Jetmir
Haxhija
,
Felix
Guischard
and
Thorsten
Koslowski
*
Institut für Physikalische Chemie, Universität Freiburg, Albertstrasse 21, 79104 Freiburg, Germany. E-mail: Thorsten.Koslowski@physchem.uni-freiburg.de
First published on 30th September 2023
We estimate the entropic contributions to the free energy of quinone unbinding in bacterial and mitochondrial respiratory chains using molecular dynamics (MD) and Monte Carlo (MC) computer simulations. For a varying length of the isoprenoid side chain, MD simulations in lipid bilayers and in unpolar solvents are used to assess the dihedral angle distributions along the chain. These form the basis of a MC estimate of the number of molecular structures that do not exhibit steric self-overlap and that are confined to the bilayer. We obtain an entropy drive of TΔS = 1.4 kcal mol−1 for each isoprene unit, which in sum is comparable to the redox potential differences involved in respiratory chain electron transfer. We postulate an entropy-driven zipper for quinone unbinding and discuss it in the context of the bioenergetics and the structure of complex I, and we indicate possible consequences of our findings for MD-based free energy computations.
Ubiquinones have 1,4-benzoquinone as a structural basis, which is substituted at the positions 5 and 6 by methoxy groups. These molecules carry an isoprenoid side chain at position 3 and a methyl group at position 2. Ubiquinones are classified according to the degree of polymerization of the side chain, n, as UQ n. In human mitochondria, UQ10 is operative as an electron and proton carrier, rodents have UQ9 and yeast UQ6.
Menaquinones are electron carriers used primarily by bacteria. Here, the methoxy groups attached to the positions 5 and 6 are replaced by a benzyl ring, cf.Fig. 1 (note that the number of atom 6 changes to 10). The introduction of methoxy groups makes the head group of the quinone slightly more demanding from a steric point of view, and also slightly more hydrophobic. Menaquinones are found in the inner membrane of bacteria and are involved in the transfer of electrons in cellular respiration and other metabolic processes. Examples include Bacillus subtilis, Streptococcus pneumoniae and Mycobacterium tuberculosis.
![]() | ||
Fig. 1 Structure formulae of menaquinone and menaquinol, with the enumeration of atoms as used in the text. |
In photosynthesis, plastoquinones play a similar role. They are located in the thylakoid membrane of chloroplasts and are involved in the transfer of electrons from photosystem II to photosystem I. This electron transfer is part of the reactions that generate ATP and NADPH, which are used in the light-independent reactions of photosynthesis. Plastoquinones do not carry a methyl group in position 2, but a simple hydrogen. The two methoxy moieties become methyl groups.
Ubiquinone is involved in several human diseases and disorders. Q10 levels are often low in individuals with cardiovascular disease7,8 and diabetes,9 although cause and effect are not always clear. As ubiquinone is vital to the respiratory chain, deficiencies can lead to mitochondrial disorders such as mitochondrial encephalomyopathy.10
We identify the quinones by their chemical structure (menaquinone MK, ubiquinone UQ or plastoquinone PQ), state of reduction (the respective hydroquinones or quinoles are referenced as MHK, UHQ and PHQ) and the chain length, which will be added to the identifier as an integer. We focus on MKn with n = 4, 6 and 8 in the MD simulations described below, and on n = 2 to 10 in the MC approach.
In the last years considerable effort has been devoted to investigate the exact mechanism of the respiratory complex I. Mechanistically and structurally, the complex can be divided into two domains. In the hydrophilic part, electrons are transferred from NADH to FMN, which transfers them to ubiquinone via FeS clusters.11,12 In the membrane part, four protons are then translocalized to the periplasmatic side. The interplay of both mechanisms is referred to as proton coupled electron transfer, although proton pumping is still not fully understood. Due to the availability of high resolution structures and the advances in computing power, molecular dynamics simulations have been widely used to investigate the exact mechanism of respiratory complex I.13,14 Varying proposals for the proton–pumping mechanism have been reported.15–17 In combination with biochemical data the reduction of quinone actives a reaction cascade combined of electrostatic and conformational changes that lead to the proton transfer.16
Complementary to these large-scale MD simulations, we focus on a single aspect of the bioenergetics of quinones, the contribution of the conformational entropy of the isoprenoid side chain to the free energy of protein–hydroquinone unbinding. We consider the confining environment of a lipid bilayer an assume that the number of conformations in the quinone binding pocket of the protein is small compared to those within the layer. In this way, we are able to estimate a lower boundary to the entropy.
The remaining part of the article is organized as follows. In the next section we describe molecular dynamics simulations of menaquinones embedded in a lipid bilayer and in n-hexane. The resulting dihedral angle distributions of the menaquinones’ oligoisoprene chain and the concentration profiles are used as the input to a Monte Carlo microcanonical computation of the entropic contributions to ubiquinone unbinding. We discuss our findings and draw conclusions in the final section of this paper.
As described above, molecular dynamics simulations have been performed for MHK4, MHK6 and MHK8 in a lipid bilayer, which is embedded into an aqueous environment. To address the question to which extent geometrical constraints play a role, MHK6 was also simulated in n-hexane as a hydrophobic solvent. Test simulations have also been performed for the oxidized quinones, we find no significant differences to the hydroquinones.
Concentration profiles for the MHK8 menaquinol simulations within a lipid bilayer are shown in Fig. 3. The arrangement of lipid and water molecules shows the familiar picture of a lipid double layer in contact with the aqueous phase via the hydrophobic head groups and a hydrophobic core formed by a double layer of the hydrophobic lipid tails. For the quinone, we find a binodal profile. The first maximum – representing the quinone head group – is embedded in a neighbourhood of lipid molecule head groups and water molecules. The second, corresponding to the oligoisoprene chain, is slightly broader and is located ∼7 Å deeper within the bilayer, i.e. in a purely hydrophobic environment. It does not show any overlap with the water or the hydrophilic head group profile. As a consequence, we consider the isoprenoid side chain of MHK8 to be confined to the hydophobic core of the lipid bilayer. The side chain is attached to a naphtoquinone that is localized in an amphipilic environement between the head and the tail groups of the lipid. We will use this information as input to the Monte Carlo model described below. Throughout the simulations, no flips of the quinone molecules have been observed. The side chain profile is typical of large-chain polymers confined to two dimensions, where the probability distribution strongly peaks at the center.30
In the following, we focus on the dihedral angles of the isoprenoid chain beyond the first double bond. Within the accuracy of the simulations and the analysis, they reflect the periodicity of the chemical structure. Sample distributions are displayed in Fig. 4. The complete side chain distributions are supplied as ESI.† As expected, the conformational space of the molecules is not fully explored for all but the smallest side chains, an effect that has also been described for proteins.31 On the time scale of the simulation, phase space is not sampled properly, a phenomenon sometimes referred as ergodicity breaking. This behaviour is evident from the number of unique dihedral angle conformations as a function of the simulation time, as presented in the ESI.† This finding is part of our motivation for switching from molecular dynamics to Monte Carlo simulations.
The first dihedral angle, ϕdcba, always exhibits two maxima at ∼80 and ∼272 degrees. Regardless of the environment – lipid or n-hexane – an additional maximum shows up at zero degrees for MHK6 and MHK8, whichs is missing for MHK4. As the latter dihedral angle corresponds to a U turn of the chain, we assume that its introduction to a small chain is sterically prohibited, while it can evade this situation for longer chains. We are mainly interested in the energetics of the larger chains, and will consequently use all three dihedral angles ϕdcba ∈ {0,86,272} in the Monte Carlo setup. As the maxima are all of a similar height, we can safely assume that they belong to isoenergetic minima of the dihedral angle potential energy surface.
For ϕadcb, we find three well-separated maxima of the angle probability distribution at ∼62, ∼178 and ∼295 degrees with a dominant peak at 178 degrees. As the maxima exhibit different heights, we have to check whether this non-uniformity stems from a steric overlap that is also taken into account in the Monte Carlo model, or whether it has a different origin. As we will demonstrate below, the first assumption is valid, and we consistently recover the weights of the maxima in the Monte Carlo simulations. Hence, we can use all three dihedral angles, idealized as ϕadcb ∈ {60,180,300} in the microcanonical Monte Carlo setup with an equal weight.
For the dihedral angle ϕbadc, we observe two maxima at ∼80 and ∼272 degrees for MHK6 (lipid bilayer and n-hexane environment) and MHK8. Only for MHK4, an additional maximum around 180 degrees can be observed. For larger chains, vestiges can be found as a small nonzero contribution to the dihedral angle distribution, which renders the dominant parts asymmetric. Focussing on the larger chains, we can conservatively use idealized angles ϕbadc ∈ {90,270} in the Monte Carlo setup.
The dihedral angle ϕcbad is always fixed due to the presence of a double bond. For the same reason, any dihedral angle involving e is also fixed, such as ϕebad.
As a model for the molecule, we use a rigid benzoquinone as a basis to randomly grow a chain. Motivated by the localization patterns obtained from the MD simulations, we place the two oxygens in the x–y-plane, which we define as separating the hydrophilic head and the hydrophobic tail groups of the lipids. The minor axis of rotation is oriented along the z axis, with the methyl group and the chain pointing towards negative z values. This orientation is motivated by our analysis of the MD simulations for MHK4, where we find a direction cosine of the minor rotational axis of z/r = 0.92 ± 0.02. To grow the chain, we make use of standard single (1.54 Å) and double (1.34 Å) bond lengths and idealized sp2 and sp3 bond angles of 120 and 109.45 degrees, respectively. For each carbon atom added to the chain, the dihedral angle is either fixed (if involving a double bond) or randomly drawn from a set of two or three values. They stem from our analysis of the MD simulations and are – again idealized – given by ϕdcba ∈ {0,90,270} ϕadcb ∈ {60,180,300} and ϕbadc ∈ {90,270} degrees. The geometries are build making use of the xyzatm subroutine of the TINKER molecular modeling program package.34 Hydrogen positions are not generated explicitely, as we consider these atoms in the frame of a united atom approach. They are characterized by fused parameters with the carbon atoms, and they share their positions. For each fully grown chain we check for a steric overlap within the chain or of the chain and the benzoquinone. Optionally, we record whether a chain atom is located within a layer of −20 Å < z < 0, corresponding to a very conservative estimate of the size of the hydrophobic core of the lipid bilayer, cf.Fig. 3. As an overlap criterion, we take a united atom carbon–hydrogen van der Waals diameter of 3.5 Å, as for example used in the OPLS/UA force field.35 Overlaps up to and including 1–4 interactions are not taken into account.
We assume that all non-overlapping chains have the same energy. This implies that – within statistical inaccuracies – all dihedral angle distributions originating from the MD simulations have pronounced maxima of the same height (or that non-uniform distributions are self-consistently recovered), that the dihedral angles are uncorrelated and that details of the interactions do not play a significant role. In this regime, we have two energies, one that is identical for all non-overlapping conformations and can be set to the zero of the energy scale, and one that is infinite for all overlapping conformations. As a consequence, the system can be described thermodynamically by a microcanonical ensemble of Ω non-overlapping microstates, and the entropy is given by the Boltzmann expression S = kBln
Ω. Due to the large number of states, this expression can not be evaluated directly, but we have to resort to sampling a randomly generated subset. We make use of a simple hit-and-miss approach36 with mshot attempts to create a chain, which leads to mhit chains that are accepted w.r.t. to the criteria listed above. We can then estimate Ω ≃ m × mhit/mshot or
ln![]() ![]() ![]() ![]() | (1) |
Self-overlapping configurations and those protruding out of the hydrophobic core of the lipid bilayer do not contribute to mhit, and we are confident that to reject them is correct, as none of these are encountered in the MD simulations, either.
A sample realization of a hit-and-miss algorithm is provided by Allen and Tildesley as a Fortran program (code 4.1.).36 Applications of hit-and-miss methods date back to Lazzarini's experimental approach to Buffon's needle problem, a geometrical problem related to the determination of the number π.37 Apocryphically, Ulam's idea to use a hit and miss method to estimate the number of decks that lead to a successful solution of the solitaire card game has initialized Monte Carlo methods on electronic computers.38
n | m hit | −TS (kcal mol−1) |
---|---|---|
2 | 2221080 | −1.49 |
3 | 955493 | −2.96 |
4 | 399675 | −4.42 |
5 | 149060 | −5.80 |
6 | 56240 | −7.19 |
7 | 20620 | −8.57 |
8 | 7757 | −9.96 |
9 | 2913 | −11.35 |
10 | 1129 | −12.76 |
A least squares linear regression analysis assuming −kBTS = a − bn leads to a = 1.24 ± 0.03 kcal mol−1 and b = 1.402 ± 0.005 kcal mol−1. As expected, the entropy vanishes for a chain consisting of a single isoprene unit, as we fix the first two and the final dihedral angle of each chain. As an entropic contribution to the free energy of quinone unbinding from a situation in which all dihedral angles are fixed, we observe a gain of 1.4 kcal mol−1 per isoprene unit. The analysis is shown in the ESI,† including the theoretical limit of a model that permits overlap and is not confined.
The confinement of the chain to the lipid bilayer has a small effect on the entropy. For n = 10 – the maximum chain length studied here – we observe mhit = 1129 out of 5 × 106 configurations that do neither exhibit collisional overlap nor protrude out of the hydrophilic core of the membrane. Dispensing with confinement to the membrane, we find mhit = 3340, which amounts to a free energy difference of 0.65 kcal mol−1, which is close to kBT at room temperature.
We recover the non-uniform distribution of the dihedral angle ϕadcb in the number of ideal dihedral angles that contribute to successful realizations of the isoprenoid random walk model. For n = 10, we have 1168 hits for ϕadcb = 90 degrees, 1142 hits for ϕadcb = 270 degrees and 7851 hits for ϕadcb = 180 degrees. We note that ϕbadc is symmetric within the limit of fluctuations with 5151 hits for 90 degrees and 5010 hits for 270 degrees. Notably, for ϕdcba a zero dihedral angle can be less frequently observed (2616 hits) than those of 90 and 270 degrees (3855 and 3690 hits, respectively). On the other hand, the MD simulations show ϕdcba dihedral angle maxima of roughly equal height. Thus, the MC model exhibits less U turns than the MD simulations, which suggests that overlap is slightly penalized too strong in our MC model.
Using a simple hit-and-miss algorithm, we have constructed and counted non-overlapping isoprenoid random walks confined to a lipid bilayer, thus counting the number of accessible dihedral angle microstates from a subset of all possible states. By the Boltzmann relation, we find a corresponding free energy release of 1.4 kcal mol−1 per isoprenoid unit.
In the following, this number is compared to other energy scales relevant to charge transfer in complex I, e.g. the redox potential, which is proportional to the thermodynamic driving force, −eΔE0 = ΔG0 = ΔH0 − TΔS, or to the strength of hydrogen bonds. From our calculations, we cannot provide a serious estimate of ΔH. This is due to the lack of an experimentally determined structure with a quinone localized within the protein binding pocket and its chemically specific interactions, and to the possible dissipation of energy along the electron transport chain.
In terms of the redox potential, the entropic contribution to quinone unbinding amounts to a substantial value of 61 mV per isoprene unit. Hence, quinone redox potentials in protein binding–unbinding reactions are not only affected by quinone–amino acid interactions, but also by a release of conformational entropy. Other factors may also play a role in reducing the number of conformations during the release of hydroquinones from a protein: the isoprenoid chain may overlap with atoms from the host protein, or other protein complexes from a hypothetical supercomplex. Consequently, our computations serve as an estimate of the entropy in a more complex environment.
The entropy contribution to the free energy generated by releasing menaquinoles is considerably larger than the redox potential differences of the redox pair of NAD+/NADH and MK/MHK. Although the exact numerical value of this difference in a complex biological environment is not yet known, it is save to assume that it is close to the corresponding difference in the standard potentials:39–41 we have an electronic driving force of ∼400 meV or ∼8 kcal mol−1, a value slightly smaller than the dihedral angle entropy of the MHK10 isoprenoid chain. This implies that the release of conformational entropy has to be coupled to an endergonic process in the vicinity of the MK binding pocket, which has to occur almost simultaneously. Breaking the two hydrogen bonds between the menaquinol hydroxyl groups and their coordinating Nqo4 amino acids amounts to an estimated 12 kcal mol∼, as we have a O⋯H–O bond to Tyr87 and a O⋯H–N bond to His38 with ∼7 and 5 kcal mol−1, respectively.42
There is, however, an additional angle, as the driving force is an entropic one. In each cycle of MK reduction, 2–3 protons are pumped against a pH gradient. These protons are in turn required to catalytically convert one ADP into an ATP, a reaction that stores ∼6 kcal mol−1 and that takes place in a different enzyme. It is difficult to imagine that ATP synthesis is accompanied by a similar entropic waste generated by the MHK release from its binding pocket. Consequently, the MHK release has not only to fulfill a free energy balance, as outlined above, but also an entropic one. This is subject to the same criteria, viz. those of a spatial proximity and a immediate coupling in the chain of chemical reactions or structural rearrangements.
In their determination of the structure of the entire complex I of Th. thermophilus, Baradaran et al. have noted an unusual feature of the MK binding funnel:13 it is lined with amino acids that have a charged or strongly dipolar character: they find His34, Arg216, Glu223, Glu225, Glu248, Tyr249, Arg294 and Arg299, in the Nqo8 subunit and His1, Arg36 and Asp62 in the Nqo6 subunit of the complex (cf. Fig. 4(d) of ref. 13). The latter amino acid is not resolved in the protein database (PDB) entry 4HEA. These amino acids may be of some assistance in transiently binding to the MHK head group in an energetically downhill cascade while leaving the binding site. Their interaction with the hydrophobic tail, however, can only be assumed as weak, unspecific and non-directional. Only upon the release of MHK, they are likely to form pairs or other aggregates with a reduced number of degrees of freedom, simultaneously compensating the monomer-by-monomer entropy production of an oligomer chain leaving its binding pocket. We refer to this model as an entropic zipper, and its essential elements are illustrated in Fig. 5 as a cartoon model. Verification or falsification of this hypothesis has to await structural models that compare vacant binding pockets to those populated by long-chain quinones or comparable inhibitors.
Finally, we note that entropy – or heat – production is essential for endothermic organisms like mammals or birds, which may explain an evolutionary drive towards elongated isoprenoid chains for these classes of animals.
In molecular dynamics simulations, free energies can be computed by thermodynamic integration or umbrella sampling techniques. The latter requires the introduction of a potential – usually harmonic – to a subset of the atoms. In this way, an oligomer chain can be gradually pulled out of its binding pocket. It is crucial to select the atoms affected by the external potential in a way that retains the full flexibility of the chain upon unbinding to recover the full entropic contribution to the free energy. As an alternative, our computations may serve as a correction, e.g. for the unbinding of a stretched chain.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d3cp03466f |
This journal is © the Owner Societies 2023 |