Abstract
Selective 15N isotope labeling of the cytochrome bo3 ubiquinol oxidase from E. coli with auxotrophs was used to characterize the hyperfine couplings with the side-chain nitrogens from R71, H98, and Q101 residues and peptide nitrogens from R71 and H98 residues around the semiquinone (SQ) at the high-affinity QH site. The 2D ESEEM (HYSCORE) data have directly identified the Nε of R71 as an H-bond donor carrying the largest amount of the unpaired spin density. In addition, weaker hyperfine couplings with the side-chain nitrogens from all residues around the SQ were determined. These hyperfine couplings reflect a distribution of the unpaired spin density over the protein in the SQ state of the QH site and strength of interaction with different residues. The approach was extended to the virtually inactive D75H mutant, where the intermediate SQ is also stabilized. We found that the Nε from a histidine residue, presumably H75, carries most of the unpaired spin density instead of the Nε of R71, as in the wild-type bo3. However, the detailed characterization of the weakly coupled 15Ns from selective labeling of R71 and Q101 in D75H was precluded by overlap of the 15N lines with the much stronger ~1.6 MHz line from quadrupole triplet of the strongly coupled 14Nε from H75. Therefore, a reverse labeling approach, in which the enzyme was uniformly labeled except for selected amino acid types, was applied in order to probe the contribution of R71 and Q101 to the 15N signals. Such labeling has shown only weak coupling with all nitrogens of R71 and Q101. We utilize density functional theory based calculations to model the available information about 1H, 15N and 13C hyperfine couplings for the QH site and to describe the protein-substrate interactions in both enzymes. In particular, we identify the factors responsible for the asymmetric distribution of the unpaired spin density and ponder the significance of this asymmetry to the quinone’s electron transfer function.
Cytochrome bo3 ubiquinol oxidase (cyt bo3) from Escherichia coli catalyzes the reduction of molecular oxygen to water using ubiquinol as the electron donor.1 The enzyme located in the cytoplasmic membrane also functions as a proton pump, conserving much of the energy available from the redox reaction as the proton motive force.2,3 Cyt bo3 is a member of heme-copper superfamily. Of the four subunits of cyt bo3, the catalytic subunit and two other subunits are analogous to the mitochondrially encoded subunits of aa3-type cytochrome c oxidase (cyt aa3).4 Despite having different electron donors, the proton pumps of cyt bo3 and cyt aa3 likely operate in a similar manner.5
Previous work has established that cyt bo3 isolated in the detergent n-dodecyl β-D-maltoside (DDM) is associated with a tightly bound ubiquinone-8 (UQ8) at the QH site, whereas purification in Triton X-100 yields the enzyme without any bound UQ8.6,7 The UQ8 bound at the QH site does not exchange with the substrate ubiquinone pool during turnover. Hence, the QH site is distinct from the QL site, where the oxidation of substrate ubiquinol takes place.7,8 The QH site has been shown to be able to stabilize the one-electron reduced semiquinone (SQH), which can be detected by EPR spectroscopy.9,10 The comparison of the kinetics between the enzyme preparations with and without the tightly bound UQ8 or with the QH site inhibitors led to the conclusion that the UQ8 at the QH site facilitates the fast electron transfer process from the substrate ubiquinol to the low spin heme b.6,11–13
A crystal structure of cyt bo3 without a bound UQ8 has been reported.14 Based on the crystal structure and mutagenesis studies, residues R71, D75, H98 and Q101 from subunit I were proposed to interact with the bound UQ8 at the QH site (Fig. 1).14,15 Mutating any of the four residues led to the loss of catalytic activity. Notably, the D75H mutant, despite its inability to reduce molecular oxygen, was found to stabilize a SQ radical with a midpoint potential similar to that of the wild type (WT) enzyme.15 Hence, an environment resulting in the stabilization of the SQ radical is necessary but not sufficient for proper function. A precise spatial arrangement of the SQ radical and the surrounding residues at the QH site is crucial for the successful electron transfer process.
Figure 1.
The current model of UQ at QH site of cyt bo3. The strong H-bonds are shown in dashed lines and the weak one in a dotted line. The figure was generated according to the model based on the X-ray crystal structure by Abramson et al.14 Adapted from ref. (20).
Previous X-band 1D and 2D ESEEM experiments have offered insights into the nature of interactions between the SQH and the QH site residues (Fig. 1).16–20 Hyperfine couplings with methyl group protons and exchangeable, hydrogen bonded protons, are consistent with a neutral SQH species in the WT cyt bo3, indicating significant asymmetry in the distribution of the unpaired spin density.17 The 14N ESEEM spectra reveal one strongly coupled nitrogen participating in a hydrogen bond with the SQH.16–18 In a previous communication, we reported selective 15N-labeling of cyt bo3, identifying the Nε of R71 as the strongly coupled nitrogen in the WT cyt bo3 and have shown that the other nitrogens in R71, H98 and Q101 possess substantially smaller couplings.19 In contrast, a strongly coupled nitrogen that is observed in the 14N ESEEM spectra of the SQH radical of the inactive D75H mutant was shown to be from a different amino acid residue.18 The SQH is stabilized by hydrogen bonds different from the ones identified with the WT enzyme and its characteristics are shifted towards an anion-radical.
In the current work we provide the comparative analysis of the 2D ESEEM (HYSCORE) spectra of 15N uniformly labeled (15N-U) WT and D75H cyt bo3. We report the complete results of selective 15N-labeling of the residues at the QH site in these two enzymes. In addition to the strongly coupled Nε of R71, the weakly coupled Nε of H98 was also found to carry unpaired spin density in the WT cyt bo3. It is also shown that in the D75H mutant, the Nε of H75 is the strongly coupled nitrogen, and the Nε of R71 displays significantly weaker hyperfine interaction with the SQH. The principal values of the rhombic hyperfine tensors for the strongly coupled nitrogens in WT and D75H cyt bo3 were determined using a newly developed method of HYSCORE cross-peak lineshape analysis. In addition to 15N couplings from the present work, 1H and 13C couplings from previous studies of the WT and D75H cyt bo317,18,20 were utilized in QM/MM calculations to calculate the spatial environment and electronic structure of the SQH. In particular, these calculations identify the factors responsible for the asymmetric distribution of the unpaired spin density and charge at the QH site and ponder the significance of this asymmetry to the quinone’s electron transfer function.
Materials and Methods
Materials
Isopropyl-β-D-thiogalactopyranoside (IPTG) was bought from Fisher Scientific (Pittsburgh, PA). n-dodecyl-β-D-maltoside (DDM) was purchased from Anatrace (Maumee, OH). 15NH4Cl and 15N-enriched amino acids for the growth of E. coli to express isotope labeled cyt bo3 were ordered from Cambridge Isotope Laboratories (Andover, MA). Other chemicals used in the preparation of growth medium and buffers were from Sigma-Aldrich (St. Louis, MO).
Bacterial Strains
E. coli C43(DE3) strains with deletions of genes involved in amino-acid biosynthetic pathways were constructed with the λ-Red recombination system as described previously.20,21 The C43(DE3) auxotroph strains used in the current work are listed in Table 1.
Table 1.
The genotype of E. coli C43(DE3) auxotrophs used for the preparation of selectively 15N labeled cyt bo3 samples and amino acids supplemented to the minimal medium to grow these auxotrophs.
|
E. coli C43(DE3) strains |
Genes knocked out | Amino acids added to the minimal medium |
Cyt bo3 samples prepared |
|---|---|---|---|
| C43(DE3) | none | none |
|
| CLY | cyo | none |
|
| ML8 | cyo, argH | 105 mg/L Arg |
|
| ML17 | glnA | 750 mg/L Gln |
|
| ML21 | tyrA, hisG | 18 mg/L Tyr, 15 mg/L His |
|
| ML22 | cyo, ilvE, avtA, aspC, hisG | 40 mg/L Ile, 40 mg/L Leu, 35 mg/L Val, 90 mg/L Tyr, 50 mg/L Phe, 40 mg/L Asp, 16 mg/L His |
|
| ML26 | cyo, ilvE, avtA, aspC, hisG, argH | 40 mg/L Ile, 40 mg/L Leu, 35 mg/L Val, 90 mg/L Tyr, 50 mg/L Phe, 40 mg/L Asp, 16 mg/L His, 105 mg/L Arg |
|
| ML30 | cyo, ilvE, avtA, aspC, hisG, argH, glnA | 40 mg/L Ile, 40 mg/L Leu, 35 mg/L Val, 90 mg/L Tyr, 50 mg/L Phe, 40 mg/L Asp, 16 mg/L His, 105 mg/L Arg, 750 mg/L Gln |
|
| ML32 | cyo, argH, glnA | 105 mg/L Arg, 750 mg/L Gln |
|
| ML34 | cyo, argH, glnA, hisG | 105 mg/L Arg, 750 mg/L Gln, 16 mg/L His |
|
Preparation of Amino-acid Selective Isotope Labeled Cyt bo3
Selectively 15N labeled cyt bo3 samples were expressed from the various C43(DE3) auxotroph strains as summarized in Table 1. Each auxotroph to be used as an expression host was transformed with the pET-17b vector (Novagen) encoding the WT or D75H mutant cyo operon engineered to encode cyt bo3 with a 6xHis tag at the C terminus of subunit II.18 The enzymes overexpressed from the C43(DE3) cultures were solublized in DDM detergent, purified with Ni-NTA affinity chromatography and reduced with sodium ascorbate under anaerobic condition to generate the SQH radical as described previously.20
EPR Measurements
The CW EPR measurements were performed on an X-band Varian EPR-E122 spectrometer. The pulsed EPR experiments were carried out using an X-band Bruker ELEXSYS E580 spectrometer equipped with Oxford CF 935 cryostats. Unless otherwise indicated, all measurements were made at 50 K. Some pulsed EPR measurements were also performed at higher temperatures, up to 120 K, and the results were similar to those obtained at 50 K. Several types of ESE experiments with different pulse sequences were employed with appropriate phase cycling schemes to eliminate unwanted features from the experimental echo envelopes. Among them are two-dimensional three-pulse and four-pulse sequences, which are described in detail elsewhere.17 Spectral processing of three- and four-pulse ESEEM patterns, including subtraction of relaxation decay (fitting by 3–6 degree polynomials), apodization (Hamming window), zero filling, and fast Fourier transformation, was performed using Bruker WIN-EPR software.
Some HYSCORE spectra are presented as contour plots in Matlab R14. For the 2D data representation, the echo decay was eliminated by low-order polynomial (up to fourth-order) base-line corrections in each dimension and apodizied with a Hamming window subsequently. Before the 2D Fourier transformation, the data was zero-filled up to a 1024 × 1024 matrix. The evaluated values were subsequently used to simulate the 2D HYSCORE spectra applying the Kazan Viewer package22 or home written PC software developed by Dr. A. Tyryshkin (now at Princeton University).23
HYSCORE Spectra from 15N Nuclei
The experimental data regarding the ligand environment of the semiquinone were obtained in this work from the two-dimensional (2D) 15N ESEEM (HYSCORE) spectra.24 The HYSCORE technique creates 2D spectra with off-diagonal cross-peaks. 15N nucleus with I=1/2 has two hyperfine frequencies να and νβ from opposite ms=±1/2 electron spin manifolds. These may produce a pair of cross-features (να, νβ) and (νβ, να) in (++) quadrant, as well as another pair (−να, νβ) and (να, −νβ) in the (+−) quadrant. The appearance of cross-peaks in (++) or (+−) quadrants is governed by the relative values of hyperfine couplings 15A and Zeeman frequency 15νN.25,26 Peaks in the (+−) quadrant come primarily from strong hyperfine interaction, i.e. |15A|>2(15νN), whereas the peaks in the (++) quadrant appear predominantly from interactions with |15A|<2(15νN). Peaks may appear in both quadrants simultaneously in the intermediate case when both parts of the inequalities are comparable.
Orientation-disordered (i.e. powder) 2D spectra of I=1/2 nuclei also reveal, in the form of cross-ridges contour projections, the interdependence between να and νβ values in the same orientation. The two coordinates of the arbitrary point at the cross-ridge, described in the first order by equation
| (1) |
can be used for the first-order estimate of the corresponding hyperfine coupling constant A:
| (2) |
On the other hand, analysis of the cross-ridges in (να)2 vs. (νβ)2 coordinates and spectral simulations allows in many cases for simultaneous determination of the isotropic aiso and anisotropic T components of the hyperfine tensor.25,26
Computational methods
All density functional calculations were performed using Gaussian 09.27 All calculations including geometry optimization and hyperfine coupling were performed using the B3LYP functional and the EPR-II basis set. Specific details concerning hyperfine coupling calculations are as previously described.28,29 Details of specific models used are given in the text.
Results and discussion
14N and 15N ESEEM spectra of Wild Type and D75H Mutant of Cyt bo3
The interaction of the SQH with the protein environment in WT and D75H cyt bo3 with natural abundance of nitrogen (14N isotope - 99.63 %) has been previously studied in detail by pulsed EPR spectroscopy.16–18 1D and 2D 14N ESEEM spectra show the contribution from only a single nitrogen in each protein (Figs. S1 and S2). These possess different characteristics, i.e. quadrupole coupling constant (qcc) K=0.93 MHz, asymmetry parameter η=0.51, and hyperfine coupling 14A=1.8 MHz for WT cyt bo316,17 and K=0.43 MHz, η =0.73, and 14A=2.7 MHz for D75H.18 The values of K and η characterize the chemical type and electronic configuration of 14N atom interacting with the SQH. For instance, K=e2qQ/4h=0.93 MHz, most closely corresponds to the nitrogen from an NH or NH2 group.17,18 This value is ~10% larger than the qcc for the peptide amide nitrogen and more than two times the qcc of the protonated imidazole nitrogens in histidine. Therefore, it was suggested that the most likely candidates for the H-bond donor in WT cyt bo3 are the nitrogens from the side chains of R71 or Q101. Likewise, a protonated imidazole nitrogen of a histidine residue H75 or H98 was suggested as the H-bond donor in the D75H mutant.18 14N spectra do not show any lines from other side-chain and peptide nitrogens of the nearby environment. These nitrogens are coupled more weakly and do not produce well defined lines in the 14N powder-type spectra, because of the nuclear quadrupole interactions (nqi) influence.30,31 In contrast, the lines from weakly coupled nitrogens (Nwc) are well observed in 2D 15N ESEEM spectra, which are not complicated by nqi.
Two presentations of 15N HYSCORE spectrum of the SQH in 15N-U WT cyt bo3 are shown in Fig. 2A,C (see also Fig. S3). The (++) quadrant exhibits a pair of intensive cross-peaks 1 located symmetrically around the diagonal point (15νN,15νN), along the antidiagonal with maxima at (2.74, 0.34) MHz, which correspond to hyperfine coupling 15A=2.4 MHz. This coupling is in agreement with the expected 15A of 2.5 MHz rescaled from the hyperfine coupling 14A of 1.8 MHz measured from 14N ESEEM spectra. In addition, the (++) quadrant contains feature Nwc with the maximum near the diagonal point (15ν N,15νN) with the decaying shoulders symmetrically extended up to ~0.8 MHz along the antidiagonal.
Figure 2.
Contour (A,B) presentation of the 15N HYSCORE spectra of the SQH in 15N uniformly labeled WT (A) and D75H (B) cyt bo3 (magnetic field 345.2 mT (A) and 346.1 mT (B), time between first and second pulses τ=136 ns, microwave frequency 9.702 GHz (A) and 9.704 GHz (B)). Stacked (C,D) presentation of the (++) quadrant of the WT (C) and D75H (D) spectra. Stacked presentation of the full (A) and (B) spectra is shown in Figure S4. Red dashed straight segments are defined by |ν1 ± ν2|= 2(15νN).
The 15N HYSCORE spectrum of the 15N-U D75H mutant (Fig. 2B,D) shows a pair of intensive cross-peaks 1 in the (+‒) quadrant with maxima at ~(±3.3, ‒/+0.4) MHz defining 15A=3.7 MHz or 14A=2.6 MHz, also consistent with the coupling estimated from 14N spectra. The (++) quadrant of the spectrum (Fig. 2B,D) exhibits also Nwc feature located around the (15νN,15νN) point. The Nwc feature in D75H has a different shape compared to that of the WT enzyme. For D75H, the Nwc feature is a triplet, including a central peak at (15νN,15νN) and two other lines symmetrically located around the antidiagonal with the splitting ~0.6 MHz.
The 14N and 15N ESEEM spectra show the presence of one strongly coupled nitrogen in the SQH environment, which is different in the WT and D75H cyt bo3 proteins. In contrast, the Nwc features resolved in 15N spectra of both the WT and D75H mutant result from multiple non-equivalent contributions of weakly-coupled nitrogen nuclei from in the immediate vicinity of the SQH. The shapes of the Nwc features indicate differences in the individual interactions for the SQH of the WT and D75H mutant. To further resolve the interactions with nitrogens in the SQH environment, selective 15N labeling in different residues, as well as 15N uniform labeling was employed.
Selective 15N labeling of the WT cyt bo3 protein
Arg, His and Gln were targeted for selective 15N labeling because the corresponding residues are involved in the current model of QH-site (Fig. 1). The molecular structure and atomic numbering of each of the three amino acids are displayed in Fig. S4. The following samples of WT cyt bo3 were prepared with 15N labels as follows. 1) Arg: a) uniform labeling; b) the two Nη positions; c) the peptide position Nα; 2) His: a) uniform labeling; b) ring-15N (Nδ and Nε); c) the Nδ position only; 3) Gln with 15N in the Nε position; 4) uniformly labeled with 15N except for Arg, Gln and His. It is assumed that only R71, Q101 and H98 are significant in interpreting the interactions with SQH.
The following results were obtained.
As shown previously, a dramatic change of the ESEEM spectra, accompanied by the complete disappearance of the 14N peaks, is observed with the WT cyt bo3 with uniformly 15N-labeled R71 (ref.19, Fig. S5A). The HYSCORE spectrum of the SQH in this protein contains two intense cross-peaks 1 similar to ones in the spectrum of 15N-U WT cyt bo3 (Fig. 2A,C). In contrast, the Nwc feature observed with 15N-U R71 WT cyt bo3 is very different (Fig. S5A), and resolves only a weak doublet centered around the (15νN, 15νN) diagonal point with the splitting 15A=0.15 MHz. A similar doublet was observed in the spectrum of the WT cyt bo3 with the selectively labeled 15Nη ositions in R71 (ref.19, Fig. S5B), whereas no 15N resolved peaks were observed in a sample in which the peptide nitrogen Nα of R71 was selectively labeled with 15N. These observations show a weak interaction in the WT cyt bo3 between the SQH with the 15Nη of R71, and confirm that the Nε of R71 possesses the largest hyperfine coupling and is responsible for the 14N spectral features in the WT cyt bo3.
It was previously shown that the spectrum of cyt bo3 with uniformly 15N-labeled H98 shows the Nwc feature with a maximum at the (15νN,15νN) point accompanied by extended shoulders up to 0.6 MHz with very poorly resolved maxima.19 This line can be produced by the interactions with up to three 15Ns and additional selective 15N labeling was performed to clarify the origins of this feature. The spectra of the sample with ring-15N (Nδ and Nε) labeled His is identical to that with uniformly 15N labeled His (Fig. S5C and Fig. S5D), but a very week antidiagonal 15N line is found in the spectrum of the protein with 15Nδ His. Thus, it is concluded that the extended 15Nwc feature is primarily due to the Nε of H98.
It is noted that a peak of very low intensity located at the diagonal point (15νN,15νN) was previously observed for the WT cyt bo3 with 15N labeled Nε of Q101, indicating very weak dipolar interaction between the unpaired electron and a distant 15N nucleus.19 No 15N signal was found in the spectra of WT cyt bo3 uniformly labeled with 15N except for Arg, Gln and His. Based on these observations, it is concluded that the only contributions to the Nwc feature are from R71, Q101 and H98.
Selective 15N labeling of D75H
Selectively 15N labeled D75H cyt bo3 samples were also examined. The 14N signals completely disappeared in the HYSCORE spectrum of the D75H with uniformly 15N(3)-labeled His (Fig. S6). This result is consistent with the prediction, based on the value of the qcc, that the 14N ESEEM spectrum of the D75H mutant arises from the protonated imidazole ring nitrogen of a histidine residue, presumably either H75 or H98.18 In order to definitively identify whether the Nδ or Nε of His contribute to the 14N ESEEM spectra, a D75H with 15Nδ His was prepared. The 14N HYSCORE spectrum of this sample is identical to the spectrum of unlabeled D75H in Fig. S2, so it can be concluded that these spectroscopic features originate from the Nε of a His residue.
The Nwc feature in the HYSCORE spectrum from the enzyme with uniformly labeled 15N(3) His possesses the maximum intensity at the (15νN,15νN) diagonal point and shoulders extended up to 0.5 MHz in both directions from the diagonal (Fig. 3B). However, the total intensity of the diagonal peak and shoulders is substantially lower than in the 15N-U enzyme (Fig. 3A), using the cross-peaks 1 as the reference for the intensity comparison. The larger intensity and different shape of the Nwc feature that is only observed in the uniformly labeled 15N-U D75H mutant, must be due to the additional contributions from nitrogens in R71 and Q101 residues.
Figure 3.
Nwc feature in 15N HYSCORE spectra of SQH in 15N -U D75H (A), D75H with uniformly 15N(3) labeled His (B), (C) obtained as a difference between the spectra of 15N -U D75H and 15N-U D75H except Arg, (D) obtained as a difference between the spectra of 15N-U D75H except Arg and D75H with uniformly 15N(3) labeled His.
Unfortunately, the HYSCORE spectra collected from the D75H selectively labeled with 15N Arg or 15N Gln do not resolve the signals from the weakly coupled 15N because any such signal is overshadowed by much more intense peak ν+ =1.61 MHz from the strongly coupled 14Nε of His (Fig. S7). In order to visualize the Nwc from R71 and Q101 in the D75H mutant, a reverse labeling approach was employed to prepare 15N-U D75H except for Arg. The contributions of the nitrogen interactions from R71 and Q101 to the 15N spectrum were then obtained using the spectra of (a) 15N-U D75H; (b) D75H with 15N-His(3); and (c) 15N-U D75H with 14N Arg. The subtractions a–c and c–b give the 15Nwc difference spectra from R71 and Q101, respectively (Fig. 3 C,D). The difference spectrum isolating the contribution from R71 shows extended shoulders with the maximum corresponding to the splitting ~0.6 MHz. The difference spectrum showing the contribution from Q101 consists of a weak peak at the diagonal point (15νN,15νN). In addition we have prepared the 15N-U D75H except for the Nε and Nα of Arg. The shape of Nwc for this sample is similar with one for D75H with 15N-His(3). This result indicates that the cross-peaks with the splitting 0.6 MHz are produced by the Nε of R71 that forms H-bond with SQH in D75H as well.
In contrast to the WT cyt bo3, there are two histidines, i.e. H75 and H98, around the SQH in the D75H mutant. Experimental 14,15N spectra do not give any indication which of them carries largest spin density on the Nε atom. However, in our previous work we have provided arguments based on the comparison of the hyperfine couplings with methyl protons reflecting asymmetry in spin density distribution with the QA site SQ of the reaction center that stronger interaction with the Nε of H75 is more preferable.18 Taking into account this assignment one can conclude that the Nwc pattern in D75H with 15N-His(3) includes contributions from Nδ of H75 and Nδ and Nε of H98. The Nwc shoulders in this sample (Fig. 3B) show slight increase of intensity corresponding to the splitting ~0.8±0.2 MHz. We suggested that the Nε from H98 forms the H-bond with the SQH similarly with the WT cyt bo3 and produces this splitting. The Nδ s in H75 and H98 separated by two bonds from H-bonded Nεs would have much smaller couplings and would contribute to the central part of the Nwc line around the diagonal point. This suggestion is supported by about 1/20 ratio of the hyperfine couplings for the remote and coordinated imidazole nitrogens in complexes with metals and clusters.32 We also suggested that Nαs from both His residues produce negligible influence on the spectra similarly as in the WT protein.
Hyperfine tensors of strongly coupled nitrogens
So far in our description of the experimental spectra we have used the hyperfine couplings determined from the position of cross-peak maxima using first-order expressions for two nuclear frequencies (Eqs.1, 2) of 15N nuclei. However, the lineshape of the cross-peaks from strongly coupled nitrogens in the WT and D75H cyt bo3 allows one to determine all of the principal values of hyperfine tensors. The cross-peaks from these nuclei possess the horn-like lineshape that indicates a significant rhombicity of the corresponding hyperfine tensors.25,26 The principal values of the rhombic hyperfine tensor can be defined as follow: Ax = aiso − T (1 + δ), Ay = aiso − T (1 − δ), Az = aiso + 2T with 0 ≤ δ ≤ 1, where aiso, T are the isotropic and anisotropic components of hyperfine coupling and δ is a rhombic parameter. The two nuclear frequencies of 15N (I = 1/2) from opposite mS = ±½ electron spin manifolds for each principal value i = x, y, z are ναi = |15νN + |Ai|/2| and νβi = |15νN − |Ai|/2|. An estimate of the principal components can be performed using theoretical predictions of the lineshape of the cross-peaks in powder-type spectra. The borders of the ideal cross-peak horn in such spectra are formed by three arc-type ridges between the pairs of three points (ναx, νβx), (ναy, νβy) and (ναz, νβz) located on the |ν1±ν2|=2(15νN) lines. The shape of these ridges is described by the general equation (where Q and G are coefficients which are functions of aiso,T, δ and 15νN)25,26:
| (3) |
The arc-type ridges transform to straight segments in ((ν1)2 vs. (ν2)2) plots, producing a triangle lineshape of the cross-peak with triangle vertexes at ((να(β)x)2, (νβ(α)x)2), ((να(β)y)2, (νβ(α)y)2), and ((να(β)z)2, (νβ(α)z)2).25 It should be noted that HYSCORE intensity at (ναx, νβx), (ναy, νβy) and (ναz, νβz) points corresponding to orientations of magnetic field along the principal directions (x, y, z) of the hyperfine tensor is equal to zero and significantly suppressed in the orientations around principal directions.26 Therefore, in HYSCORE spectra only the central part of the border cross-ridges, which correspond to orientations of the magnetic field substantially different from the principal directions, will possess substantial intensity.26 It means that in real spectrum the cross-peak borders should not cross the |ν1±ν2|=2(15νN) line(s) and the crossing points (ναx, νβx), (ναy, νβy) and (ναz, νβz) can be obtained through the linear regression of the observed parts of border arcs in ((ν1)2 vs. (ν2)2) presentation of the spectrum.
Fig. 4 (top) shows a presentation of the (++) quadrant of the 15N HYSCORE spectrum of the SQH in WT cyt bo3 in ((ν1)2 vs. (ν2)2) coordinates. The borders of the cross-peaks can be estimated by the area of the sharp increase of the peak intensity, i.e. where the blue background transformed into colored area in Fig. 4. In the spectrum shown in Fig. 4 (bottom), only one border segment with a well-defined linear portion can be defined (pink line segment). Linear regression of this segment (pink line in Fig. 4, bottom) gives two crossing points ((νβx)2, (ναx)2) and ((νβz)2, (ναz)2), corresponding to the minimal and maximal principal values. The larger coordinate of the crossing points is estimated by the values 5.5–5.8 MHz2 and 9.7–10.0 MHz2 that defines ναx =2.3–2.4 MHz or |Ax|=1.7±0.1 MHz and ναz=3.1–3.2 MHz or |Az|=3.3±0.1 MHz, respectively. The intermediate principal value was determined from the simulations of 15N HYSCORE spectra. Simulations varying the tensor anisotropy from axial (δ=1) to completely rhombic (δ=0) show that best agreement in the cross-peaks 1 location (Fig. S8) was achieved with δ=0.64±0.05, and defines the complete hyperfine tensor as |Az|=3.3 MHz, |Ay|=2.3 MHz |Ax|=1.7 MHz (±0.1MHz) with aiso=2.42 MHz and T=(0.88, −0.16, −0.72) MHz (signs a and T components are relative). All principal values Ai should have the same sign. Only this selection correctly describes the location of the cross-peak and gives the isotropic coupling consistent with the values estimated from 14N and 15N spectra. This tensor is assigned to the Nε of R71. A similar analysis was performed for cross-peaks 1 in the (+−) quadrant of the 15N HYSCORE spectrum of the SQH in the D75H mutant (see Figs. S9 and S10). Determined principal values of the hyperfine tensor assigned to the Nε of H75 are shown in Table 2.
Figure 4.
(top) Contour presentation of the (++) quadrant from the 15N HYSCORE spectrum of the SQH in 15N uniformly labeled WT cyt bo3 (Fig.2A) in ((ν1)2 vs. (ν2)2) coordinates. The red curve is defined by |ν1 ± ν2|= 2(15νN).
(bottom) Analysis of the cross-peak 1 contour lineshape. Linear regression of the border segment (pink line) gives two crossing points ((νβx)2, (ναx)2) and ((νβz)2, (ναz)2) with |ν1 ± ν2|= 2(15νN) curve corresponding to minimum and maximum principal values of hyperfine tensor.
Table 2.
Hyperfine tensors of 15N nuclei at the QH-site of cyt bo3 proteins.
| Cytochrome bo3 | Residue | Nitrogen | Hyperfine tensors (MHz) |
|---|---|---|---|
| Wild-type | R71 | Nε Nη Nα |
aiso=2.42, T=(0.88, −0.16, −0.72) aiso ~0.15, T<0.05–0.1 ~0 |
| H98 | Nδ Nε Nα |
~0 aiso=0.3, T~0.3–0.4 ~0 |
|
| Q101 | Nε | Weak dipolar coupling <0.05–0.1 MHz |
|
| D75H | H75 | Nε | aiso=3.5, T= (0.9, −0.2, −0.7) |
| H98 | Nε Nδ |
aiso=0.8, |T| ~ 0.3–0.4 | |
| R71 | Nε Nη |
aiso=0.6, |T| ~ 0.4–0.5 ~0 |
|
| Q101 | Nε | Weak dipolar coupling <0.05–0.1 MHz |
Hyperfine tensors of other nitrogens
In addition simulations of the 15N HYSCORE spectra were used to estimate the isotropic and anisotropic components of the hyperfine tensors for weakly coupled nitrogens, i.e. Nη of R71 and Nε of H98 in the WT cyt bo3 and Nε of H98 and Nε of R71 in the D75H cyt bo3. The results of simulations are provided in Table 2. The Nwc spectra are extended along antidiagonal of the (++) quadrant. They possess narrow linewidth in the direction normal to the antidiagonal and symmetrical line shapes with the cross-peak maximum corresponding to the undefined orientation of the magnetic field relative to the principal axes. The intensity is suppressed at the cross-peak wings corresponding to field orientations along or near the axes with maximum and minimum principal values of the tensor. Our previous analysis of the 15N HYSCORE spectra23 and simulations have shown that the position of the maximum gives accurate estimate of the isotropic coupling (~±0.05 MHz) but the relative signs of the isotropic and anisotropic components and symmetry of the tensor (i.e., axial or rhombic) are uncertain from the NWC lineshape. This uncertainty influences the accuracy in the anisotropic tensor estimate in larger degree, i.e. for T~0.3–0.4 MHz, and the accuracy in its determination is ~0.1–0.15 MHz.
Density Functional Studies
Previous modeling studies used water molecules and N-methylformamide groups as hydrogen bond donors to the O1 and O4 atoms of the SQH and did not take into account the varying strength of hydrogen bond interaction with the SQH by different groups.33–35 In our work, based on the X-ray structure, Fig. 1, and the EPR and mutational data for the WT enzyme, the SQH was modeled to have hydrogen bonds to the OH of the carboxylic acid group of D75, the NεH of the imidazole group of H98 and the NεH group of the guanidium group of R71. In the absence of accurate data from the X-ray crystal structure, we explored idealized small models with geometry optimization. Therefore, the overall significance and relative strength of each interaction can be assessed by the spin density distribution, but the correct orientation of the hydrogen bonds with the SQH will not be well reproduced.
Computed Geometries
Table 3 shows the calculated geometry of the isolated SQH and the effect of hydrogen bonding on this geometry in the WT and D75H models. For the WT model, the hydrogen bond lengths of O1 with the R71 NεH group (1.65 Ǻ) and the D75 carboxylic acid group (1.60 Ǻ) are considerably shorter than those formed at O4 with the imidazole NH representing H98 (1.83 Ǻ). These bond distance trends suggest stronger hydrogen bonding to the O1 atom of the SQH caused principally by the strong interaction with the carboxylic acid group and the positively charged guanidium group. The NH group of the H98 imidazole is a weaker hydrogen bond donor. For the D75H model, the hydrogen bond length to the imidazole NH group from H75 is also significantly longer (1.79 Ǻ) than that found for the carboxylic acid group of D75 in the WT model (1.60 Ǻ) suggesting a weaker hydrogen bond formed by the imidazole group in the mutant.
Table 3.
Optimised geometries. All distances in angstroms.
| Isolated | WT | D75H | |
|---|---|---|---|
| C1-O1 | 1.27 | 1.30 | 1.30 |
| C4-O4 | 1.27 | 1.27 | 1.27 |
| O1-H (R71) | - | 1.65 | 1.61 |
| O1-H (D75/H75) | - | 1.60 | 1.79 |
| O4-H (H98) | - | 1.83 | 1.82 |
Spin densities and populations
Fig. 5 shows the spin density distribution changes that occur on going from the isolated SQ to the WT and D75H models. Fig. 6 provides a more quantitative picture using spin populations obtained from a Mulliken population analysis. The spin density distribution of the isolated SQ is symmetric whereas, in contrast, it is highly asymmetric for both the WT and D75H. For simpler semiquinone models, the primary effect of hydrogen bonding to either of the semiquinone carbonyl oxygen atoms is a redistribution of spin density from the oxygen atom position to the adjacent carbon.36 In our WT model (Fig. 6), the spin density redistribution is mainly from O1 to C1 with a much smaller redistribution occurring between C4 and O4. The charged guanidinum and carboxylic acid groups polarize the C1-O1 bond. The increased spin population at C1 leads via spin polarization to a lower spin population at positions C2 and C6 and a higher spin population at positions C3 and C5. This “domino” spin polarization effect should lead to a significantly lower spin population at C4, but this is offset by the presence of a hydrogen bond from the imidazole group of H98 to the O4 atom, which partially balances the spin polarization effect on C4. The spin populations obtained for WT can, therefore, be explained by the presence of strong hydrogen bonding to the O1 atom of the SQH and a significantly weaker hydrogen bond to the O4 atom, in accord with EPR studies. For the D75H model, the spin populations are slightly less asymmetric compared to the WT model. As noted above the hydrogen bond made by the imidazole is longer than that calculated for the carboxylic acid group suggesting a weaker hydrogen bond is formed by this group. The predicted spin density distributions are also in accord with the lower 5-CH3 1H and 13C hyperfine couplings measured for the D75H mutant, as discussed below.
Figure 5.
Spin density contour plots (0.004e/au) for (a) Isolated, (b) WT and (c) D75H models. Atom numbering is as shown in Fig. 1.
Figure 6.
Mulliken spin populations for (a) Isolated, (b) WT and (c) D75H models. Atom numbering is as shown in Fig. 1.
Comparison of calculated and experimental hyperfine couplings
In addition to the spin density distribution, hyperfine couplings can be calculated and compared with available 15N, 1H and 13C experimental values. Table 4 shows the 15N isotropic and anisotropic hyperfine couplings calculated for the WT and D75H models. In addition, the calculated 14N qcc, K, and asymmetry parameters, η, are compared with experimental values. The experimental 15N isotropic hyperfine couplings for the R71 Nε and Nη (2.4 MHz and 0.15 MHz, respectively), as well as K (0.93 MHz) and η (0.51) for the R71 Nε are excellently reproduced by the WT model. For the H98 imidazole NH, the calculations give aiso = 0.4 MHz, in reasonable agreement with the experimental value of 0.3 MHz. The value of the isotropic hyperfine couplings for these nitrogens is very sensitive to the angle made by the NH donor with respect to the SQ ring plane.37 The very close correspondence between calculated and experimental value for the WT model suggests that the orientation of these hydrogen bond donors in the optimized model is very similar to that adopted in the actual QH binding site.
Table 4.
15N isotropic, aiso, and anisotropic, Tii, hyperfine tensors calculated for the QH site models. Calculated values for the 14N Nuclear Quadrupole Coupling Constant (K) and the asymmetry parameter, η are also given. Experimental values are given in brackets. All values given in MHz.
| WT Model | D75H | |||||||
|---|---|---|---|---|---|---|---|---|
| Pos. | TzzTyyTxx | aiso | K | η | TzzTyyTxx | aiso | K | η |
| R71 - Nε | 0.6(0.9) −0.3(−0.6) −0.3(0.3) |
2.4 (2.4) | 0.9 (0.9) | 0.7 (0.5) | 0.0 0.0 0.0 |
1.4, 0.5* (0.6) |
0.9 | 0.8 |
| R71 - Nη | 0.0 0.0 0.0 |
0.2 (0.15) | 1.1 | 0.4 | 0.0 0.0 0.0 |
0.0 (0.0) | 1.1 | 0.5 |
| R71 - Nη | 0.0 0.0 0.0 |
0.0 | 1.2 | 0.2 | 0.0 0.0 0.0 |
0.0 (0.0) | 1.2 | 0.3 |
| H98 - N | 0.4 −0.1 −0.3 |
0.4 (0.3) | 0.6 | 0.5 | 0.5 −0.2 −0.3 |
0.6 (0.8) | 0.6 | 0.6 |
| H75 - N | - | - | - | - | 0.4, 0.7*(0.9) −0.2,−0.4*(−0.7) −0.2,−0.3*(−0.2) |
0.8, 2.5* (3.5) |
0.5(0.4) | 0.6 (0.7) |
Values using adjusted orientation, see text for details.
For the D75H model, the optimized geometry gives rise to 15N isotropic hyperfine couplings which deviate somewhat from the experimental determinations. The R71 Nε has a calculated value of 1.4 MHz whereas the experimental assignment is 0.6 MHz. The calculated value for the NεH group of H75 is 0.8 MHz whereas the experimental assignment is 3.5 MHz. For the H98 imidazole group, the calculated value, 0.6 MHz, is in good agreement with the experimental assignment of 0.8 MHz. In the optimized D75H model, the imidazole NH group is oriented 42° out of the SQ ring plane. The sensitivity of the 15N value to this orientation is demonstrated by changing the orientation of the H75 NH group from the optimized value of 42° to 90°. This changes the calculated 15N isotropic hyperfine coupling from 0.8 to 2.5 MHz, much closer to the experimentally measured 3.5 MHz. Likewise changing the out of plane orientation of the guanidium group NεH from the optimized value of 41° to 35 ° changes the 15N isotropic coupling from 1.4 MHz to 0.6 MHz in exact agreement with the experimental assignment. For the H75 imidazole NH group the calculated 14N qcc K has a value of 0.5 MHz and the asymmetry parameter, η is calculated to be 0.6. These are in excellent agreement with our experimental determinations of 0.4 MHz and 0.7, respectively. Our conclusion, therefore, is that the D75H model is a good representation of the SQH in the mutant but the orientation of the hydrogen bond donors is different from the optimized small model calculation.
Calculated hyperfine couplings can also be compared with previous experimental determinations of 1H hyperfine couplings for the WT and D75H mutant. Rotating methyl groups are observed readily using ENDOR spectroscopy as they give a strong ENDOR response and have been used as a primary marker in gauging the spin density asymmetry within SQs in photosynthetic reaction centers.38 The 5-CH3 1H hyperfine coupling has been measured in numerous studies and indeed is the principal indicator that the spin density of the WT SQH is highly asymmetric compared with the radical generated in vitro. For SQ in a non-polar solvent, and, therefore, not involved in hydrogen bonding, aiso is 6.0 MHz.38 Our calculated value for the isolated model is 5.9 MHz, which is in excellent quantitative agreement with this determination. For our WT model of SQH, the calculated 1H aiso value is 9.2 MHz, which is in good accord with experimental studies which vary from 9.5–10.0 MHz.34 The elevated hyperfine coupling for the 5-CH3 position is a result of the elevated spin population calculated for the ring C5 position shown in Fig. 6. Previous DFT studies33–35 using SQ anion models have been unable to match the experimental value for this position and this can be mainly attributed to the use of water or amide groups as the hydrogen bonding groups in these modeling studies. Because the elevated 5-CH3 hyperfine coupling was similar to that found for neutral semiquinone free radicals, it has been proposed that the QH site SQ is a neutral radical.17,18 No evidence for formation of the neutral SQ form has been obtained with the current computational model although this cannot be ruled out with a different larger model of the QH site. The asymmetry in spin density distribution for both scenarios is very similar. Our current computational studies indicate that the elevated 5-CH3 coupling arises from strong hydrogen bonding of a semiquione anion radical to a positively charged guanidium group and a carboxylic acid group, i.e. R71 and D75 in the WT SQH.
For the D75H model, the calculated 5-CH3 isotropic hyperfine coupling of 8.2 MHz is lower than that of the WT model and is in excellent agreement with the value found experimentally for the D75H mutant of 8.0 MHz.18 The lower value for this hyperfine coupling, compared to the WT, can be ascribed to the more symmetric spin density distribution for the D75H model compared with the WT (Fig. 5 and 6). Replacement of the carboxylic acid group of D75 with the imidazole group of H75 results in significantly weaker hydrogen bonding to the O1 atom of SQH. The more symmetric spin density distribution leads to a lower spin density at C5 and a lower 1H hyperfine coupling value for the 5-CH3 group compared with the WT model. The good reproduction of this well characterized coupling suggests that the spin density distribution of the D75H model is reasonably accurate.
More recently, 2D ESEEM studies have obtained 13C isotropic hyperfine coupling values for the 5-CH3 carbon nucleus for both WT and D75H samples.20 For the WT and D75H models, the calculated isotropic values −5.3 MHz and −4.4 MHz are in good agreement with the experimental values of −6.1 MHz and −4.7 MHz, respectively.
1H hyperfine couplings have also been resolved from exchangeable, presumably hydrogen bonded, protons in both the WT and D75H mutant. Compared with methyl group hyperfine couplings, hydrogen bonded protons are difficult to interpret in complex systems where there can be significant overlap of spectral lines. For the WT SQH, 2D ESEEM identified three protons having isotropic and anisotropic values (aiso, T) of (−/+ 0.7, ± 6.3) MHz; (−/+ 1.2, ±4.2) MHz; (−/+ 4.2, ±1.7) MHz.17 The calculated values of (aiso, T=T11(max)/2) for the three hydrogen bonded protons from D75, R71 and H98 are, respectively: (0.0, 4.5) MHz, (−0.1, 4.3) MHz and (−0.6, 4.1) MHz (Table 5). The good agreement between the calculated R71Nε, 15N aiso, value and experimental value indicates that the NH orientation relative to the SQH ring plane is accurately modeled for the WT. The most likely candidate for the strong H-bond observed experimentally is the COOH group of D75, which exhibits the shortest optimized hydrogen bonding distance. In the optimized model of the WT, this is oriented 40° out of the ring plane. By changing this orientation to 90° we calculate that the anisotropic coupling increases to 5.4 MHz with an isotropic coupling value of 0.8 MHz, which is in better agreement with the experimental values. Therefore, we assign the largest (−0.7, +6.3) MHz proton hyperfine coupling to the COOH proton of D75. The (−1.2, +4.2) MHz coupling could be due to an overlap of the H98 NH and R71 NεH lines. None of our calculated values match the very unusual coupling of (−/+ 4.2, ±1.7) MHz. The unusually large isotropic coupling together with the relatively small anisotropic component would suggest that this coupling does not arise from a hydrogen bonded proton, and further experimental characterization is required.
Table 5.
5-Methyl and Hydrogen bonded 1H isotropic, aiso, and anisotropic, Tii, hyperfine tensors calculated for the QH site model. Experimental values are given in brackets. All values are given in MHz.
| Isolated | WT | D75H | ||||
|---|---|---|---|---|---|---|
| Pos. | TzzTyyTxx | aiso | TzzTyyTxx | aiso | TzzTyyTxx | aiso |
| 5- CH3 | 2.6 −1.3 −1.3 |
5.9 (6.0) | 3.0 −1.5 −1.5 |
9.2 (9.5–10.0) | 2.7 −1.4 −1.4 |
8.2 (8.0) |
| R71 - NHε | - | - | 8.6 −4.1 −4.5 |
−0.1 | 8.4 −4.1 −4.3 |
−0.2 |
| R71 - NHη | - | - | 1.8 −0.6 −1.2 |
0.0 | 1.8 −0.6 −1.2 |
0.0 |
| R71 - NHη | - | - | 0.8 −0.3 −0.5 |
0.0 | 0.8 −0.3 −0.5 |
0.0 |
| H98 - NH | - | - | 8.2 −3.9 −4.2 |
−0.6 | 8.3 −4.0 −4.3 |
−0.7 |
| D75/H75 – COOH/NH | - | - | 8.9 −4.4 −4.5 |
0.0 | 6.5 −3.2 −3.3 |
−0.4 |
For the D75H mutant exchangeable protons have been experimentally assigned to 1H hyperfine couplings (aiso, T) of (−/+ 1.0, ±4.6) MHz and (−/+ 4.3, ±1.2) MHz. Three hydrogen bond donors are present in our D75H model: R71 Nε 1H, H75 N1H and H98 N1H (Table 5). The calculated (aiso, T) values of each of these are in reasonable agreement with the (1.0 MHz, +4.6 MHz) experimental value, so all three may contribute to this experimentally determined coupling. As found with the WT model, the other experimentally observed coupling with a large isotropic and small anisotropic component, (−/+ 4.3, ±1.2) MHz, has no counterpart in the calculated values and is unlikely to be due to a proton(s) hydrogen bonded with carbonyls. One can suggest that exchangeable resonances with weaker hfi couplings observed in the spectra result from the overlap of the signals from the several protons in the SQ environment. Overlap of the cross-peak with different (but weak) hyperfine couplings would give cross-peaks with the shape and length substantially different from contributing signals. As the result the analysis would give some “effective” couplings not related directly to the real values. We suggest that further experiments using fully deuterated protein and cofactor as well as HYSCORE and ENDOR experiments at Q-band will better resolve these couplings, which would identify their source. Q-band will also address specific questions about the orientation of the H-bonds around the SQH site based on the hyperfine (and nuclear quadrupole) tensors of exchangeable protons (deuterons) that would lead to more adequate computational models.39
Relevance to quinone one-electron transfer role
For in vitro systems, in protic solvents, quinone reduction or quinol oxidation proceeds via sequential loss or gain of two electrons via the semiquinone free radical form with consequent loss or gain of two protons. In biological systems, quinones can function additionally as one-electron transfer agents. For example the QA-site quinone in Type II photosynthetic reaction centers acts as a one-electron transfer agent between a pheophytin molecule and the quinone reductase site QB.28,38 In the Photostsystem I reaction center, a phylloquinone molecule, A1, also acts as a one-electron transfer agent between a chlorophyll molecule and an Fe4S4 center.40,41 For QB, which undergoes two electron reduction and subsequent protonation to the quinol, as in in vitro systems, a more symmetric spin density distribution is found.38 The QB quinone is in dynamic equilibrium with a quinone pool and binds to receive two electrons and two protons before unbinding in the reduced quinol form. It is noteworthy that for the quinones functioning as one-electron transfer agents, an extremely asymmetric spin density distribution has been found for the SQ form, whereas in a true substrate binding quinone site, such as QB, a more symmetric spin density has been found i.e. similar to that found in protic solvents. To fulfill the role of a one-electron transfer agent it may be a requirement of the quinone binding site to produce a highly asymmetric spin density distribution for the SQ intermediate. The asymmetry in the spin density distribution for the semiquinone reflects a more contracted/localized electron density distribution for the singly occupied molecular orbital, SOMO, of the free radical anion. This is clearly shown by the spin density contours in Fig. 5. This contracted electron density distribution would be expected to lead to a less stable SQ form compared with the symmetrical form but would also imply that further one electron reduction to the two electron reduced dianion or quinol form would be unfavourable as now two electrons have to occupy the contracted asymmetric orbital. While the exact mechanism of electron transfer and reduction at the QH site is still unknown, it is believed to shuttle electrons from a true quinone substrate binding site QL to heme b for eventual reduction of oxygen at the heme o3-CuB active site.7,13 It would appear to perform a similar function as the CuA site in cytochrome c oxidase. As for the QA and A1 quinones described above, the high asymmetry in spin density distribution observed for SQH, would indicate that it performs a similar one-electron transfer role and that high asymmetry in the SOMO electron density distribution may be a crucial factor in determining the quinone role as an effective one-electron transfer agent. In the present study therefore one possibility for the non-activity exhibited by the D75H mutant could be related to the more symmetrical spin density distribution of the QH site SQ compared with the WT. The asymmetry in spin density distribution is still quite large however for the D75H mutant, compared to the symmetrical situation. This would argue against the above hypothesis, unless rapid one-electron transfer requires the extreme asymmetry which is present in the WT sample and also present in the QA and A1 sites.
Conclusions
In this study, the pulsed EPR experiments performed with the selectively 15N labeled cyt bo3 samples have led to the unambiguous assignment of the Nε of R71 as the nitrogen that is strongly coupled to the QH-site SQ in the wild-type enzyme and gives rise to the observed 14N ESEEM signals. In addition, selective 15N labeling of cyt bo3 has enabled the detection of even weak interactions between the SQH and individual residues at the ubiquinone binding site as summarized in Table 2. The HYSCORE data suggest that H98 at the QH-site is weakly coupled to the semiquinone radical and there is no direct interaction between the SQH and the Nε of Q101. The 15N selective labeling has also identified the Nε of H75 as the nitrogen hydrogen bonded with the SQH of the D75H mutant enzyme.
Density functional calculations on models of the active site for both the WT and D75H systems show good agreement between for experimentally observed and calculated values of available 14,15N, 1H and 13C couplings. The model studies indicate that a very strong hydrogen bonding interaction occurs between the O1 atom of the SQH and the hydrogen bond donor groups of R71 and D75 with a relatively weaker interaction occurring for O4 and the imidazole NH group of H98. This is mainly responsible for the highly asymmetric spin density distribution observed experimentally. Replacement of D75 by H75 in the D75H mutant model leads to a lower spin density asymmetry for the SQH. The HYSCORE results also imply that in the D75H mutant, the Nεof R71 possesses significantly smaller hyperfine coupling than that in the WT protein despite practically the same length of H-O bond in optimized structure. This suggests that the orientation of the hydrogen bond donors is different from the optimized small model calculation for D75H. Based on previous studies of quinone one-electron transfer sites, we further postulate that the highly asymmetric spin density distribution of the WT SQH may be a significant factor in its role as a one-electron transfer agent between the substrate binding site and heme b.
Overall, the combined pulsed EPR experiments and electronic structure calculations carried out in this study constitute a major step towards complete characterization of the distribution of the unpaired spin density of the SQ in the QH site. The interactions between the SQ and nearby residues unraveled in this study are crucial to understand how the radical is stabilized inside the protein’s binding pocket and establish a foundation for future studies of quinone structure-function relationships in bioenergetics.
Supplementary Material
Acknowledgment
PJOM acknowledges the use of the EPSRC UK National Service for Computational Chemistry Software (NSCCS) in carrying out this work.
Abbreviations
- 2D
two-dimensional
- CW EPR
continuous wave electron paramagnetic resonance
- cyt bo3
cytochrome bo3 ubiquinol oxidase from E. coli
- DDM
n-dodecyl-β-D-maltoside
- DFT
Density Functional Theory
- ESEEM
electron spin echo envelope modulation
- HYSCORE
hyperfine sublevel correlation
- nqi
nuclear quadrupole interactions
- Nwc
weakly coupled nitrogens
- qcc
quadrupole coupling constant
- QH
the high affinity quinone-binding site
- QL
the low affinity quinone-binding site
- SQ
semiquinone
- SQH
semiquinone at the QH site
- UQ8
ubiquinone-8
- WT
wild-type
Footnotes
This investigation was supported by the DE-FG02-08ER15960 (S.A.D.) and DE-FG02-87ER13716 (R.B.G.) Grants from Chemical Sciences, Geosciences and Biosciences Division, Office of Basic Energy Sciences, Office of Sciences, US DOE, and the NIH GM062954 Grant (S.A.D) and NCRR/NIH Grants S10-RR015878 and S10-RR025438 for instrumentation.
Supporting Information Available: model of the QH-site, 14N,15N three-pulse and HYSCORE spectra of the QH site SQ in WT and D75H cyt bo3, the structures of amino acids selectively labeled in this work showing the atomic numberings, complete ref. 27. This material is available free of charge via the Internet at http://pubs.acs.org.
REFERENCES
- 1.Anraku Y. Bacterial electron transport chains. Annu. Rev. Biochem. 1988;57:101–132. doi: 10.1146/annurev.bi.57.070188.000533. [DOI] [PubMed] [Google Scholar]
- 2.Puustinen A, Finel M, Virkki M, Wikstrom M. Cytochrome o (bo) is a proton pump in Paracoccus denitrificans and Escherichia coli. FEBS Lett. 1989;249:163–167. doi: 10.1016/0014-5793(89)80616-7. [DOI] [PubMed] [Google Scholar]
- 3.Puustinen A, Finel M, Haltia T, Gennis RB, Wikstrom M. Properties of the two terminal oxidases of Escherichia coli. Biochemistry. 1991;30:3936–3942. doi: 10.1021/bi00230a019. [DOI] [PubMed] [Google Scholar]
- 4.Garcia-Horsman JA, Barquera B, Rumbley J, Ma J, Gennis RB. The superfamily of heme-copper respiratory oxidases. J. Bacteriol. 1994;176:5587–5600. doi: 10.1128/jb.176.18.5587-5600.1994. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 5.Trumpower BL, Gennis RB. Energy transduction by cytochrome complexes in mitochondrial and bacterial respiration: The enzymology of coupling electron transfer reactions to transmembrane proton translocation. Annu. Rev. Biochem. 1994;63:675–716. doi: 10.1146/annurev.bi.63.070194.003331. [DOI] [PubMed] [Google Scholar]
- 6.Puustinen A, Verkhovsky MI, Morgan JE, Belevich NP, Wikstrom M. Reaction of the Escherichia coli quinol oxidase cytochrome bo3 with dioxygen: The Role of a bound ubiquinone molecule. Proc. Natl. Acad. Sci. U. S. A. 1996;93:1545–1548. doi: 10.1073/pnas.93.4.1545. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 7.Yap LL, Lin MT, Ouyang H, Samoilova RI, Dikanov SA, Gennis RB. The quinone-binding sites of the cytochrome bo3 ubiquinol oxidase from Escherichia coli. Biochim. Biophys. Acta. 2010;1797:1924–1932. doi: 10.1016/j.bbabio.2010.04.011. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 8.Sato-Watanabe M, Mogi T, Ogura T, Kitagawa T, Miyoshi H, Iwamura H, Anraku Y. Identification of a novel quinone-binding site in the cytochrome bo complex from Escherichia coli. J. Biol. Chem. 1994;269:28908–28912. [PubMed] [Google Scholar]
- 9.Sato-Watanabe M, Itoh S, Mogi T, Matsuura K, Miyoshi H, Anraku Y. Stabilization of a semiquinone radical at the high-affinity quinone-binding site (QH) of the Escherichia coli bo-type ubiquinol oxidase. FEBS Lett. 1995;374:265–269. doi: 10.1016/0014-5793(95)01125-x. [DOI] [PubMed] [Google Scholar]
- 10.Ingledew WJ, Ohnishi T, Salerno JC. Studies on a stabilisation of ubisemiquinone by Escherichia coli quinol oxidase, cytochrome bo. Eur. J. Biochem. 1995;227:903–908. doi: 10.1111/j.1432-1033.1995.tb20217.x. [DOI] [PubMed] [Google Scholar]
- 11.Sato-Watanabe M, Mogi T, Miyoshi H, Anraku Y. Characterization and functional role of the QH site of bo-type ubiquinol oxidase from Escherichia coli. Biochemistry. 1998;37:5356–5361. doi: 10.1021/bi9727592. [DOI] [PubMed] [Google Scholar]
- 12.Mogi T, Sato-Watanabe M, Miyoshi H, Orii Y. Role of a bound ubiquinone on reactions of the Escherichia coli cytochrome bo with ubiquinol and dioxygen. FEBS Lett. 1999;457:223–226. doi: 10.1016/s0014-5793(99)01047-9. [DOI] [PubMed] [Google Scholar]
- 13.Kobayashi K, Tagawa S, Mogi T. Transient formation of ubisemiquinone radical and subsequent electron transfer process in the Escherichia coli cytochrome bo. Biochemistry. 2000;39:15620–15625. doi: 10.1021/bi0014094. [DOI] [PubMed] [Google Scholar]
- 14.Abramson J, Riistama S, Larsson G, Jasaitis A, Svensson-Ek M, Laakkonen L, Puustinen A, Iwata S, Wikstrom M. The structure of the ubiquinol oxidase from Escherichia coli and its ubiquinone binding site. Nat. Struct. Biol. 2000;7:910–917. doi: 10.1038/82824. [DOI] [PubMed] [Google Scholar]
- 15.Hellwig P, Yano T, Ohnishi T, Gennis RB. Identification of the residues involved in stabilization of the semiquinone radical in the high-affinity ubiquinone binding site in cytochrome bo3 from Escherichia coli by site-directed mutagenesis and EPR spectroscopy. Biochemistry. 2002;41:10675–10679. doi: 10.1021/bi012146w. [DOI] [PubMed] [Google Scholar]
- 16.Grimaldi S, MacMillan F, Ostermann T, Ludwig B, Michel H, Prisner T. QH*− ubisemiquinone radical in the bo3-type ubiquinol oxidase studied by pulsed electron paramagnetic resonance and hyperfine sublevel correlation spectroscopy. Biochemistry. 2001;40:1037–1043. doi: 10.1021/bi001641+. [DOI] [PubMed] [Google Scholar]
- 17.Yap LL, Samoilova RI, Gennis RB, Dikanov SA. Characterization of the exchangeable protons in the immediate vicinity of the semiquinone radical at the QH site of the cytochrome bo3 from Escherichia coli. J. Biol. Chem. 2006;281:16879–16887. doi: 10.1074/jbc.M602544200. [DOI] [PubMed] [Google Scholar]
- 18.Yap LL, Samoilova RI, Gennis RB, Dikanov SA. Characterization of mutants that change the hydrogen bonding of the semiquinone radical at the QH site of the cytochrome bo3 from Escherichia coli. J. Biol. Chem. 2007;282:8777–8785. doi: 10.1074/jbc.M611595200. [DOI] [PubMed] [Google Scholar]
- 19.Lin MT, Samoilova RI, Gennis RB, Dikanov SA. Identification of the nitrogen donor hydrogen bonded with the semiquinone at the QH site of the cytochrome bo3 from Escherichia coli. J. Am. Chem. Soc. 2008;130:15768–15769. doi: 10.1021/ja805906a. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 20.Lin MT, Shubin AA, Samoilova RI, Narasimhulu KV, Baldansuren A, Gennis RB, Dikanov SA. Exploring by pulsed EPR the electronic structure of ubisemiquinone bound at the QH site of cytochrome bo3 from Escherichia coli with in vivo 13C-labeled methyl and methoxy substituents. J. Biol. Chem. 2011;286:10105–10114. doi: 10.1074/jbc.M110.206821. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 21.Datsenko KA, Wanner BL. One-step inactivation of chromosomal genes in Escherichia coli K-12 using PCR products. Proc. Natl. Acad. Sci. U. S. A. 2000;97:6640–6645. doi: 10.1073/pnas.120163297. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22.Epel B, Silakov A. KAZAN viewer for Matlab™: Visualisation and data processing GUI for EPR and NMR. http://www.boep.specman4epr.com/kv_intro.html. [Google Scholar]
- 23.Martin E, Samoilova RI, Narasimhulu KV, Wraight CA, Dikanov SA. Hydrogen bonds between nitrogen donors and the semiquinone in the QB site of bacterial reaction centers. J. Am. Chem. Soc. 2010;132:11671–11677. doi: 10.1021/ja104134e. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 24.Höfer P, Grupp A, Nebenführ H, Mehring M. Hyperfine sublevel correlation (hyscore) spectroscopy: a 2D ESR investigation of the squaric acid radical. Chem. Phys. Lett. 1986;132:279–282. [Google Scholar]
- 25.Dikanov SA, Bowman MK. Cross-peak lineshape of two-dimensional ESEEM spectra in disordered S = 1/2, I = 1/2 spin systems. J. Magn. Reson. A. 1995;116:125–128. [Google Scholar]
- 26.Dikanov SA, Tyryshkin AM, Bowman MK. Intensity of cross-peaks in HYSCORE spectra of S = 1/2, I = 1/2 spin systems. J. Magn. Reson. 2000;144:228–242. doi: 10.1006/jmre.2000.2055. [DOI] [PubMed] [Google Scholar]
- 27.Frisch MJ, et al. Wallingford CT: Gaussian, Inc.; 2009. [Google Scholar]
- 28.Lin T-J, O’Malley PJ. An ONIOM study of the spin density distribution of the QA site plastosemiquinone in the photosystem II reaction center. J. Phys. Chem. B. 2011;115:4227–4233. doi: 10.1021/jp112163w. [DOI] [PubMed] [Google Scholar]
- 29.Martin E, Samoilova RI, Narasimhulu KV, Lin TJ, O'Malley PJ, Wraight CA, Dikanov SA. Hydrogen bonding and spin density distribution in the QB semiquinone of bacterial reaction centers and comparison with the QA site. J. Am. Chem. Soc. 2011;133:5525–5537. doi: 10.1021/ja2001538. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 30.Dikanov SA, Samoilova RI, Kappl R, Crofts AR, Hüttermann J. The reduced [2Fe-2S] clusters in adrenodoxin and Arthrospira platensis ferredoxin share spin density with protein nitrogens, probed using 2D ESEEM. Phys. Chem. Chem. Phys. 2009;11:6807–6819. doi: 10.1039/b904597j. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 31.Dikanov SA, Shubin AA, Kounosu A, Iwasaki T, Samoilova RI. A comparative, two-dimensional 14N ESEEM characterization of reduced [2Fe-2S] clusters in hyperthermophilic archaeal high- and low-potential Rieske-type proteins. J Biol. Inorg. Chem. 2004;9:753–767. doi: 10.1007/s00775-004-0571-y. [DOI] [PubMed] [Google Scholar]
- 32.Iwasaki T, Kounosu A, Uzawa T, Samoilova RI, Dikanov SA. Orientation-selected 15N-HYSCORE detection of weakly coupled nitrogens around the archaeal Rieske [2Fe−2S] center. J. Am. Chem. Soc. 2004;126:13902–13903. doi: 10.1021/ja045898x. [DOI] [PubMed] [Google Scholar]
- 33.Boesch SE, Wheeler RA. Isotropic 13C hyperfine coupling constants distinguish neutral from anionic ubiquinone-derived radicals. Chem. Phys. Chem. 2009;10:3187–3189. doi: 10.1002/cphc.200900503. [DOI] [PubMed] [Google Scholar]
- 34.Kacprzak S, Kaupp M, MacMillan F. Protein–cofactor interactions and EPR parameters for the QH quinone binding site of quinol oxidase. A density functional study. J. Am. Chem. Soc. 2006;128:5659–5671. doi: 10.1021/ja053988b. [DOI] [PubMed] [Google Scholar]
- 35.MacMillan F, Kacprzak S, Hellwig P, Grimaldi S, Michel H, Kaupp M. Elucidating mechanisms in haem copper oxidases: The high-affinity QH binding site in quinol oxidase as studied by DONUT-HYSCORE spectroscopy and density functional theory. Faraday Discuss. 2011;148:315–344. doi: 10.1039/c005149g. [DOI] [PubMed] [Google Scholar]
- 36.O'Malley PJ. Effect of hydrogen bonding on the spin density distribution and hyperfine couplings of the p-benzosemiquinone anion radical in alcohol solvents: A hybrid density functional study. J. Phys. Chem. A. 1997;101:9813–9817. [Google Scholar]
- 37.O'Malley PJ. A density functional study of the effect of orientation of hydrogen bond donation on the hyperfine couplings of benzosemiquinones: relevance to semiquinone-protein hydrogen bonding interactions in vivo. Chem. Phys. Lett. 1998;291:367–374. [Google Scholar]
- 38.Lubitz W, Feher G. The primary and secondary accepters in bacterial photosynthesis III. Characterization of the quinone radicals QA−˙ and QB−˙ by EPR and ENDOR. Appl. Magn. Reson. 1999;17:1–48. [Google Scholar]
- 39.Flores M, Isaacson R, Abresch E, Calvo R, Lubitz W, Feher G. Protein-cofactor interactions in bacterial reaction centers from Rhodobacter sphaeroides R-26: II. Geometry of the hydrogen bonds to the primary quinone Q˙A− by 1H and 2H ENDOR spectroscopy. Biophys. J. 2007;92:671–682. doi: 10.1529/biophysj.106.092460. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 40.Lin T-J, O’Malley PJ. Binding site influence on the electronic structure and electron paramagnetic resonance properties of the phyllosemiquinone free radical of photosystem I. J. Phys. Chem. B. 2011;115:9311–9319. doi: 10.1021/jp203484w. [DOI] [PubMed] [Google Scholar]
- 41.Srinivasan N, Golbeck JH. Protein-cofactor interactions in bioenergetic complexes: The role of the A1A and A1B phylloquinones in Photosystem I. Biochim. Biophys. Acta. 2009;1787:1057–1088. doi: 10.1016/j.bbabio.2009.04.010. [DOI] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.






