (2011). Escherichia coli dihydrofolate reductase. Proteins 80, 2369–2383 (2012).
7. D. D. Boehr, R. Nussinov, P. E. Wright, The role of dynamic conformational ensembles in 41. K. G. Daniels, Y. Suo, T. G. Oas, Conformational kinetics reveals affinities of protein conformational
biomolecular recognition. Nat. Chem. Biol. 5, 789–796 (2009). states. Proc. Natl. Acad. Sci. U.S.A. 112, 9352–9357 (2015).
8. A. Malmendal, J. Evenäs, S. Forsén, M. Akke, Structural dynamics in the C-­terminal domain of 42. K. S. Chakrabarti et al., A litmus test for classifying recognition mechanisms of transiently binding
calmodulin at low calcium levels. J. Mol. Biol. 293, 883–899 (1999). proteins. Nat. Commun. 13, 3792 (2022).
9. E. Z. Eisenmesser et al., Intrinsic dynamics of an enzyme underlies catalysis. Nature 438, 117–121 43. A. G. Palmer, Chemical exchange in biomacromolecules: Past, present, and future. J. Magn. Reson.
(2005). 241, 3–17 (2014).
10. D. D. Boehr, D. McElheny, H. J. Dyson, P. E. Wright, The dynamic energy landscape of dihydrofolate 44. J. C. Hoch et al., Biological magnetic resonance data bank. Nucleic Acids Res. 51, D368–D376 (2023).
reductase catalysis. Science 313, 1638–1642 (2006). 45. H. M. McConnell, Reaction rates by nuclear magnetic resonance. J. Chem. Phys. 28, 430–431
11. H. Beach, R. Cole, M. L. Gill, J. P. Loria, Conservation of us-­ms enzyme motions in the apo-­and (1958).
substrate-­mimicked state. J. Am. Chem. Soc. 127, 9167–9176 (2005). 46. P. Li, I. R. S. Martins, M. K. Rosen, The feasibility of parameterizing four-­state equilibria using
12. K. Okazaki, S. Takada, Dynamic energy landscape view of coupled binding and protein relaxation dispersion measurements. J. Biomol. NMR 51, 57–70 (2011).
conformational change: Induced-­fit versus population-­shift mechanisms. Proc. Natl. Acad. Sci. U.S.A. 47. O. Stenström, C. Diehl, K. Modig, M. Akke, Ligand-­induced transition state stabilization of protein
105, 11182–11187 (2008). conformational change switches the binding pathway from conformational selection to induced
13. O. F. Lange et al., Recognition dynamics up to microseconds revealed from an RDC-­derived fit -­APO. Biological Magnetic Resonance Data Bank. https://doi.org/doi:10.13018/BMR52174.
ubiquitin ensemble in solution. Science 320, 1471–1475 (2008). Deposited 15 October 2023.
14. S.-­R. Tzeng, C. G. Kalodimos, Dynamic activation of an allosteric regulatory protein. Nature 462, 48. E. L. Kovrigin, J. G. Kempf, M. J. Grey, J. P. Loria, Faithful estimation of dynamics parameters from
368–372 (2009). CPMG relaxation dispersion measurements. J. Magn. Reson. 179, 178–189 (2006).
15. N. J. Anthis, M. Doucleff, G. M. Clore, Transient, sparsely populated compact states of apo and 49. A. G. Palmer, C. D. Kroenke, J. P. Loria, Nuclear magnetic resonance methods for quantifying
calcium-­loaded calmodulin probed by paramagnetic relaxation enhancement: Interplay of microsecond-­to-­millisecond motions in biological macromolecules. Methods Enzymol. 339,
conformational selection and induced fit. J. Am. Chem. Soc. 133, 18966–18974 (2011). 204–238 (2001).
16. L. Ye, N. Van Eps, M. Zimmer, O. P. Ernst, R. Scott Prosser, Activation of the A2A adenosine G-­protein-­ 50. O. Stenström, C. Diehl, K. Modig, M. Akke, Ligand-­induced transition state stabilization of protein
coupled receptor by conformational selection. Nature 533, 265–268 (2016). conformational change switches the binding pathway from conformational selection to induced fit
17. V. Venditti, G. M. Clore, Conformational selection and substrate binding regulate the monomer/ -­LACTOSE. Biological Magnetic Resonance Data Bank. https://doi.org/doi:10.13018/BMR52175.
dimer equilibrium of the C-­terminal domain of Escherichia coli enzyme I. J. Biol. Chem. 287, Deposited 15 October 2023.
26989–26998 (2012). 51. G. Schreiber, G. Haran, H. X. Zhou, Fundamental aspects of protein -­Protein association kinetics.
18. A. Sekhar et al., Conserved conformational selection mechanism of Hsp70 chaperone-­substrate Chem. Rev. 109, 839–860 (2009).
interactions. eLife 7, e32764 (2018). 52. O. G. Berg, Orientation constraints in diffusion-­limited macromolecular association. The role of
19. V. Z. Miloushev et al., Dynamic properties of a type II cadherin adhesive domain: Implications for the surface diffusion as a rate-­enhancing mechanism. Biophys. J. 47, 1–14 (1985).
mechanism of strand-­swapping of classical cadherins. Structure 16, 1195–1205 (2008). 53. S. Wernersson, S. Birgersson, M. Akke, Cosolvent dimethyl sulfoxide influences protein-­ligand
20. G. G. Hammes, Y. C. Chang, T. G. Oas, Conformational selection or induced fit: A flux description of binding kinetics via solvent viscosity effects: Revealing the success rate of complex formation
reaction mechanism. Proc. Natl. Acad. Sci. U.S.A. 106, 13737–13741 (2009). following diffusive protein-­ligand encounter. Biochemistry 62, 44–52 (2023).
21. K. S. Chakrabarti et al., Conformational selection in a protein-­protein interaction revealed by 54. C. Diehl et al., Protein flexibility and conformational entropy in ligand design targeting the
dynamic pathway analysis. Cell Rep. 14, 32–42 (2016). carbohydrate recognition domain of galectin-­3. J. Am. Chem. Soc. 132, 14577–14589 (2010).
22. T. R. Weikl, F. Paul, Conformational selection in protein binding and function. Protein Sci. 23, 55. M. L. Verteramo et al., Interplay between conformational entropy and solvation entropy in protein-­
1508–1518 (2014). ligand binding. J. Am. Chem. Soc. 141, 2012–2026 (2019).
23. K. Saraboji et al., The carbohydrate-­binding site in galectin-­3 is preorganized to recognize a sugar-­ 56. M. Ubbink, The courtship of proteins: Understanding the encounter complex. FEBS Lett. 583,
like framework of oxygens: Ultra-­high resolution structures and water dynamics. Biochemistry 51, 1060–1066 (2009).
296–306 (2012). 57. C. Diehl, S. Genheden, K. Modig, U. Ryde, M. Akke, Conformational entropy changes upon lactose
24. D. G. Davis, M. E. Perlman, R. E. London, Direct measurements of the dissociation-­rate constant binding to the carbohydrate recognition domain of galectin-­3. J. Biomol. NMR 45, 157–169 (2009).
Downloaded from https://www.pnas.org by 102.188.205.138 on April 3, 2024 from IP address 102.188.205.138.

for inhibitor-­enzyme complexes via the T1r and T2 (CPMG) methods. J. Magn. Reson. Ser. B 104, 58. J. P. Loria, M. Rance, A. G. Palmer, A relaxation-­compensated Carr-­Purcell-­Meiboom-­Gill sequence for
266–275 (1994). characterizing chemical exchange by NMR spectroscopy. J. Am. Chem. Soc. 121, 2331–2332 (1999).
25. R. Schneider et al., Visualizing the molecular recognition trajectory of an intrinsically disordered 59. D. F. Hansen, P. Vallurupalli, L. E. Kay, An improved N-­15 relaxation dispersion experiment for the
protein using multinuclear relaxation dispersion NMR. J. Am. Chem. Soc. 137, 1220–1229 (2015). measurement of millisecond time-­scale dynamics in proteins. J. Phys. Chem. B 112, 5898–5904
26. O. Stenström, C. Diehl, K. Modig, U. J. Nilsson, M. Akke, Mapping the energy landscape of (2008).
protein–ligand binding via linear free energy relationships determined by protein NMR relaxation 60. G. N. B. Yip, E. R. P. Zuiderweg, A phase cycle scheme that significantly suppresses offset-­dependent
dispersion. RSC Chem. Biol. 2, 259–265 (2021). artifacts in the R-­2-­CPMG N-­15 relaxation experiment. J. Magn. Reson. 171, 25–36 (2004).
27. V. A. Feher, E. P. Baldwin, F. W. Dahlquist, Access of ligands to cavities within the core of a protein is 61. F. A. A. Mulder, N. R. Skrynnikov, B. Hon, F. W. Dahlquist, L. E. Kay, Measurement of slow
rapid. Nat. Struct. Biol. 3, 516–521 (1996). (microseconds-­milliseconds) time scale dynamics in protein side chains by 15N relaxation
28. D. Zeng, V. S. Bhatt, Q. Shen, J. H. Cho, Kinetic insights into the binding between the nSH3 domain dispersion NMR spectroscopy: Application to Asn and Gln residues in a cavity mutant of T4
of CrkII and proline-­rich motifs in cAbl. Biophys. J.111, 1843–1853 (2016). lysozyme. J. Am. Chem. Soc. 123, 967–975 (2001).
29. E. Z. Eisenmesser, D. A. Bosco, M. Akke, D. Kern, Enzyme dynamics during catalysis. Science 295, 62. N. R. Skrynnikov, F. W. Dahlquist, L. E. Kay, Reconstructing NMR spectra of “invisible” excited states
1520–1523 (2002). using HSQC and HMQC experiments. J. Am. Chem. Soc. 124, 12352–12360 (2002).
30. K. Sugase, J. C. Lansing, H. J. Dyson, P. E. Wright, Tailoring relaxation dispersion experiments for 63. F. Delaglio et al., NMRPipe: A multidimensional spectral processing system based on UNIX pipes.
fast-­associating protein complexes. J. Am. Chem. Soc. 129, 13406–13407 (2007). J. Biomol. NMR 6, 277–293 (1995).
31. D. Tolkatchev, P. Xu, F. Ni, Probing the kinetic landscape of transient peptide-­protein interactions by 64. A. Ahlner, M. Carlsson, B. H. Jonsson, P. Lundström, PINT: A software for integration of peak volumes
use of peptide 15N NMR relaxation dispersion spectroscopy: Binding of an antithrombin peptide to and extraction of relaxation rates. J. Biomol. NMR 56, 191–202 (2013).
human prothrombin. J. Am. Chem. Soc. 125, 12432–12442 (2003). 65. M. Niklasson et al., Comprehensive analysis of NMR data using advanced line shape fitting.
32. S. De et al., Complete thermodynamic and kinetic characterization of the isomer-­specific interaction J. Biomol. NMR 69, 93–99 (2017).
between Pin1-­WW domain and the amyloid precursor protein cytoplasmic tail phosphorylated at 66. J. P. Carver, R. E. Richards, A general two-­site solution for the chemical exchange produced
Thr668. Biochemistry 51, 8583–8596 (2012). dependence of T2 upon the Carr-­Purcell pulse separation. J. Magn. Reson. 6, 89–105 (1972).
33. A. Furukawa, T. Konuma, S. Yanaka, K. Sugase, Quantitative analysis of protein–ligand interactions by 67. J. L. Devore, Probability and Statistics for Engineering and the Sciences (Brooks/Cole Publishing
NMR. Prog. Nucl. Magn. Reson. Spectrosc. 96, 47–57 (2016). Company, ed. 5, 1999).
34. H. Koss, M. Rance, A. G. Palmer, General expressions for Carr-­Purcell-­Meiboom-­Gill relaxation 68. W. H. Press, B. P. Flannery, S. A. Teukolsky, W. T. Vetterling, Numerical Recipes. The Art of Scientific
dispersion for N-­site chemical exchange. Biochemistry 57, 4753–4763 (2018). Computing (Cambridge University Press, 1986).
35. H. Koss, M. Rance, A. G. Palmer, Algebraic expressions for Carr-­Purcell-­Meiboom-­Gill relaxation 69. O. Stenström, K. Modig, M. Akke, Matlab scripts to fit ligand-­concentration dependent CPMG
dispersion for N-­site chemical exchange. J. Magn. Reson. 321, 106846 (2020). relaxation dispersion data to a 4-­state exchange binding model. GitHub. https://github.com/Akke-­
36. A. D. Vogt, E. Di Cera, Conformational selection or induced fit? A critical appraisal of the kinetic group/4_state_fits. Accessed 20 February 2024.
mechanism. Biochemistry 51, 5894–5902 (2012). 70. A. C. F. Ribeiro et al., Binary mutual diffusion coefficients of aqueous solutions of sucrose, lactose,
37. A. D. Vogt, E. Di Cera, Conformational selection is a dominant mechanism of ligand binding. glucose, and fructose in the temperature range from (298.15 to 328.15) K. J. Chem. Eng. Data 51,
Biochemistry 52, 5723–5729 (2013). 1836–1840 (2006).
38. F. Paul, T. R. Weikl, How to distinguish conformational selection and induced fit based on chemical 71. P. Bernado, J. Garcia de la Torre, M. Pons, Interpretation of 15N NMR relaxation data of globular
relaxation rates. PLoS Comput. Biol. 12, e1005067 (2016). proteins using hydrodynamic calculations with HYDRONMR. J. Biomol. NMR 23, 139–150 (2002).

8 of 8 https://doi.org/10.1073/pnas.2317747121 pnas.org

You might also like