Introduction

Structural plasticity in proteins is intimately linked to their function. For example, the activity of many signal transducing enzymes is associated with large conformational changes between active and inactive states1,2. Moreover, many proteins undergo significant conformational changes on interaction with target proteins or small ligands3,4. In analogy to the energy landscape model of protein folding5, transitions between native conformational states in folded proteins are thought to be achieved through multiple pathways6,7. Qualitative descriptions of conformational changes can be inferred from analysis of the stable ground states. Hinge and shear motions, and combinations thereof, have been proposed to facilitate conformational changes8. What is not yet clear is the molecular mechanism, including information on transition states and other transient intermediates, of conformational change in the native states of proteins. It is difficult to probe intermediates and transition states involved in conformational changes, as these states are short lived and not directly accessible by spectroscopy. Here, we describe experiments that allow us to decipher the molecular mechanism of a large-scale conformational change in adenylate kinase (AK).

AK catalyses the reversible phosphoryl transfer reaction Mg2+ATP+AMP↔Mg2+ADP+ADP and has the primary role of maintaining the energy balance in cells. The importance of AK is reflected in its ubiquitous presence in many different organisms and tissues9. AK isolated from E. coli (AKe) uses a random bi–bi mechanism during catalysis10,11 and the phosphoryl transfer chemistry likely follows an associative mechanism12. A simplified nucleotide-binding mechanism for AK is shown in Figure 1.

Figure 1: Kinetic and structural models for substrate binding by AK.
figure 1

(a) Nucleotide binding mechanism for AK. E corresponds to enzyme, S to substrate, ESO to a transient substrate-bound encounter complex, ESC to the active conformation of AK. Superscripts 'O' and 'C' refer to AK in open (inactive) and closed (active) conformations. The rate constants for association and dissociation of ESO are given by kon and koff, whereas the rate constants for conformational change are given by kclose and kopen. (b, c) Structures of AKe during the extreme stages of catalysis. ATPlid (residues 113–176) and AMPlid (residues 28–72) are coloured gold and the INSERT segment of the ATPlid (residues 122–15943) is coloured blue. (b) Substrate-free open AKe (4AKE.pdb). Residues isoleucine 116 and leucine168, which are mutated into glycine in this study, are highlighted with ball and stick representations, and the location of alpha helices, α6 and α7, is indicated. (c) Closed AKe with bound Ap5A (1AKE.pdb). All molecular graphic images in this report were prepared with MolMol44.

The structure of AKe has been solved in substrate-free (open)13 and inhibitor P1,P5–Di(adenosine-5′)pentaphosphate (Ap5A14)-bound (closed) conformations15 (Fig. 1). NMR experiments conducted in our laboratory have shown that the closed conformation represents the catalytically active state in which phosphoryl transfer occurs16. AKe is a modular enzyme and consists of a CORE subdomain and ATP- and AMP-binding subdomains (Fig. 1). Both ATP and AMP subdomains (ATPlid and AMPlid) undergo significant conformational changes in response to substrate binding. The large conformational change has at least two functional origins: first, the active site must be dehydrated to avoid non-productive hydrolysis of ATP and, second, the substrates are aligned optimally for the phosphoryl-transfer chemistry. The local thermodynamic stability of the nucleotide-binding subdomains is significantly lower compared with the CORE, and the subdomains fold and unfold in a non-cooperative manner17. From a functional perspective, this feature allows the rearrangement in backbone hydrogen bonding patterns (observed during the open-to-closed transition) without provoking global unfolding of the entire protein. It was recently reported that, at temperatures above 35 °C, the INSERT segment in ATPlid (Fig. 1) selectively unfolded, resulting in reduced Ap5A-binding affinity18. The state with an unfolded INSERT segment was denoted as a 'binding incompetent state' and appears to be an intermediate on the global unfolding pathway. Selective unfolding of INSERT agrees with our previous observation that ATPlid can be selectively unfolded without perturbing the structure of other parts of the enzyme17.

The structural plasticity of AKe is strongly coupled to its catalytic efficiency, as reopening of substrate binding domains with bound nucleotides is rate limiting for catalysis19. This observation is consistent with the finding that the rate-limiting step involves dissociation of bound nucleotides10. Notably, the conformational changes of ATPlid and AMPlid occur even in the absence of substrates20. Addition of ATP and AMP gradually shifts the conformational equilibrium towards the closed state16. Apparently, the conformational dynamics that is relevant for substrate binding is an intrinsic property of AKe. This feature of enzyme function has also been observed for dihydrofolate reductase21, cyclophilin A22 and RNase A23.

There exist two contradicting models for the open/closed conformational change of the nucleotide-binding subdomains in AK: (i) 'induced fit' and (ii) 'cracking' models. On the basis of an analysis of crystallographic structures, the induced fit model has been proposed for closure of both ATPlid and AMPlid. In this model, segments in between hinge regions are spatially translated as rigid bodies24. Closure of ATPlid is accomplished by rotation of the INSERT segment around helices α6 and α7, which in turn rotate around the CORE (Fig. 1). In the 'cracking' model, significant strain energy is built up for residues in α6 and α7 during the reaction, and, at the transition state, this strain is released by a local unfolding event. At the downhill side of the activation energy barrier, the protein refolds and the thermodynamic minimum corresponding to the closed state is reached25,26. The cracking model was inferred from coarse-grained modelling and the principle of minimal frustration26. There are no experimental studies verifying which mechanism is used by AK. We previously identified a hydrophobic cluster in ATPlid that consists of isoleucine (Ile) 116, valine (Val) 117, Val 164 and leucine (Leu) 168. Simultaneous mutation of these positions to glycine selectively unfolded the entire ATPlid without perturbing the structures of the CORE and AMPlid subdomains17. Surprisingly, this variant is capable of binding the inhibitor Ap5A (unpublished results). The hydrophobic cluster constituted of these four residues is conserved in the AK family, as evident from a published sequence alignment of 250 AK sequences27.

On the basis of our observations on the quadruple mutant, we have, in this study, designed single (Ile116Gly, 'M1') and double (Ile116Gly and Leu168Gly, 'M2') mutated variants of AKe in order to probe the reaction mechanism of ATPlid closure. Our data show that segments in ATPlid unfold and subsequently refold during the open-to-closed conformational transition, and show that folding and function are closely coupled events in AKe.

Results

Structure of mutated AKe in apo states

The structures of M1 and M2 AKe variants were assessed with far-ultraviolet circular dichroism (CD) spectroscopy (Supplementary Fig. S1). Wild-type (WT) AKe displays two distinct minima in the CD signal at 208 and 222 nm, in agreement with previous results28. Single and double mutations decrease the helical content of AKe, as inferred from a reduction in the intensity of CD signals at 208 and 220 nm. As discussed below, this effect is dependent on selective destabilization/unfolding of α6 and α7. A qualitative analysis of the CD spectra using CDNN software (copyright Dr Gerald Böhm) provided deconvoluted helical contents of 64, 54 and 39 (±5) % for WT, M1 and M2, respectively. The reduction of helical content in M2 compared with WT AKe is consistent with a complete loss of the helical structure of α6 and the section of α7 belonging to the ATPlid. Taking the helical contents of WT AKe and M2 as benchmarks for folded and unfolded α6 and α7, M1 is placed in between with a fraction of folded α6, α7 of 60%.

1H-15N Heteronuclear Single-Quantum Coherence (HSQC) spectra of the two mutants superimposed on the WT are shown in Figure 2a,b. A visual inspection shows that M1 resembles the WT enzyme with many overlapping peaks distributed over the whole spectrum. In contrast, M2 shows marked spectral differences, compared with the WT enzyme.

Figure 2: NMR spectra of substrate-free WT AKe, M1 and M2 mutants.
figure 2

In (a) and (b), WT AKe is shown with blue contouring and the two mutants are contoured red. (a) Superimposition of 1H-15N HSQC spectra of WT AKe and M1 acquired at 25 °C. (b) Superimposition of 1H-15N HSQC spectra of WT AKe and M2 acquired at 25 °C. Selected resonances in the INSERT segment of the ATPlid that are missing in the M2 mutant are indicated. (c) Superimposition of 1H-15N HSQC spectra of M2 acquired at 25 °C (red contours) and 15 °C (blue contours). Selected resonances (same set as in b) corresponding to residues in the INSERT segment in ATPlid are indicated.

Combined absolute differences in 1H and 15N chemical shifts between the M1 mutant and WT AKe are shown in Figure 3a, and these results are plotted on the open AKe structure in Figure 3b. Significant chemical shift differences are found in α6, α7, and these are propagated to a large part of the CORE. In light of the results from CD spectroscopy and from our previous analysis of a quadruple AKe mutant17, the chemical shift perturbations in α6 and α7 depend on a local unfolding event. The differences observed in the CORE are not dependent on unfolding, as we have shown that the quadruple mutant does not affect the topology of this domain17. The region of the CORE that shows shift changes (Fig. 3b,c) is part of a dense interaction network, and the observed changes may be reflecting minute changes in the packing of hydrophobic side chains. Interestingly, residues in the INSERT segment show no significant chemical shift perturbations in M1 (Fig. 3d), showing that this part of the molecule is structurally unaffected by the mutation. This result suggests that ATPlid is composed of two cooperatively folding units: the INSERT and α6, α7 parts.

Figure 3: Propagation of structural perturbations in the M1 mutant.
figure 3

Perturbations resulting from the mutation were quantified on the basis of combined absolute 15N and 1H chemical shift differences between WT and mutant, calculated according to δω=0.2 × |Δ15N|+|Δ1H| (p.p.m.)16. As a result of the point mutation, structural rearrangements are propagated throughout the entire ATPlid, with the exception of residues in the INSERT segment (indicated with a dashed oval). (a) Chemical shift differences for non-mutated residues plotted against the primary sequence; threshold values used in (b) are indicated with red and orange lines. (b) Chemical shift perturbations displayed on the open AKe structure (4AKE.pdb); perturbations are colour coded as follows: red δω>0.4 p.p.m., orange 0.4>δω>0.2 p.p.m.; residues that are unassigned in either WT or M1 and proline residues are coloured blue. The mutation site and residues F19, E143 and L178 are shown as dark blue ball and sticks. (c, d) Expansions of 1H-15N HSQC spectra of WT (blue) and M1 (red) AKe. (c) Chemical shift perturbations of L178 and F19 (both 0.27 p.p.m., calculated as stated above); both these residues are accordingly coloured orange in (b). (d) The chemical shift of E143 is identical in the mutant and WT spectra, consistent with identical structures in the INSERT segment.

Partial assignments were obtained for the M2 mutant. The resonances corresponding to the ATPlid could not be assigned because of unfolding of this subdomain. On the basis of spectral similarities with published NMR spectra of a quadruple mutant17, we conclude that M2 has an unfolded ATPlid subdomain (in agreement with the CD results). On the other hand, many residues in the CORE and AMPlid could be assigned. In a spectrum recorded at 15 °C, resonances corresponding to the INSERT segment appeared at resonance frequencies very close to that of WT at the same temperature (Fig. 2c). This result implies that the INSERT retains the WT structure at 15 °C and that the absence of these resonances at 25 °C is due to conformational fluctuations on the μs–ms time scale. Notably, resonances corresponding to α6 and α7 could not be observed at 15 °C. The decoupling between folding of α6, α7 and the INSERT is thus a conserved feature of both M1 and M2 variants.

Protein stability was studied by measuring thermal stability, which provides melting points (Tm) and enthalpies of unfolding (ΔHvH) at Tm. Unfolding curves for the three AKe variants (Fig. 4) were fitted to a two-state unfolding transition according to Equation (3). Both mutated variants show a modest decrease in Tm compared with the WT (Supplementary Table S1); however, the ΔHvH is significantly lower for both mutants. The large enthalpic difference is consistent with an unfolded ATPlid for the mutants at Tm. The small effect on Tm in light of the large enthalpic differences between WT and mutants is explained with the thermal transition being dominated by the CORE subdomain.

Figure 4: Thermal denaturation of AKe variants.
figure 4

Unfolding was monitored with the CD signal at 220 nm for WT (filled circles), M1 (filled triangles) and M2 (open triangles). Solid lines represent nonlinear fits to Equation (3). Fitted parameters are summarized in Supplementary Table S1.

Local unfolding promotes ATP-binding affinity

ATP-binding affinities and enthalpies between AKe variants and ATP were determined using isothermal titration calorimetry (ITC; Fig. 5 and Table 1). All experiments were performed in duplicate and the reported parameters are average values. Surprisingly, the binding affinities of the two mutants, M1 and M2, towards ATP are increased markedly compared with WT AKe. The observed dissociation constant (Kappd) for WT is 51±13 μM (pH 7, 25 °C), with a calorimetric enthalpy (ΔHITC) of −2.2±0.4 kJ mol−1. Binding between ATP and AKe has previously been quantified with other biophysical techniques16,29,30. These independent studies provided Kappd values in the range of 35–53 μM in good agreement with our ITC data. As ATP binding is associated with a low enthalpy change and baseline distortions, we performed a control experiment using Ap5A (Supplementary Fig. S2). Ap5A binds AKe with a Kappd of 0.14±0.5 μM and a ΔHITC of −19±5 kJ mol−1, in good agreement with published results18. Kappd values quantified for M1 and M2 are 0.78±0.2 and 0.98±0.3 μM, respectively. All AK variants bind to ATP with a one-to-one stoichiometry as evident by fitted n values (Table 1), in agreement with previous NMR studies that showed one-to-one binding between WT AKe and ATP16. Notably, binding of ATP to M1and M2 is associated with significantly increased ΔHITC, −51±2 and −110±10 kJ mol−1, respectively. This difference is explained with a coupled folding and binding event in the case of M1 and M2, in which ΔHITC is a cumulative sum of ATPlid folding (ΔHfol) and ATP-binding enthalpies (ΔHbind), according to Equation (1) (fu refers to the degree of ATPlid unfolding).

Figure 5: Increased binding affinities towards ATP for mutated enzymes relative to WT AKe.
figure 5

Apparent dissociation constants and enthalpy on binding were measured using ITC. The baseline-corrected instrumental response is shown in the upper panels, and the integrated data (squares), together with the best fits (solid lines) according to a one-site binding model, are shown in the lower panels for (a) WT AKe, (b) M1 and (c) M2. Fitted parameters are summarized in Table 1.

Table 1 ITC data for ATP binding to AKe variants at pH 7 and 25 °C

The structural analysis showed that helices α6 and α7 are partially unfolded in M1 and M2; thus, ΔHITC of ATP binding includes folding of these helices. The difference in calorimetric enthalpy for the two mutants scales with the degree of unfolding in ATPlid observed using CD and NMR spectroscopy. The entropic term is in favour of ATP binding in the case of WT AKe, presumably because of the release of solvent molecules hydrating both the substrate and the enzyme. In contrast, both M1 and M2 display strong enthalpy/entropy compensation for ATP binding. The large enthalpic components in M1 and M2 are paralleled with a significant entropic penalty. This may be explained by restriction of the main chain as a consequence of the coupled folding/binding event in the ATPlid in these mutants.

Preservation of conformational dynamics in mutated AKe

Chemical shifts can be used to distinguish between open and closed conformations of AK. WT AKe in complex with Ap5A mainly populates the closed conformation16. The 1H-15N HSQC spectrum of M2, in complex with Ap5A (Supplementary Fig. S3), is of high quality and closely resembles the spectrum of Ap5A-bound WT AKe16. Residues with significant chemical shift perturbations are localized to regions surrounding the sites of mutations (Supplementary Fig. S4). The similarity between WT and M2 HSQC spectra is a strong indication that M2 is adopting a structure in complex with Ap5A that closely resembles the closed AKe conformation. Interestingly, most of the resonances that are missing (because of unfolding) in ATPlid in the M2 apo state are visible and assigned in the Ap5A complex. Apparently, M2 undergoes a coupled folding and binding event on interaction with Ap5A. To directly verify that M2 is populating a closed AKe conformation in complex with Ap5A, we made use of a chemical shift-based method previously developed in our laboratory16. There exists a strong linear correlation with a slope of close to unity when chemical shift perturbations between M2 and WT (both bound to Ap5A) are compared (Supplementary Fig. S5). The linear correlation (with a slope near one) is evidence for a closed M2 conformation in complex with Ap5A, and enables the use of chemical shifts of M2 in complex with Ap5A as residue-specific markers of a fully closed enzyme. As α6 and α7 are unfolded to a degree in between M2 and WT, it is safe to assume that M1 populates a closed conformation in complex with Ap5A as well. The results show that M2 and M1 are functional enzymes that retain the ability to adopt the closed state.

Structure of AKe:ATP complexes

The structural response to ATP binding in M1 and M2 was addressed by quantifying chemical shift perturbations in protein samples saturated with ATP. The quantitative analysis is restricted to the M1 variant as assignments of the ATPlid are missing in M2. However, the 1H-15N HSQC spectra of ATP-saturated M1 and M2 are very similar (Supplementary Fig. S6), implying that the corresponding ATP complexes adopt similar structures. Notably, the number of observable resonances of M2 is significantly increased in the M2:ATP complex compared with the apo state (compare Fig. 2b and Supplementary Fig. S6), which is complementary evidence of a coupled folding/binding reaction, in agreement with the ITC data. In Supplementary Figures S7 and S8, the chemical shift perturbations in response to ATP binding for M1 and WT AKe are summarized. The spatial locations of residues that are most strongly affected by ATP binding are very similar in both enzyme variants. Apparently, M1 binds to ATP with the same interaction surface as WT and as such is a functional enzyme with respect to ATP binding. This inference is valid for M2 as well, in light of the resemblance of M1 and M2 NMR spectra in ATP-saturated states.

Mutations modulate the open-to-closed equilibrium constant

We have previously shown that the ATPlid in AKe interconverts between open and closed conformations in the presence of bound ATP16. A detailed description of the method used to quantify populations is provided in the Supplementary Methods and in Supplementary Figures S9, S10. Equilibrium open and closed populations of both WT and the M2 variant in their ATP-bound states were determined by comparing chemical shifts in apo (open), Ap5A-bound (closed) and ATP-saturated states (Fig. 6). As assignments of ATPlid in the M2 apo state was missing, the analysis was centred on the N- and C-terminal α-helices (inset Fig. 6a). Both these α-helices respond structurally to ATP binding in concert with ATPlid and are valid markers of the open/closed equilibrium (Supplementary Information Fig. S2 and Table S2 in Ådén and Wolf-Watz16). For WT AKe, the population of closed ATPlid in the ATP complex is 0.6, in good agreement with previous analysis16. The fraction of closed ATPlid is significantly increased in the M2:ATP complex, and, within the experimental uncertainty, the closed population is near 1. Apparently, the conformational equilibrium for the open/closed transition in the presence of ATP is strongly skewed towards the closed state as a consequence of the structural perturbation. As the NMR spectra of the ATP-saturated M1 and M2 variants are very similar, it appears likely that a similar shift in the open/closed equilibrium is valid for M1 as well.

Figure 6: Shift of open/closed equilibrium in ATP-saturated M2 versus WT AKe.
figure 6

Normalized chemical shift perturbations in ATP- and Ap5A-bound states referenced to the apo states are calculated according to δω=0.2Δ15N+Δ1H (p.p.m.). (a) Correlation between ATP and Ap5A chemical shift perturbations for WT AKe. The best-fitted straight line (red) has a slope of 0.62±0.07 with an R coefficient of 0.97. In all, 20 residues were used in the analysis. Inset: positions of the residues analysed in WT and M2 are coloured gold on the open AKe structure. (b) Correlation between ATP and Ap5A chemical shift perturbations for the M2 variant. The best-fitted straight line (red) has a slope of 1.07±0.1 with an R coefficient of 0.97; 19 residues were used in the analysis.

Catalytic activities of AKe variants

As product release is rate limiting for AKe catalysis10,19, the increased ATP-binding affinities of the M1 and M2 mutants are expected to hamper catalytic activities. Kinetic parameters with ATP as the variable and AMP as the constant substrate were quantified for M1, M2 and compared with that of WT AKe (Table 2). As expected, kcat values were significantly reduced (40-fold) for the two mutants, and the reduction in kcat was very close to the (50-fold) increase in ATP-binding affinity. In addition, the increased ATP-binding affinities were reflected in decreased KATPM values for the two mutants.

Table 2 Kinetic parameters for AKe catalysis with ATP as variable substrate at 25 °C

Discussion

The two AKe variants studied here have segmentally disordered ATPlid domains at 25 °C. We have previously proposed that the entire ATPlid functions as one cooperatively folding unit17. In this study, we identified additional layers in complexity in the folding behaviour of ATPlid, namely, that α6 and α7 forms one folding unit and the INSERT segment forms a second. The mutated variants have significantly reduced vant't Hoff enthalpies of thermal unfolding but with only modest reduction in Tm. This small perturbation in Tm is a consequence of the weak thermodynamic coupling between ATPlid and CORE subdomains. From a functional perspective (in terms of ATP binding), the mutated variants display WT-like properties. First, the chemical shift perturbation patterns inflicted by ATP binding are strikingly similar to the WT, showing that the same interaction surface is used for ATP interaction in mutated and WT AKe. Second, both mutated variants can populate fully closed conformations in complex with the inhibitor Ap5A; thus, the enzymes retain the ability to close its substrate-binding motifs.

As a consequence of the coupled folding/binding equilibrium in the M1 and M2 variant, the enthalpic contribution for ATP binding is significantly larger compared with that of WT AKe because of the cumulative effects of folding and binding enthalpies. The decreased enthalpies of ATP binding (favouring binding) are accompanied by increased entropic components (disfavouring binding) in the mutants. Such enthalpy/entropy compensation has been observed in other proteins that mechanistically utilize coupled folding/binding reactions to interact with substrates31.

It is remarkable that structural mutations that destabilize the ATPlid are paralleled with an 50-fold increase in ATP-binding affinity. To identify the underlying molecular explanation, we analysed the formal expression for Kappd following Figure 1a and obtained Equation (2) (see Supplementary Methods for derivation).

In Equation (2), Kd is the dissociation constant and Kconf is the equilibrium constant for the open-to-closed equilibrium in the presence of bound ATP (Supplementary Equation S4). An inspection of Equation (2) shows that the decreased Kappd can originate from reduced Kd or increased Kconf values, relative to WT AKe. It is unlikely that Kd is decreased in the mutants, as the structure of ATPlid is perturbed in both M1 and M2. On the other hand, the equilibrium constant for the open–closed transition (Kconf) in the presence of bound ATP is significantly skewed in the M2 variant compared with WT AKe. In WT AKe and the M2 variant, the experimentally determined values of fractions closed ATPlid are 0.62±0.07 and 1.07±0.1, respectively. Assuming that Kd adopts equal values in WT, M1 and M2, and with a WT Kconf of 1.5 (0.6/0.4), Equation (2) was used to calculate a value of Kconf equal to 124 that accounts for the observed reduction in Kappd (from 50 μM in WT to 1.0 μM in the mutants). This value translates to fractions of open and closed conformations of around 0 and 1, respectively, in excellent agreement with our experimental observations. Apparently, the increased affinity for ATP in the mutant AKe variants is fully explained by a large shift in the open-to-closed equilibrium constant. If, on the other hand, mutations would stabilize a 'binding incompetent state', that is, a disordered ATPlid on a reaction pathway towards global protein unfolding, the observed binding affinity would decrease (Supplementary Equation S6). As the ATP-binding affinity is increased as a result of the ATPlid destabilization, our mutations modulate the native free-energy landscape on the reaction trajectory for subdomain closure. The effect of the mutations is to redistribute states that are present in WT AKe. It might seem intuitive that the chemical shifts of the mutant AKe variants should fall in between the chemical shifts of open and closed WT AKe. However, a detailed consideration of chemical shifts (Supplementary Note and Supplementary Fig. S11) shows that this correlation is not expected. In good agreement with our model for catalysis (that is, rate-limiting dissociation of bound nucleotides), the M1 and M2 variants display 40-fold decrease in catalytic efficiency compared with WT AKe.

The fact that mutations redistribute already existing states is in agreement with arguments discussed previously7. On the basis of these arguments, we propose that local disordering of α6 and α7 is part of the reaction trajectory for closure of the ATPlid (Fig. 7). As both the stable ground states (open and closed) contain well-folded α6 and α7 helices, the conformational change follows an order-disorder-order mechanism, in agreement with the cracking model25,26. To the best of our knowledge, this is the first experimental observation of such a mechanism. The energetic prerequisite for our mechanistic model is that the local thermodynamic stability of α6 and α7, is sufficiently low to accommodate an unfolding/folding reaction. This boundary condition has been observed for AK isolated from three different organisms17,32. Thus, the energetic prerequisite for our model is generally met in the AK fold. The low structural stabilities in these key regions are also reflected in high B-factors in the substrate-free AKe structure13,17. Both residues mutated here (Ile 116 and Leu 168) are strongly conserved within the AK family27. In fact, the entire hydrophobic cluster (positions 116, 117, 164 and 168) in ATPlid identified by us previously17 is conserved with a consensus sequence equivalent to the AKe sequence. In our proposed model, disruption of this cluster is part of the mechanism for ATP binding and, given the strong sequence conservation, the mechanism is likely to be general in the AK family. Protein folding is usually described using funnel-shaped energy landscapes5 in which the native state exhibits minimal frustration33; our results with an unfolding/folding reaction during a functional transition point towards an overlap between folding and functional energy landscapes. It is possible that the novel mechanism used by AKe during the open-to-close transition is also used by other proteins that undergo conformational changes.

Figure 7: Model for ATPlid closure in AKe.
figure 7

Relative free-energy differences between open and closed ATPlid conformations in the presence of saturating ATP concentration for WT (a) and the M2 variant (b). The relative change in free energy (ΔΔGconf) is calculated from Kconf values of 1.5 and 124 (124 equals the minimum value that accounts for the increased ATP-binding affinity in M1 and M2 compared with WT) for WT and the M2 variant. The y axis in (a) and (b) is set arbitrarily such that the absolute values of the free energy for open, closed and transition state (TS) structures may in fact differ for the two enzymes. (c) Structural model of the conformational change in the ATPlid in the presence of bound ATP. The ATP complex is represented with yeast AKe in complex with the non-hydrolysable ATP analogue AMPPCF2P45. The structures going from open to partially unfolded to closed are arranged as a function of the reaction coordinate. The relative positions of the WT and the M1, M2 variants are indicated on the reaction coordinate and should be interpreted as a qualitative measure of the distance to the transition state.

A subpart of the mechanism identified for the conformational change in AKe, that is, disorder–order transitions, has been found in interactions between proteins and a diverse set of substrates. Disorder–order mechanisms coupled to substrate binding have been observed, for instance, in protein–DNA34, enzyme–substrate31,35, protein–peptide36 and protein–protein37 interactions.

The folding of small proteins into a well-defined and compact active state is most often a highly cooperative process38 without a significant population of non-native states. In a recent study, it was suggested that the cooperativity of the folding reaction is the result of natural selection39. It is intriguing that AKe uses cooperative unfolding/refolding to accommodate a large-scale conformational change during catalysis. An appealing hypothesis is that nature has made use of cooperative unfolding/refolding to add functional properties (such as conformational plasticity) to proteins.

Methods

Protein preparation

All AKe variants were prepared as described previously19. The concentrations of all proteins were determined using an extinction coefficient at 280 nm of 10.000 M−1 cm−1.

Site-specific mutagenesis

Primer sequences for the M1 and M2 mutants were ordered from CyberGene. PCR-amplified DNA sequences encoding the mutant proteins were subcloned into the expression vector pET-3a (Novagen) using restriction enzymes Nde1 and BamHI. The final constructs were verified using DNA sequencing (Eurofins MWG Operon).

Kinetic assays

Enzyme kinetic parameters for ATP turnover in the direction of ADP formation were quantified at 25 °C using a coupled spectroscopic assay10. The ATP concentration was varied and the AMP concentration was fixed at 0.5 mM.

NMR spectroscopy

All NMR experiments were recorded on samples containing 0.7–1.2 mM 15N-labelled proteins in 50 mM NaCl and 30 mM MOPS at pH 7.0 with 10% (v/v) 2H2O. NMR experiments were carried out on a Bruker DRX 600 MHz spectrometer equipped with a 5-mm triple resonance z-gradient cryoprobe. The temperature was calibrated using an external probe inserted into the sample compartment of the NMR probe. Backbone resonance assignment of apo and substrate-saturated M1 and M2 mutants was obtained with 15N nuclear overhauser effect spectroscopy-HSQC and HNHA experiments40. ATP and Ap5A (Sigma-Aldrich)-saturated enzymes were obtained by adding 20- and 2-mM substrate to protein samples, respectively. NMRPipe software41 and ANSIG for Windows42 were used for NMR data processing and visualization of spectra, respectively.

ITC

ITC experiments were carried out at 25 °C using an iTC200 Microcalorimeter (Microcal). The reference cell was filled with sample buffer, and protein-containing solutions were filled in the sample cell (200 μl volume) and titrated with ATP (40 μl). Substrate solutions were prepared in buffer (same batch) from the final step of protein purification (that is, gel filtration). Protein concentrations were 1.5 mM and 40 μM for WT titrations with ATP and Ap5A, respectively. M1 and M2 were at 20 μM for ATP titrations. The substrate concentration in the injection syringe was 10 times higher than the concentration of protein. A typical experiment consisted of an initial control injection of 1 μl, followed by 16 injections, each of 2.5 μl and 5 s duration, with a 180 s interval in between. ITC measurements were routinely taken in 50 mM NaCl and 30 mM MOPS at pH 7.0. In control experiments, the ligand (ATP) was injected into the buffer. Raw data were collected, corrected for ligand heats of dilution and integrated using the Microcal Origin software supplied with the instrument. A single-site binding model was fit to the data by a nonlinear regression analysis to yield binding constants (Kd), observed enthalpies of binding (ΔHITC) and stoichiometry of binding (n).

CD spectroscopy

Far-ultraviolet CD spectra (190–260 nm, 25 °C, 1-mm path length) and temperature scans (15–85 °C at 222 nm, 0.1-mm path length) were collected in 10 mM Na-phosphate and 50 mM NaCl (pH 7.0) on a Jasco J-810 Spectropolarimeter. Thermal denaturation curves were fitted to Equation (3) assuming a two-state reaction. The reversibility of the thermal transitions was larger than 80%. Protein concentrations in CD experiments ranged between 5 and 10 μM.

Where

Here, Sobs is the temperature-dependent spectroscopic signal, mf, mu and Sf, Su are the slopes and intercepts of the folded and unfolded baselines, respectively, T is the absolute temperature, Tm is the melting point, Ku is the unfolding equilibrium constant and ΔHvH is the van't Hoff enthalpy of unfolding at Tm.

Additional information

How to cite this article: Olsson, U. & Wolf-Watz, M. Overlap between folding and functional energy landscapes for adenylate kinase conformational change. Nat. Commun. 1:111 doi: 10.1038/ncomms1106 (2010).