Abstract
Visible and ultraviolet resonance Raman (RR) spectra are reported for FeIII(NO) adducts of myoglobin variants with altered polarity in the distal heme pockets. The stretching frequencies of the FeIII–NO and N–O bonds, νFeN and νNO, are negatively correlated, consistent with backbonding. However, the correlation shifts to lower νNO for variants lacking a distal histidine. DFT modeling reproduces the shifted correlations, and shows the shift to be associated with the loss of a lone-pair donor interaction from the distal histidine that selectively strengthens the N–O bond. However, when the model contains strongly electron-withdrawing substituents at the heme β-positions, νFeN and νNO become positively correlated. This effect results from FeIII–N–O bending, which is induced by lone pair donation to the NNO atom. Other mechanisms for bending are discussed, which likewise lead to a positive νFeN/νNO correlation, including thiolate ligation in heme proteins and electron-donating meso-substituents in heme models. The νFeN/νNO data for the Fe(III) complexes are reporters of heme pocket polarity and the accessibility of lone pair, Lewis base donors. Implications for biologically important processes, including NO binding, reductive nitrosylation and NO reduction, are discussed.
Introduction
Heme protein sensors1–3 of the gaseous ligands CO, NO and O2 play key regulatory roles in biology. Other heme proteins catalyze oxidative or reductive chemistry of the same ligands. The nature of the heme-ligand adducts and their interactions with the surrounding protein are of continuing interest.
Among these ligands, NO is unique in being able to bind heme in both Fe(II) and Fe(III) oxidation states. The odd electron on NO gives it flexible electronic properties. FeIII(NO) adducts are often transient, being easily reduced to the more stable FeII(NO) adducts. However, FeIII(NO) is stabilized in some heme proteins. These include the nitrophorins,4 used by blood-sucking insects to release NO as a vasodilator into the victim’s blood stream. Maintaining the Fe(III) oxidation state assures reversible binding and release of NO. FeIII(NO) is also involved in the catalytic cycle of NO synthases,5 and of NO redox enzymes, including nitrite reductase6 and NO reductase.7
The dominant mechanism for XO binding to heme (X = C, N, O) is backbonding, because the ligands have low-lying π* orbitals, well matched to the filled dπ orbitals on Fe(II). These orbitals are contracted on Fe(III), which is unable to bind CO or O2. However, the odd electron on NO can readily transfer to Fe(III), producing the FeII(NO+) resonance structure, isoelectronic with FeII(CO). DFT computations on model heme adduct show that FeII(NO+) is indeed the appropriate description of the FeIII(NO) ground state.8 However, Lehnert and coworkers report computation of a low-energy FeIII(NO) radical state, in which the FeIII–NO bond is slightly elongated and the odd electrons on Fe(III) and NO are antiferromagnetically coupled.8 This state is likely involved in the mechanism of FeIII–NO dissociation. Walker discussed this state previously and pointed out that it should be stabilized by porphyrin ruffling, a structural feature of the nitrophorins.4
Backbonding produces anticorrelation of the Fe–XO and X–O bond strengths, in response to polarization effects. Enhancement of backbonding strengthens Fe–XO while weakening X–O. Negative correlations between Fe–XO and X–O stretching frequencies, gathered from Raman and/or IR spectra, have been observed for Fe(II) adducts of all three XO ligands.9 However, FeIII(NO) adducts are exceptional. When the available literature data are plotted (Figure 1), the dominant νFeN/νNO correlation seems to be positive; the FeIII–NO and N–O bond strengths increase or decrease in concert for oxidized iron. Rodgers and coworkers10,11 have investigated this phenomenon computationally, and have found positive νFeN/νNO correlations when electron donating and withdrawing heme meso-substituents were used to polarize the FeIII(NO) adduct, a technique used previously,12,13 both experimentally and computationally to explore the negative correlations found for FeII(XO) adducts. They attributed the contrary behavior of FeIII(NO) to a unique occupied molecular orbital that is antibonding with respect to both the FeIII–NO and N–O bonds.10,11
Figure 1.
νFeN/νNO data for six-coordinate FeIII(NO) heme proteins and model complexes (see Table 1). “” indicates histidine-ligated heme-protein adducts, while “
” indicates cysteine-ligated heme-protein adducts; “
” and “
” represent FeIII(NO) porphyrin model complexes. The blue line is a linear correlation for thiolate-ligated adducts, with r = 0.84, after excluding the data for model complexes (SR and SR-HB).
However, the available experimental data are sparse (Table 1), owing to the limited stability of FeIII(NO) adducts. The data describing the positive correlation in Figure 1 are all from thiolate-ligated heme proteins and models. The available data for non-thiolate adducts appear as a scattered cluster in the upper right corner of the figure. In order to examine this cluster more closely, we have obtained a new data set from resonance Raman (RR) spectra of a series of myoglobin (Mb) variants, used previously to explore backbonding in FeII(CO)14 and FeII(NO)15 complexes (Figure 2). These data reveal negative νFeN/νNO correlations for FeIII(NO) adducts, fully consistent with backbonding. They also reveal a specific N–O strengthening effect that is associated with the distal histidine of Mb. DFT modeling confirmed the latter effect, showing it to result from a lone pair interaction of the imidazole with the positively charged FeII(NO+), an effect that was originally postulated by Miller et al.16 Further modeling revealed that a positive νFeN/νNO correlation stems from FeIII–N–O bending, which can be induced by several mechanisms, including thiolate ligation. Implications of these results for heme protein function are discussed.
Table 1.
FeIII–NO and N–O Stretching Frequencies (cm−1) for Six-Coordinate FeIII(NO) Complexes of Heme Proteins and Porphyrin Models.
Protein/porphyrin | ν(Fe–NO) | ν(N–O) | Refs |
---|---|---|---|
(His)FeIII(NO) complexes | |||
NORa | 594 | 1904 | 7 |
HRPb | 604 | 1903 | 17, 18 |
hHO-1c | 596 | 1918 | 19 |
NP1d (pH 8.0) | 591 | 1917 | 20, 21 |
NP1 (pH 5.6) | 591 | 1904 | 20, 21 |
NP4e (pH 7.5) | 593 | 1921 | 22, 23 |
NP4 (pH 5.5) | 590 | 1904 | 22, 23 |
HbNf | 591 | 1914 | 24 |
HbN-Y33Fg | 592 | 1908 | 24 |
HbAh | 594 | 1925 | 17 |
Mbi | 594 | 1921 | t.w.p |
OEP(Pyr)j | 602 | 1917 | 25, 26 |
(Cys)FeIII(NO) complexes | |||
P450camk (substrate free) | 528 | 1806 | 27, 28 |
P450cam+camphor | 522 | 1806 | 27, 28 |
P450cam+norcamphor | 524 | 1818 | 27, 28 |
P450cam+adamantanone | 520 | 1818 | 27 |
CPOl | 538 | 1868 | 27, 28 |
P450norm | 529 | 1851 | 28, 29 |
P450nor (proto) | 530 | 1853 | 30 |
P450nor (meso) | 532 | 1852 | 30 |
P450nor (deutero) | 533 | 1851 | 30 |
SRn | 510 | 1828 | 31 |
SR-HBo | 515 | 1837 | 31 |
Bacterial nitric oxide reductase from Paracoccus denitrificans.
Horseradish peroxidase.
Human heme oxygenase-1.
Nitrophorin 1 from Rhodnius prolixus.
Nitrophorin 4 from Rhodnius prolixus.
Truncated hemoglobin from Mycobacterium tuberculosis.
Tyrosine 33 to phenylalanine mutant of HbN.
Human hemoglobin.
Myoglobin.
Fe(III)-octaethylporphyrin with pyridine as proximal ligand.
Cytochrome P450cam.
Chloroperoxidase.
Cytochrome P450 nitric oxide reductase from Fusarium oxysporum.
Fe(III) porphyrin-alkanethiolate complex.
Hydrogen-bonded SR.
This work.
Figure 2.
Structure of sperm whale MbII(CO) (pdb ID: 1A6G), showing the mutated distal residues in the variants studied.
Materials and Methods
Mb Samples
Preparation of the Mb variants has been described elsewhere.32,33 Mb samples were converted to the Fe(III) (metMb) form by adding ca. 10 fold excess of sodium ferricyanide (Aldrich), and removing excess ferricyanide with a Sephadex G-25 PD-10 column (GE Healthcare), pre-equilibrated with 100 mM sodium phosphate, pH 7.0 buffer. FeIII(NO) adducts were then prepared either by passing 14NO gas (Matheson Gas Products) anaerobically through a 200 μL, 0.1–0.2 mM sample in a sealed NMR tube, or by injecting the required amount of chemically generated 15NO via an air-tight syringe.
Resonance Raman Measurements
RR spectra were collected for the νFeN region using 413.1-nm excitation from a Kr+ ion laser (Spectra Physics, 2080-RS), and for the νNO region using 244-nm excitation from an intracavity doubled-Ar+ ion laser (Innova 300 FReD, Coherent Radiation, Palo Alto, CA). Photodissociation and reduction of the NO adducts were minimized by using low laser power (~1 mW at the sample) and by spinning the sample. For the visible RR experiments, the scattered light was collected and focused onto a triple spectrograph (Spex 1877) equipped with a CCD detector (Roper Scientific, Model 7375-0001) operating at −110 °C. Spectra were calibrated with dimethylsulfoxide-d6. For the ultraviolet RR (UVRR) experiments, a single spectrograph (Spex 1269) equipped with a UV-enhanced CCD detector (Princeton Instruments, Model LN/CCD-1340/400) was used to collect the scattered light. UVRR spectra were calibrated using acetone and dimethylsulfoxide-d6.
DFT Calculations
Density functional theory (DFT; B3LYP) calculations on the model complexes were performed using the Gaussian 03 program.34 The standard 6-31G* basis set was used for all the atoms except Fe, for which Ahlrichs’ valence triple-ζ (VTZ) basis set was chosen.35 Geometry optimization and frequency calculations were performed using tight convergence criteria and an ultrafine integration grid. Cs symmetry constraints were used for the FeIII(NO) porphyrin models without distal imidazole, where proximal imidazole was kept in the plane bisecting the Fe–N(pyrrole) bonds. For the models with distal imidazole, structures were optimized under Cs or no symmetry (C1) constraints as specified for each case. Vibrational frequency values were taken directly from the Gaussian program without scaling. In all calculations, a singlet ground state for the FeIII(NO) complexes was assumed.
Results
Resonance Raman Spectra of FeIII(NO) Mb Variants
Although a number of RR spectra for FeIII(NO) heme adducts have been reported (Table 1), the data are sparse. Facile autoreduction to the more stable FeII(NO) adducts is undoubtedly responsible for the relative paucity of literature reports. Although we hoped to obtain systematic data on model complexes, as we had for FeII(CO)12 and FeII(NO)13,36 adducts, our efforts have been frustrated by autoreduction and high background scattering.
We have, however, obtained high quality visible and ultraviolet RR spectra of carefully prepared FeIII(NO) adducts of Mb using low laser intensities. The samples included Mb variants with distal residue substitutions (Figure 2), designed to vary the electric field distribution around the bound NO.14,37 These variations have been shown to produce systematic vibrational shifts in FeII(CO)38 and FeII(NO)15 adducts of Mb, and we sought similar effects for the isoelectronic FeIII(NO).
RR spectra are illustrated in Figure 3 for wild-type (WT) Mb. The spectrum obtained with 413.1-nm excitation, in resonance with the heme Soret absorption, reveals the FeIII–NO stretching and bending vibrations at 594 and 571 cm−1, respectively, assigned17 by the shifts induced upon substituting 15NO for 14NO. To detect the N–O stretching vibration, we utilized excitation at 244 nm, where enhancement of this mode occurs, probably via resonance with an Fe(III)←NO ligand-to-metal charge-transfer (LMCT) transition.18
Figure 3.
RR spectra for FeIII(NO) adducts of wild-type Mb containing 14NO or 15NO, and the difference spectra (14NO – 15NO).
Spectra of similar quality were obtained for the remaining Mb variants; some of their 14NO – 15NO difference spectra are displayed in Figures 4 and 5, which reveal a novel and unexpected pattern. The νFeN bands all fall in a narrow range, 590–600 cm−1, but the νNO bands occupy two distinct regions, one at ~1920 cm−1, and one at ~1900 cm−1. The common thread is that the distal histidine residue, His64, is present in the ~1920 cm−1 variants (Figure 4), but is replaced by hydrophobic residues (Ile or Leu) in the ~1900 cm−1 variants (Figure 5). However, two of the His64-containing variants, F46L and F46V, show both νNO bands. In these variants the bulky Phe46 residue, which buttresses the His64 side chain (Figure 2), is replaced by smaller residues. Other studies have shown that this substitution increases the mobility of His64, whose side chain can occupy alternative positions, an inward conformer neighboring the heme NO ligand or an outward conformer away from it.14,39,40 Thus, the two νNO bands in these mutants reflect populations in which His64 does or does not interact with the bound NO.
Figure 4.
RR difference spectra (14NO – 15NO) for FeIII(NO) adducts of the indicated Mb variants, which retain the distal His64.
Figure 5.
RR difference spectra (14NO – 15NO) for FeIII(NO) adducts of the indicated Mb variants, in which distal His64 is replaced by hydrophobic residues leucine or isoleucine.
These data are listed in Table 2, along with data from additional variants (spectra in Figure S1), in which His64 is replaced by the polar residue Gln, or in which Arg45 is replaced with smaller or oppositely charged amino acids. Again the ~1900- and 1920-cm−1 bands show up simultaneously. The R45A and R45E substitutions eliminate an anchoring salt-bridge between Arg45 and a heme propionate, and increase the mobility of His64,14,41,42 thus accounting for the two populations. In the R45K variant, this salt-bridge is preserved by the positively charged Lys45, and only the WT-like population is seen. For the H64Q variants, we infer that there are again two populations, this time with Gln64 interacting or not with the bound NO, in the same manner as His64. The Gln64 side chain can donate a lone pair from the carbonyl Oε atom, but is expected to be more mobile than that of His64.
Table 2.
FeIII–NO and N–O Stretching Frequencies (cm−1) of Six-Coordinate FeIII(NO) Myoglobin Mutants.
Mb mutant | ν(Fe–NO) | ν(N–O) (%)a |
---|---|---|
WT | 594 | 1921 |
V68F | 590 | 1923 |
F46L | 595 | 1920 (61), 1901 (39) |
F46V | 596 | 1920 (48), 1902 (52) |
L29F | 598 | 1918 |
H64I | 597 | 1900 |
H64L | 596 | 1898 |
H64L/V68F | 595 | 1901 |
H64L/L29F | 599 | 1899 |
H64L/V68N | 598 | 1911 |
H64Q | 597 | 1920 (44), 1903 (56) |
H64Q/V68F | 594 | 1921 (75), 1903 (25) |
H64Q/L29F | 598 | 1919 (36), 1900 (64) |
H64Q/L29F/V68F | 596 | 1919 (82), 1901 (18) |
R45A | 594 | 1923 (81), 1905 (19) |
R45E | 594 | 1923 (75), 1905 (25) |
R45K | 594 | 1922 |
For Mb variants with two νNO frequencies, the percentage area for each band is shown in parentheses.
When νFeN is plotted against νNO (Figure 6), it is apparent that there are two parallel correlations (solid lines), one in which His64 or Gln64 interacts with the NO, and one in which this interaction is absent. Both correlations are negative, as expected for backbonding, and the values of the slopes, −1.1 and −1.2, respectively, are somewhat larger than those observed when νFeC is plotted against νCO for FeII(CO) adducts, −0.73.43 Side chain substitutions that increase the positive polarity near the bound NO, e.g., via the Phe multipole in L29F,44 shift the point higher on either line.
Figure 6.
νFeN/νNO data for FeIII(NO) adducts of the indicated Mb variants, without () and with (
) lone-pair donation to the bound NO from a distal His64 or Gln64. Asterisks indicate minor components when two νNO populations are detected. The red and blue lines are linear correlations with r = 0.88 and 0.84, respectively.
DFT Modeling
Density functional theory was used to analyze trends in structure and vibrational frequencies for FeIII(NO) heme adducts. The basic model was (ImH)FeIII(NO)P, where ImH is imidazole and P is porphine. The level of theory was the same (see Material and Methods section) as was used successfully for FeII(CO) adducts.36 The calculated structure (see bond parameters for FeP in Table 3) agrees well with those reported by others,8,10,45 using different levels of theory, and also with available model complex crystallographic data.46,47 No MbIII(NO) crystal structure is available, but EXAFS analysis48 has indicated a linear FeIII–N–O unit, and bond distances consistent with the models.
Table 3.
Selected Bond Lengths (Å), Bond Angles (deg), and Vibrational Stretching Frequencies (cm−1) Calculated for Six-Coordinate (ImH)FeIII(NO), (ImH)FeIII(NO)···(ImH) and (ImH)FeIII(NO)···(H2O) Adducts.a
Complex | Fe–NO | N–O | ∠FeNO | Fe–NImp | NNO···NImd | ONO···NImd | ν(Fe–NO) | ν(N–O) |
---|---|---|---|---|---|---|---|---|
(ImH)–FeIII–NO, Cs symmetry | ||||||||
FeP(NH2)8 | 1.630 | 1.138 | 178.4 | 2.027 | – | – | 666 | 2057 |
FeP(CH3)8 | 1.632 | 1.137 | 179.7 | 2.025 | – | – | 666 | 2063 |
FeP | 1.636 | 1.136 | 179.8 | 2.020 | – | – | 660 | 2070 |
FePF8 | 1.642 | 1.134 | 179.8 | 2.019 | – | – | 649 | 2077 |
FePCl8 | 1.643 | 1.133 | 179.8 | 2.014 | – | – | 647 | 2079 |
FeP(NO2)8 | 1.656 | 1.130 | 179.9 | 2.005 | – | – | 629 | 2092 |
(ImH)–FeIII–NO←:N(ImH), Cs symmetry | ||||||||
FeP(NH2)8 | 1.642 | 1.132 | 175.4 | 2.026 | 3.454 | 2.855 | 650 | 2084 |
FeP(CH3)8 | 1.645 | 1.131 | 175.3 | 2.023 | 3.426 | 2.832 | 647 | 2088 |
FeP | 1.650 | 1.130 | 175.8 | 2.020 | 3.401 | 2.777 | 639 | 2096 |
FePF8 | 1.657 | 1.128 | 177.3 | 2.019 | 3.353 | 2.706 | 626 | 2103 |
FePCl8 | 1.660 | 1.127 | 177.2 | 2.014 | 3.346 | 2.685 | 623 | 2104 |
FeP(NO2)8 | 1.683 | 1.126 | 178.5 | 2.005 | 3.264 | 2.535 | 575 | 2101 |
(ImH)–FeIII–NO←:N(ImH), C1 symmetry | ||||||||
FeP(NH2)8 | 1.638 | 1.133 | 176.1 | 2.028 | 3.351 | 2.966 | 656 | 2080 |
FeP(CH3)8 | 1.642 | 1.132 | 174.5 | 2.025 | 3.238 | 2.929 | 650 | 2085 |
FeP | 1.648 | 1.131 | 172.2 | 2.021 | 3.118 | 2.920 | 641 | 2089 |
FePF4 | 1.654 | 1.130 | 169.4 | 2.019 | 3.010 | 2.921 | 627 | 2088 |
FePF8 | 1.662 | 1.130 | 166.8 | 2.016 | 2.922 | 2.917 | 609 | 2083 |
FePCl8 | 1.672 | 1.130 | 163.4 | 2.010 | 2.794 | 2.897 | 587 | 2072 |
FeP(NO2)8 | 1.713 | 1.131 | 156.0 | 2.003 | 2.529 | 2.797 | 519 | 2041 |
(ImH)–FeIII–NO←(:OH2), C1 symmetry | ||||||||
FeP | 1.647 | 1.131 | 173.4 | 2.020 | 2.949 | 2.807 | 644 | 2092 |
(ImH)–FeIII–NO←(:OH2)2, C1 symmetry | ||||||||
FeP | 1.655 | 1.126 | 174.6 | 2.022 | 2.97/3.12 | 2.74/2.78 | 635 | 2117 |
(ImH)–FeIIIPH–NO, constrained ∠FeIIINO, Cs symmetry | ||||||||
FeP | 1.755 | 1.146 | 140° | 1.981 | – | – | 440 | 1938 |
FeP | 1.701 | 1.141 | 150° | 1.989 | – | – | 505 | 1993 |
FeP | 1.664 | 1.138 | 160° | 2.002 | – | – | 594 | 2036 |
FeP | 1.643 | 1.136 | 170° | 2.014 | – | – | 645 | 2062 |
FeP | 1.641 | 1.136 | 172° | 2.017 | – | – | 649 | 2065 |
FeP | 1.639 | 1.136 | 174° | 2.018 | – | – | 657 | 2067 |
NImp and NImd are the nitrogen atoms of proximal and distal imidazoles, respectively (see Figure 9).
The computed νFeN and νNO frequencies, 660 and 2070 cm−1, are overestimated, as is usually the case with DFT, for which empirical scaling factors are often applied.49–51 The required scaling factor for νNO, ~0.92, is indeed the same as determined by Pulay and coworkers50 and consistent with the range generally encountered for molecules of the first row elements.51 Scaling factors for metal–ligand bonds have not been collected. Since we are interested in frequency trends, and not in absolute values, we have not applied scaling factors.
(a) Inductive Effects of Peripheral β-Substituents
To examine the effects of backbonding, we resorted to the same method as employed previously for FeII(CO)12,36 and FeII(NO)36 adducts, attaching electron donating or withdrawing substituents to the porphine periphery. Electron donors are expected to increase backdonation to the axial ligand, via their inductive effect, while electron-withdrawing substituents should decrease backdonation. This computational device has been applied to (ImH)FeIII(NO)P by Linder and Rodgers,10 who found a positive correlation of νFeN and νNO, contrary to the negative correlation expected for backbonding. However, the positive correlation was based on a set of substituents attached to the meso-carbon atoms of porphine (Figure 7). Their calculations for substituents attached to the β-carbon atoms did not follow this correlation, and indeed the β-substituent points appeared to be negatively correlated.
Figure 7.
Fe porphine with eight substituents on the β-carbon atoms.
For FeII(CO) adducts the position of substitution does not materially affect the slope of the νFeC/νCO correlation,36 but for FeIII(NO) there is a specific electronic effect of meso-substituents, because of an altered orbital energy structure, as discussed below. Consequently, we limited the position of substitution to the β-carbon atoms. Table 3 lists structures and frequencies for (ImH)FeIII(NO)P with eight β-substituents, –NH2, –CH3, –H, –F, –Cl, and –NO2, arranged from the most donating to the most withdrawing. The resulting νFeN/νNO correlation (Figure 8) gives a straight line with a slope of −1.1, in excellent agreement with the experimental correlations found for the MbIII(NO) adducts. Thus, the experimental data behave according to the backbonding model.
Figure 8.
νFeN/νNO correlation plots computed for the six-coordinate (ImH)FeIII(NO)PX8 complexes, without (blue points) and with a distal ImH, held in Cs symmetry (red triangles). “○” indicates (ImH)FeIII(NO)P with a distal water molecule. Linear correlations are shown for models without distal ImH (blue line, r = 0.98), and with distal ImH in Cs symmetry (red line, r = 0.98), but excluding the –NO2 data point.
(b) Distal Imidazole Interaction
Next we modeled the effect of a distal histidine interaction, by placing an imidazole ring near the bound NO ligand. In the first set of calculations, the distal ImH was restricted to the same plane as the proximal ImH, which bisected the Fe–N(pyrrole) bonds (Cs overall symmetry, Figure 9a).
Figure 9.
DFT-optimized molecular structures and atom labeling for (ImH)FeIII(NO)P with a distal ImH held in Cs symmetry (a), or allowed to relax to C1 (b).
The distal ImH was initially placed with its N–H proton pointing at the NO, simulating H-bonding. This orientation is stable for FeII(CO)37,52–54 and FeII(NO)55,56 adducts, but for (ImH)FeIII(NO)P no stable structure could be found with this orientation. However, a stable structure was found when the imidazole N: lone pair was oriented toward the NO (Figure 9a). The imidazole ring was canted to one side of the NO, but the (ImH)N: atom was much closer to the ONO than the NNO atom, 2.78 vs. 3.40 Å (Table 3).
This stable structure supports the suggestion advanced some time ago by Miller et al.16 that the distal His64 in Mb engages in a lone pair donor interaction with FeIII(NO). Their suggestion was based on the observation, confirmed in the present work, that the νNO (measured via FTIR spectroscopy at 4 °C to prevent autoreduction; νFeN was not reported16) shifted down when His64 was replaced by a hydrophobic residue, whereas an upshift would have been the direction expected on the basis of backbonding, if His64 exerts positive polarity, as it does for FeII(CO).9,14
Next we computed structures and frequencies for β-substituted porphines, with the distal ImH in the same Cs-symmetry orientation (Figure 9a). The imidazole N: moves farther from the ONO atom as the substituents become more electron-donating, and the FeIII–N–O unit bends slightly, in the direction away from the distal imidazole (Table 3). The N–O distance increases while the Fe–NO distance decreases, as expected from increased backbonding.
The computed frequencies now fall on a new νFeN/νNO correlation (slope = −1.3), nearly parallel to the model without a distal ImH, and displaced ~20 cm−1 to higher νNO values (Figure 8). Thus, the pattern observed for the MbIII(NO) variants (Figure 6) is reproduced remarkably well. For a given porphyrin, the effect of the distal imidazole is to increase νNO and decrease νFeN (Table 3). Qualitatively, this effect is consistent with decreased backbonding, but the νNO increase is more pronounced, producing the observed shift in the correlation. (We note that agreement with experiment is inexact in that the Mb data indicate no effect of the distal imidazole on νFeN. For variants in which the distal His is mobile, two values of νNO are observed, but only one νFeN.)
What accounts for the shifted correlation? Figure 10 compares selected occupied molecular orbitals, computed for the distal imidazole model in Cs symmetry. The two orbitals on the left, HOMO-3 and -4, are the nearly degenerate pair that describes FeII–NO+ backbonding, and are similar to the backbonding orbitals in the absence of a distal imidazole. There is bonding overlap between Fe dπ and NO π* orbitals, and an antibonding interaction with the imidazole lone pair, that inhibits backbonding. These orbitals account for the negative slope of the νFeN/νNO correlation. The orbital on the right, HOMO-6, describes an imidazole lone-pair interaction, mainly with the NO π* orbital, with some involvement of π orbitals on the porphine, on the side opposite the distal imidazole. The imidazole lone pair pushes electrons out of the NO π* orbital onto the porphine, thereby strengthening the N–O bond, with little effect on the FeIII–NO bond. This interaction is the likely origin of the shift of the νFeN/νNO correlation to higher νNO when the distal imidazole interacts mainly with the O atom of NO.
Figure 10.
Selected HOMOs of the (ImH)N:→(NO)FeIII(ImH) porphine model (Cs).
We note that the νFeN/νNO point for distal imidazole interacting with the NO2-substituted porphine deviates from the general pattern, falling distinctly below the line for the other models (Figure 8). This deviation for the most electron-withdrawing substituent can be traced to HOMO-6 acquiring more Fe–NO antibonding character (data not shown) when the substituents are –NO2, and moves above the backbonding orbitals, thereby governing the FeIII–N–O bonding.
(c) O- Versus N-Directed Imidazole and FeIII–N–O Bending
In the course of computing vibrational frequencies, we noted that the Cs distal imidazole model yielded a negative frequency, ~ −16 cm−1, corresponding to rotation of the distal imidazole ring out of the plane to which it was constrained. While the negative frequency was low enough to preclude significant errors in the computed νNO and νFeN values, it did signal that the Cs constraint imposed an energy penalty, probably due to non-bonded contacts between the imidazole CH and the porphine macrocycle. At the close distance required for the N: lone pair interaction, this contact is within the sum of van der Waals radii (~ 3 Å).
When the Cs constraint was removed, the distal imidazole rotated out of the plane, as expected. In addition, however, it shifted toward the N atom of the NO (Table 3 and Figure 9b). The imidazole N: was now only slightly closer to the ONO than the NNO atom, 2.92 vs. 3.12 Å. Moreover, electron-withdrawing substituents pulled the (ImH)N: closer to the NNO than the ONO atom. They also induced significant FeIII–N–O bending. In the extreme case of eight –NO2 substituents, the distances of (ImH)N: to ONO and NNO were 2.80 and 2.53 Å, respectively, while the FeIII–N–O angle was reduced to 156° (Table 3).
When the computed frequencies were placed on the νFeN/νNO plot (Figure 11), a surprising pattern emerged. The data for the more electron-donating β-substituents (–NH2, –CH3, –H) fall on a shifted backbonding correlation (Figure 11, top red line), similar to that found for the Cs model (Figure 11, blue line), but the data for the more electron-withdrawing β-substituents (–F, –Cl, –NO2) fall on a positive correlation, both νNO and νFeN decreasing as the substituents became more electron-withdrawing (Figure 11, bottom red line). The break between the two correlations occurs between eight –H substituents (porphine) and four each of –F and –H substituents (H- and F4-labeled points in Figure 11). With more electron-donating β-substituents the (ImH)N: stays farther from the NNO than the ONO atom, and the FeIII–N–O unit stays nearly linear, while with more electron-withdrawing β-substituents the (ImH)N: moves closer to the NNO than the ONO atom, and the FeIII–N–O unit bends strongly (Table 3).
Figure 11.
νFeN/νNO correlation plots computed for the (ImH)FeIII(NO)PX8 and (ImH)FeIII(NO)PH4F4 complexes, without (blue points, data from Figure 8) and with a distal ImH allowed to relax to C1 symmetry (red triangles). The black line is the positive correlation of νFeN and νNO, obtained by constraining the FeIII–N–O angle at the indicated values in the (ImH)FeIII(NO)P model. Also shown is the HOMO-1 a2u-type orbital for the (ImH)FeIII(NO)P model with the ∠FeIII–N–O = 140° (electron density has been removed from the front side of the molecule, to show the Fe–N–O contribution).
A similar effect of the distal (ImH)N:→NNO interaction was observed by constraining the distance between the distal ImH and NNO at fixed values for unsubstituted FeIII(NO) porphine (C1 model). The closer the approach of the distal ImH to the N atom of NO, the more bent the FeIII–N–O angle became, reaching 154° at a 2.4 Å distance. The FeIII–NO and N–O bond lengths increased in concert as the distal (ImH)N:→NNO distance decreased (Table S1, Figure S2). We note that the energy well for the interaction is shallow; only 0.4 kcal/mol are lost on increasing the NNO···NImd distance from the bottom of the well, 3.1 Å, to 3.5 Å. When the distal ImH was constrained to interact with ONO (Cs model), the FeIII–N–O unit remained nearly linear (Table S2, Figure S3).
The positive νFeN/νNO correlation implies an effect that overrides backbonding, and that correlates with FeIII–N–O bending. To test the role of bending per se, we examined its effect on the porphine model, (ImH)FeIII(NO)P, by constraining the FeIII–N–O angle to different values (Table 3). Bending FeIII–N–O reduced both νNO and νFeN, producing a positive correlation (Figure 11, black line); a similar effect of bending was reported by Linder and Rodgers.10 Indeed, the slope of the bending correlation is the same as that produced by increasingly electron-withdrawing β-substituents (Figure 11, bottom red line). We note that a positive correlation was earlier computed for the FeII–CO adduct, (ImH)FeII(CO)P, upon bending the FeII–C–O unit.57 In both cases bending weakens the Fe–N(C) and N(C)–O bonds simultaneously.
The reason for this mutual weakening can be seen in the composition of the HOMO-1 molecular orbital, shown in Figure 11 (inset) for (ImH)FeIII(NO)P when FeIII–N–O is bent. Bending induces an antibonding combination of the dxz and π*NO orbitals, a feature noted earlier by Linder and Rodgers.10 It is this antibonding combination that produces simultaneous FeIII–NO and N–O bond weakening, and a positive νFeN/νNO correlation.
In addition, bending induces interaction of π*NO with the porphyrin a2u-type orbital (Figure 11, inset). The a2u orbital is concentrated on the pyrrole N atoms and the meso-C atoms, and is strongly affected by meso-substituents. As these substituents become electron-donating, the a2u-π*NO interaction becomes more favorable, inducing more FeIII–N–O bending, thereby accounting for the positive νFeN/νNO correlation computed by Linder and Rogers when the meso-substituents were varied. However, the a2u orbital has no amplitude at the β-carbon atoms of the pyrrole rings. Consequently, β-substituents do not induce FeIII–N–O bending. Instead their inductive effect shifts the extent of FeIII–NO backbonding, producing a negative νFeN/νNO correlation.
(d) Distal Water Interaction
Models were also constructed in which the distal imidazole was replaced by water molecules, in order to investigate a possible distal water effect on νFeN/νNO values of the nitrosyl ferriheme protein nitrophorin 4 (see below). Stable structures were found with the water O atoms pointing at the bound NO, somewhat closer to the ONO than the NNO atom (Table 3), with the FeIII–N–O unit remaining nearly linear. As with distal imidazole lone pair interacting with ONO, the effect of the interaction was to decrease νFeN but increase νNO, shifting the νFeN/νNO point above the interaction-free backbonding line. The H2O effect was qualitatively similar, placing the frequency point on the correlation line formed by the models with distal (ImH)N:→ONO interaction (Figure 8).
We note that Linder and Rodgers also found a stable structure for a formaldehyde molecule interacting with the bound NO in (CH3S−)FeIII(NO)P, modeling the interaction of a peptide carbonyl in the enzyme cytochrome P450 nitric oxide reductase (P450nor).58 As the contact distance was decreased, the computed FeIII–NO and N–O bond distances both decreased slightly, suggesting a positive νFeN/νNO correlation with variable interaction strength. This effect had been observed earlier by Scheidt and coworkers in crystal polymorphs of [FeIII(OEP)NO]ClO4 (OEP = octaethylporphyrin) having the perchlorate counterion at different distances from the NO.11,59 Both νFeN and νNO increased for the closer counterion contact.
Discussion
FeIII(NO) Backbonding and Bending
Our RR spectroscopic and DFT modeling results establish that, in the absence of forces that induce bending, backbonding dominates the behavior of FeIII(NO) adducts, consistent with the FeII(NO+) formulation of its electronic structure. A negative νFeN/νNO correlation is found, experimentally and computationally, for NO adducts of imidazole-ligated Fe(III) heme. The slope of the correlation is essentially the same whether the backbonding is varied by electron donating or withdrawing porphyrin β-substituents (computation, slope = −1.1), or by the protein electrostatic field (experiment, slope = −1.2). In addition, lone pair donation from a distal imidazole shifts the correlation to higher νNO values. This effect, seen experimentally in MbIII(NO) variants, is accurately modeled via DFT, and is attributable to a specific orbital interaction that strengthens the N–O bond, with little effect on the Fe–NO bond.
However, FeIII(NO) bending overrides the backbonding anti-correlation, and produces a positive νFeN/νNO correlation as the Fe–N–O angle is varied. This effect arises from induction of an antibonding combination of dxz and π*NO orbitals in HOMO-1 as FeIII(NO) bends, resulting in simultaneous weakening of the Fe–NO and N–O bonds. Bending can result from a distal lone pair donor that is drawn close enough to interact with the N instead of the O atom of the NO.
There are interesting parallels between distal lone pair interactions with FeIII(NO) adducts and distal H-bond interactions with FeII(NO) adducts, investigated earlier,36,55 which can best be appreciated with the aid of the valence bond diagrams in Figure 12. In both cases, interaction at the ONO atom polarizes the π bonds. Backbonding is enhanced by H-bonding to FeII(NO), but diminished by lone pair donation to FeIII(NO), which, in addition strengthens the N–O bond via a specific orbital interaction. In both Fe oxidation states, interaction at the NNO atom pulls electron density into the sp2-like antibonding orbital, inducing bending and weakening both the Fe–NO and N–O bonds.
Figure 12.
Different mechanisms of bond strength modulation through interaction with a distal imidazole lone pair (FeIII–NO adducts) or a distal imidazole H-bond (FeII–NO adducts). A. Modulation through the ONO polarization (backbonding). B. Modulation through interaction with NNO (Fe–N–O bending).
For FeII(NO), DFT computation reveals stable H-bond interactions at either ONO or NNO, depending on the disposition of the H-bond donor.55,56 Examples of both interactions are found in different heme proteins (see Ref. 55 and references therein). For FeIII(NO), the preferred lone-pair interaction is with ONO, unless the donor is drawn toward the NNO by electron-withdrawing heme β-substituents, or unless the donor is a strong nucleophile, like a hydroxide ion (see below). Since protoporphyrin IX does not have electron-withdrawing substituents, and since protein side-chains are not strong nucleophiles, the FeIII(NO) structural and vibrational data from ferriheme nitrosyls can be interpreted in terms of backbonding and lone-pair interactions with ONO, provided the proximal ligand is histidine.
Implications for Imidazole-Ligated Ferriheme Nitrosyls
With the νFeN/νNO correlations established for MbIII(NO), we can interpret reported vibrational data for FeIII(NO) adducts of other imidazole-ligated proteins and models (Figure 13). The extensively characterized nitrophorins from the saliva of the ‘kissing’ bug Rhodnius prolixus are of particular interest, because of the stability of their FeIII(NO) adducts. Two sets of FeIII(NO) stretching frequencies have been reported for nitrophorin 4 (NP4), at pH 7.5 and 5.5.22 Intriguingly, these fall on the two lines described by MbIII(NO) variants, with and without distal imidazole interactions. At the lower pH, NP4 is in a ‘closed’ form in which hydrophobic residues are packed around the bound NO. There is no potential lone pair donor, and the νFeN/νNO (590/1904 cm−1) position, at the low end of the imidazole-less backbonding line (Figure 13, red line), indicates an environment of low polarity.
Figure 13.
νFeN/νNO data for the indicated His-ligated FeIII(NO) heme proteins, which fall on the plus (blue line) or minus (red line) distal His lines of MbIII(NO) variants shown in Figure 6 or form a parallel intermediate linear correlation (dashed line with r = 0.96 and slope = −1.0). NOR = nitric oxide reductase, hHO-1 = human heme oxygenase-1, HbA = human hemoglobin, HbN = truncated hemoglobin with distal Tyr and Gln residues, NP4 = nitrophorin 4 and NP1 = nitrophorin 1 at the indicated pH values, HRP = horseradish peroxidase, and H64L/V68N = His64Leu/Val68Asn Mb double mutant. Pyridine adduct of a model complex FeIIIOEP(NO) (OEP = octaethylporphyrin) is also shown ().
An early crystal structure of the FeIII(NO) adduct of NP4 at pH 5.6 was refined as two conformers, one with a linear and one with a severely bent (110°) Fe–N–O unit.60 The severely bent conformation was later attributed to partial reduction to FeII(NO) in the X-ray beam, and a higher resolution gave a single conformer with ∠ Fe–N–O = 156°, but with thermal disorder at the O position.61 It seems likely that this determination too was influenced by X-ray-induced reduction, since the νFeN/νNO data would have shown a pronounced deviation below the backbonding line for this degree of bending (see Figure 11), and this is not observed.
At pH 7.5, a conformational change in NP4 induces disorder in a pair of loops lining the bound NO and opens the distal heme pocket to solvent. The νFeN/νNO data (593/1921 cm−1) place this form on the distal lone-pair line (Figure 13, blue line), yet there is no distal histidine, nor any other potential donor side-chain on the distal side of the heme; an aspartate residue is ~6 Å from the Fe,61 too far to donate an electron pair, though it could polarize an intervening water molecule. The actual ferriheme–NO structure is not available at pH 7.5, because of instability to X-ray-induced reduction, but crystal structures of NH3 and CN− adducts reveal an open, water-filled distal pocket.60–62 In the CN− adduct, a distal threonine residue orients a water molecule, holding it 2.5 Å from CN− (pdb ID: 1EQD). Our DFT modeling indicates that a lone-pair donating water molecule could have an effect on νFeN/νNO equivalent to a distal imidazole (Table 3 and Figure 8).
Like NP4, nitrophorin 1 (NP1) has an open binding pocket at elevated pH, but the crystal structure of the NP1(CN−) adduct (pdb ID: 3NP1) shows fewer water molecules than in NP4(CN−), and none are localized to the immediate vicinity of the CN−, in contrast to NP4(CN−) (pdb ID: 1EQD). Consistent with this difference, the νFeN/νNO data point for NP1(NO) at pH 8.020,21 is shifted to an intermediate position (Figure 13, dashed line), indicating a weaker lone-pair donor interaction.
Human heme oxygenase-1 (hHO-1)19 falls on the distal imidazole line (Figure 13, blue line), but in this case the lone pair donor is the carbonyl oxygen of distal Gly139, which the crystal structure shows to be in close contact with bound NO, as seen in the FeII(NO) adduct (3.1 and 2.7 Å distances to NNO and ONO, respectively).63 Reported data for the pyridine adduct of FeIIIOEP(NO)25,26 also fall on the distal imidazole line, indicating that excess pyridine, present to ensure adduct formation, provided a lone pair interaction.
In nitric oxide reductase (NOR) from Paracoccus denitrificans, a cytochrome bc protein, NO binds to a vacant coordination site on imidazole-ligated ferriheme b3.64,65 Spectroscopic studies indicate that a trisimidazole-ligated non-heme FeB is distal to the heme b3 in NOR,66 analogous to the CuB center in cytochrome c oxidase (CcO),67–69 and is likely involved in NO reduction. A model complex has been synthesized, and interestingly, its FeIII(NO) vibrational frequencies,70 νFeN/νNO = 610/1924 cm−1, place it far above the MbIII(NO) correlations in Figure 13. This placement is reminiscent of the elevated νFeC/νCO frequencies found for the CO adduct of CcO.71,72 In that case, a compression effect of the heme/CuB dinuclear site was invoked to explain the elevated νFeC,43 and a similar explanation may apply to the NOR model. However, NOR itself does not show this effect. Its FeIII(NO) adduct (νFeN/νNO = 594/1904 cm−1)7 falls on the imidazole-less line (Figure 13, red line), indicating that the protein site is less constrained than the model.
In addition to NP1, several other proteins occupy intermediate positions in Figure 13; a dashed line has been drawn through them to guide the eye. One of them is the Mb variant H64L/V68N, in which the distal His is absent, but a new potential donor is introduced in place of Val68, at the side of the bound NO (Figure 2). This orientation may limit lone pair donation by the carbonyl Oε atom of the Asn68 side-chain, accounting for the intermediate νFeN/νNO point. However, Asn68 could also increase positive polarity of the distal environment via its NεH atoms, thereby enhancing backbonding, as reflected in the elevated νFeN. Backbonding is even more pronounced for the FeIII(NO) adduct of horseradish peroxidase (HRP), which has unusually high νFeN (Figure 13), while having an intermediate νNO. Distal Arg, as well as His, residues enhance positive polarity in HRP. Meanwhile, the distal His (or an intervening water molecule, as judged from the crystal structure of the FeII(CO) adduct, pdb ID: 1ATJ) can provide lone pair donation. However, the νNO value may in this case be limited by the anionic character of the strongly H-bonded proximal His ligand. This effect was also seen in the FeII(CO) adduct of HRP, for which the νFeC/νCO point falls below the neutral His ligand backbonding line.43
Energetics of the Lone Pair Interaction
The switch from H-bond donation by a distal His for FeII(NO) to lone pair donation for FeIII(NO) is a natural consequence of the increase in positive charge on the bound NO ligand. The lone pair interaction stabilizes the FeIII(NO) unit modestly, as can be seen from the changes in the NO association rate (kon), dissociation rate (koff), and association equilibrium (KNO) constants, shown in Table 4. Mutation of His64 to apolar amino acids results in a ~7 to 10-fold increase in the rate constant for NO dissociation, koff, from the FeIII(NO) adduct of Mb (Table 4).73 Assuming that the slowing of NO dissociation by His64 reflects the lone pair stabilization energy, the factor of 7–10 translates to stabilization of bound NO in native ferric Mb by −1.0 to −1.3 kcal/mol (i.e., –RTln(koff,mutant/koff,wt)).39 This estimate agrees well with our DFT-computed energy released upon moving a distal imidazole from a non-bonded distance of 3.4 Å to its optimum distance, 2.8 Å, ~ −1.2 kcal/mol for the Cs model (which does not induce bending, Figure S3). This interpretation is supported by the results for the His64 to Gln64 mutation in which lone pair donation from the Oε atom is possible, perhaps accounting for the smaller 5-fold increase in koff for this variant (Table 4).
Table 4.
Kinetic Parameters of NO Binding to Ferric and Ferrous States of WT Mb and Its Mutants.a
Mb | kon (μM−1s−1) | koff (s−1) | KNO (μM−1) | Ref. |
---|---|---|---|---|
MbIII(NO) | ||||
WT | 0.08 | 12 | 0.0067 | 73, 75 |
His64Gln | 8.2 | 59 | 0.19 | 73, 75 |
His64Leu | 44 | 80 | 0.55 | 73, 75 |
His64Val | 160 | 120 | 1.3 | 73 |
MbII(NO) | ||||
WT | 22 | 0.00010 | 220,000 | 39, 73, 74 |
His64Gln | 43 | 0.00011 | 370,000 | 39, 74 |
His64Leu | 190 | 0.00013 | 1,500,000 | 39, 74 |
His64Val | 270 | 0.0011 | 250,000 | 73 |
kon = NO association rate constant, koff = NO dissociation rate constant, KNO = NO affinity constant (kon/koff).
The NO affinity of FeIIIMb is markedly increased ~30, 80, and 200-fold for the Gln64, Leu64, and Val64 variants, respectively, because the NO association rate constants, kon, increase ~100 to 2,000-fold (Table 4).73 The slower bimolecular rates for NO binding to native FeIIIMb reflect the free energy barrier for displacement of the water molecule coordinated to the iron atom, which in wild-type FeIIIMb is also stabilized by lone pair donation by His64. As described by Olson and Phillips,39 the unfavorable free energy required to disrupt water coordination can be estimated from RTln(kon,mutant/kon,wt) and the value is ~ +3 to +4 kcal/mol, which is substantially larger than the favorable FeIII(NO) lone-pair interaction energy that stabilizes NO coordinated to FeIIIMb. The net result is that the affinity of wild-type FeIIIMb for NO is very small compared to that for the apolar Val64 or Leu64 FeIIIMb mutants (Table 4). The same His64 mutations also increase the association rate constant for NO binding to FeIIMb, but in this case, the increases in kon for FeIIMb are much smaller (10 to 20-fold) than those for FeIIIMb (500 to 1000-fold, Table 4). The favorable free energy associated with proton donation by His64 to NO bound to FeIIMb, estimated from the ratio of the mutant versus wt NO dissociation rate constants (≤ ~1 kcal/mol) is equal to or only slightly smaller than the unfavorable free energy associated with displacement of non-coordinated, internal water H-bonded to the distal histidine in FeIIMb, which can be estimated from the ratio of the corresponding association constants (~ +1.3 kcal/mol). The net result is that, unlike the situation for FeIIIMb, only small changes in NO affinity for FeIIMb are observed for His64 mutations (Table 4).39
FeIII(NO) Bending by Nucleophillic Attack: Reductive Nitrosylation and NO Reduction
Because electron-withdrawing substituents are absent in protoporphyrin IX, a distal imidazole is attracted mainly to the ONO atom of FeIII(NO) in ferriheme proteins. This is true also of water molecules (see above) or peptide carbonyl groups.58 However, it is likely that stronger nucleophiles would shift to the NNO atom, bending the FeIII–N–O unit and interacting with the sp2 orbital.
Nucleophile-induced bending is the likely pathway for “reductive nitrosylation”,76–78 in which hydroxide ion attacks the bound NO, forming nitrous acid, HONO, and Fe(II) heme (Figure 14). The Fe(II) heme is trapped by coordination with excess NO. Consistent with this mechanism, the rate of FeIII(NO) autoreduction increases rapidly with increasing pH.78 To model this pathway, we carried out DFT optimization of (ImH)FeIII(NO)P with hydroxide ion placed in proximity to the bound NO. Formation of a HONO intermediate has been proposed, based on kinetic studies of reductive nitrosylation in buffered solutions of MbIII, HbIII and CytIII.76 Consistent with this proposal, the modeled HO− promptly attacked the NNO atom, causing bending of the FeIII(NO) unit and formation of a stable HONO adduct (Figure 14, inset). This adduct is characterized by elongated Fe–NHONO and Fe–NImp bonds, and a lowered Mulliken charge on Fe (1.071e vs. 1.106e for the starting FeIII(NO) porphine), consistent with Fe reduction. In principle, a similar mechanism of reductive nitrosylation could be available to other nucleophiles, including nitrite, thiols, and amines.
Figure 14.
Proposed pathway for reductive nitrosylation of FeIII(NO)P. Inset: DFT-optimized structure of the adduct obtained as a result of interaction between HO− and (ImH)FeIII(NO)P. Indicated are bond distances (Å, black), and Mulliken charges (blue).
We speculate that NO reduction to N2O might proceed by a similar pathway, with bound NO− providing the nucleophile for N–N bond formation. Collman and coworkers have shown that their model compound for the binuclear nitric oxide reductase produces N2O and a FeIII–O–FeIII adduct upon treatment of the di-ferrous compound with NO.70 They were able to detect intermediates with NO bound successively to the non-porphyrin and the porphyrin Fe(II),79 providing evidence for a ‘trans’ mechanism, in which the reacting NO molecules reside transiently on both Fe atoms of the binuclear complex. The reaction pathway for this di-NO adduct is uncertain, but might involve electron transfer from the heme to the non-heme Fe(II), producing FeBI(NO) at the non-heme FeB site, capable of nucleophillic attack on the heme FeIII(NO) (Figure 15a). The driving force would be stabilization of the Fe(III) by the porphyrinate di-anion, and of the Fe(I) by the neutral histidine ligands at the non-heme site. The much higher Fe(III/II) reduction potential of the heme than the non-heme site, 0.32 vs. 0.06 V,70 is consistent with this site preference. Alternatively (Figure 15b), the first step in the pathway could be coupling of the two Fe(II)-bound NO neutral radicals, accompanied by concerted electron transfer from the two Fe(II)’s. However, spin restriction and the asymmetry of the binuclear site argue against this symmetric mechanism. The choice between the two pathways will rest upon detailed computation of the energy landscape for electron reorganization. However, the energy landscape may differ somewhat between the model compound and the NOR enzyme, since, as noted above, the model and the protein differ significantly in their FeIII(NO) vibrational frequencies, in a manner suggesting greater rigidity in the model.
Figure 15.
Alternative pathways for N2O production from the di-ferrous complex of the NOR binuclear model compund.70,79
FeIII(NO) Bending by Thiolate Ligation: Cysteinyl Hemes
In addition to a lone pair interaction at the NNO atom, FeIII–N–O bending can be induced by mixing of the dz2 and dxz orbitals, or more precisely the dz2-σn and dxz-π*NO orbitals, as has been discussed by others.80–83 This is the source of bending when thiolate is a ligand. The strongly donating RS− drives up the dz2-σn energy, and increases its mixing with dxz-π*NO. The crystal structure of a thiolate-ligated FeIII(NO) adduct reveals a 160° Fe–N–O angle,84 consistent with structures of cysteinyl heme proteins,28,29,85 which also produce a positive νFeN/νNO correlation (Figure 1). For the model complex (CH3S−)FeIII(NO)P, Linder and Rodgers computed an Fe–N–O angle of 159°, which increased to 180° when the CH3S− was protonated, thereby lowering its donor strength.58 Paulat and Lehnert computed structures and vibrational frequencies for FeIII(NO)P, with thiolate ligands having zero, one and two H-bonds to the sulfur atom.86 Both νFeN and νNO increased as the number of H-bonds increased, reflecting diminished dz2-σn/dxz-π*NO mixing as the thiolate donor strength diminished. A similar effect underlies the cysteinyl heme protein νFeN/νNO correlation in Figure 1. The νFeN/νNO frequencies increase in the order cytochrome P450cam < cytochrome P450nor < chloroperoxidase (CPO), which is also the order of increased H-bonding to the cysteinyl sulfur, as judged by analyses of the crystal structures,5 and also by vibrational data on the corresponding CO adducts.43
Conclusions
For imidazole-ligated FeIII(NO) adducts νFeN and νNO are negatively correlated, consistent with variations in backbonding, but distal lone pair donors induce an additional increase in νNO due to a specific orbital interaction. These donors can include imidazole, but also peptide carbonyl groups and water molecules. The lone pair interaction of a distal imidazole represents approximately −1.0 to −1.3 kcal/mol of stabilization energy.
The νFeN and νNO frequencies become positively correlated when the FeIII(NO) moiety bends, since the FeIII–NO and N–O bonds are then weakened simultaneously. Bending can be induced by different stereoelectronic mechanisms, including:
1] electron-donating substituents at the heme meso-positions, which induce mixing of the porphyrin a2u-type orbital with a bending-induced antibonding combination of dxz and π*NO orbitals,
2] a distal lone pair drawn close enough by electron-withdrawing β-substituents, or by strong nucleophilicity, to interact with a bending-induced sp2-like orbital on NNO, and
3] a strongly donating proximal axial ligand that raises the dz2 orbital energy and induces mixing of the dz2-σn and dxz-π*NO orbitals.
The first two of these mechanisms are generally inapplicable to heme proteins, since the protoporphyrin IX meso-substituents (–H) are not electron-donating and the β-substituents (methyl, vinyl, propionate) are not electron-withdrawing, and since the available lone pair donors are generally weak nucleophiles. However, attack of a strong nucleophile is facilitated by bending-induced exposure of the empty orbital on the NNO atom. Examples include attack of hydroxide to form a transient HONO adduct, during reductive nitrosylation, and possibly the attack of FeB-bound NO− in the NO coupling reaction induced by NO reductase.
The third mechanism is operative in heme proteins with the strongly donating cysteinate ligand. Here, the νFeN and νNO frequencies are positively correlated, with the degree of bending modulated by H-bond donation to the cysteinate.
Supplementary Material
Acknowledgments
This work was supported financially by NIH grants GM033576 (TGS), GM035649 (JSO), and HL047020 (JSO), and by the Robert A. Welch Foundation grants C-0612 (JSO) and E-1184 (RSC).
Footnotes
Supporting Information Available: Complete reference 34, RR difference spectra (14NO – 15NO) for FeIII(NO) adducts of Mb variants having distal His64 replaced by Gln (Figure S1), distal ImH···NNO distance dependence for the (ImH)FeIII(NO)···(ImH) porphine complex in C1 symmetry (Table S1, Figure S2), distal ImH···ONO distance dependence for the (ImH)FeIII(NO)···(ImH) porphine complex in Cs symmetry (Table S2, Figure S3), and Cartesian coordinates of DFT-optimized structures for the models depicted in Figure 9 (Table S3). This material is available free of charge via the Internet at http://pubs.acs.org.
References
- 1.Chan MK. Curr Opin Chem Biol. 2001;5:216–222. doi: 10.1016/s1367-5931(00)00193-9. [DOI] [PubMed] [Google Scholar]
- 2.Gilles-Gonzalez MA, Gonzalez G. J Inorg Biochem. 2005;99:1–22. doi: 10.1016/j.jinorgbio.2004.11.006. [DOI] [PubMed] [Google Scholar]
- 3.Aono S. Dalton Transac. 2008:3137–3146. doi: 10.1039/b802070c. [DOI] [PubMed] [Google Scholar]
- 4.Walker FA. J Inorg Biochem. 2005;99:216–236. doi: 10.1016/j.jinorgbio.2004.10.009. [DOI] [PubMed] [Google Scholar]
- 5.Rousseau DL, Li D, Couture M, Yeh SR. J Inorg Biochem. 2005;99:306–323. doi: 10.1016/j.jinorgbio.2004.11.007. [DOI] [PubMed] [Google Scholar]
- 6.George SJ, Allen JWA, Ferguson SJ, Thorneley RNF. J Biol Chem. 2000;275:33231–33237. doi: 10.1074/jbc.M005033200. [DOI] [PubMed] [Google Scholar]
- 7.Pinakoulaki E, Gemeinhardt S, Saraste M, Varotsis C. J Biol Chem. 2002;277:23407–23413. doi: 10.1074/jbc.M201913200. [DOI] [PubMed] [Google Scholar]
- 8.Praneeth VKK, Paulat F, Berto TC, DeBeer George S, Nather C, Sulok C, Lehnert N. J Am Chem Soc. 2008;130:15288–15303. doi: 10.1021/ja801860u. [DOI] [PubMed] [Google Scholar]
- 9.Spiro TG, Ibrahim M, Wasbotten IH, Ghosh A. The Smallest Biomolecules. Elsevier; New York: 2008. pp. 96–123. [Google Scholar]
- 10.Linder DP, Rodgers KR. Inorg Chem. 2005;44:1367–1380. doi: 10.1021/ic049045h. [DOI] [PubMed] [Google Scholar]
- 11.Linder DP, Rodgers KR, Banister J, Wyllie GRA, Ellison MK, Scheidt WR. J Am Chem Soc. 2004;126:14136–14148. doi: 10.1021/ja046942b. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12.Vogel KM, Kozlowski PM, Zgierski MZ, Spiro TG. Inorg Chim Acta. 2000;297:11–17. [Google Scholar]
- 13.Vogel KM, Kozlowski PM, Zgierski MZ, Spiro TG. J Am Chem Soc. 1999;121:9915–9921. [Google Scholar]
- 14.Phillips J, George N, Teodoro ML, Li T, Smith B, Olson JS. J Phys Chem B. 1999;103:8817–8829. [Google Scholar]
- 15.Coyle CM, Vogel KM, Rush TS, III, Kozlowski PM, Williams R, Spiro TG, Dou Y, Ikeda-Saito M, Olson JS, Zgierski MZ. Biochemistry. 2003;42:4896–4903. doi: 10.1021/bi026395b. [DOI] [PubMed] [Google Scholar]
- 16.Miller LM, Pedraza AJ, Chance MR. Biochemistry. 1997;36:12199–12207. doi: 10.1021/bi962744o. [DOI] [PubMed] [Google Scholar]
- 17.Benko B, Yu NT. Proc Natl Acad Sci USA. 1983;80:7042–7046. doi: 10.1073/pnas.80.22.7042. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 18.Tomita T, Haruta N, Aki M, Kitagawa T, Ikeda-Saito M. J Am Chem Soc. 2001;123:2666–2667. doi: 10.1021/ja001431k. [DOI] [PubMed] [Google Scholar]
- 19.Wang J, Lu S, Moenne-Loccoz P, Ortiz de Montellano PR. J Biol Chem. 2003;278:2341–2347. doi: 10.1074/jbc.M211131200. [DOI] [PubMed] [Google Scholar]
- 20.Maes EM, Walker FA, Montfort WR, Czernuszewicz RS. J Am Chem Soc. 2001;123:11664–11672. doi: 10.1021/ja0031927. [DOI] [PubMed] [Google Scholar]
- 21.Ding XD, Weichsel A, Andersen JF, Shokhireva TK, Balfour C, Pierik AJ, Averill BA, Montfort WR, Walker FA. J Am Chem Soc. 1999;121:128–138. [Google Scholar]
- 22.Zaczek MB, Zareba AA, Czernuszewicz RS, Montfort WR. J Porphyrins Phthalocyanines. 2006;10:928. [Google Scholar]
- 23.Nienhaus K, Maes EM, Weichsel A, Montfort WR, Nienhaus GU. J Biol Chem. 2004;279:39401–39407. doi: 10.1074/jbc.M406178200. [DOI] [PubMed] [Google Scholar]
- 24.Mukai M, Ouellet Y, Ouellet H, Guertin M, Yeh SR. Biochemistry. 2004;43:2764–2770. doi: 10.1021/bi035798o. [DOI] [PubMed] [Google Scholar]
- 25.Lipscomb LA, Lee B-S, Yu N-T. Inorg Chem. 1993;32:281–286. [Google Scholar]
- 26.Mu XH, Kadish KM. Inorg Chem. 1988;27:4720–4725. [Google Scholar]
- 27.Hu S, Kincaid JR. J Am Chem Soc. 1991;113:2843–2850. [Google Scholar]
- 28.Obayashi E, Tsukamoto K, Adachi S, Takahashi S, Nomura M, Iizuka T, Shoun H, Shiro Y. J Am Chem Soc. 1997;119:7807–7816. [Google Scholar]
- 29.Shimizu H, Obayashi E, Gomi Y, Arakawa H, Park S-Y, Nakamura H, Adachi S-I, Shoun H, Shiro Y. J Biol Chem. 2000;275:4816–4826. doi: 10.1074/jbc.275.7.4816. [DOI] [PubMed] [Google Scholar]
- 30.Singh UP, Obayashi E, Takahashi S, Iizuka T, Shoun H, Shiro Y. Biochim Biophys Acta. 1998;1384:103–111. doi: 10.1016/s0167-4838(98)00006-5. [DOI] [PubMed] [Google Scholar]
- 31.Suzuki N, Higuchi T, Urano Y, Kikuchi K, Uchida T, Mukai M, Kitagawa T, Nagano T. J Am Chem Soc. 2000;122:12059–12060. [Google Scholar]
- 32.Springer BA, Egeberg KD, Sligar SG, Rohlfs RJ, Mathews AJ, Olson JS. J Biol Chem. 1989;264:3057–3060. [PubMed] [Google Scholar]
- 33.Dou Y, Maillett DH, Eich RF, Olson JS. Biophys Chem. 2002;98:127–148. doi: 10.1016/s0301-4622(02)00090-x. [DOI] [PubMed] [Google Scholar]
- 34.Frisch MJ, et al. Gaussian 03, Revision E.01. Gaussian, Inc; Wallingford, CT: 2004. [Google Scholar]
- 35.Bauernschmitt R, Ahlrichs R. Chem Phys Lett. 1996;256:454–464. [Google Scholar]
- 36.Ibrahim M, Xu C, Spiro TG. J Am Chem Soc. 2006;128:16834–16845. doi: 10.1021/ja064859d. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 37.Franzen S. J Am Chem Soc. 2002;124:13271–13281. doi: 10.1021/ja017708d. [DOI] [PubMed] [Google Scholar]
- 38.Li T, Quillin ML, Phillips GN, Olson JS. Biochemistry. 1994;33:1433–1446. doi: 10.1021/bi00172a021. [DOI] [PubMed] [Google Scholar]
- 39.Olson JS, Phillips GN. J Biol Inorg Chem. 1997;2:544–552. [Google Scholar]
- 40.Lai HH, Li T, Lyons DS, Phillips GNJ, Olson JS, Gibson QH. Proteins Struct Funct Genet. 1995;22:322–339. doi: 10.1002/prot.340220404. [DOI] [PubMed] [Google Scholar]
- 41.Harada K, Makino M, Sugimoto H, Hirota S, Matsuo T, Shiro Y, Hisaeda Y, Hayashi T. Biochemistry. 2007;46:9406–9416. doi: 10.1021/bi7007068. [DOI] [PubMed] [Google Scholar]
- 42.Carver TE, Olson JS, Smerdon SJ, Krzywda S, Wilkinson AJ, Gibson QH, Blackmore RS, Ropp JD, Sligar S. Biochemistry. 1991;30:4697–4705. doi: 10.1021/bi00233a009. [DOI] [PubMed] [Google Scholar]
- 43.Spiro TG, Wasbotten IH. J Inorg Biochem. 2005;99:34–44. doi: 10.1016/j.jinorgbio.2004.09.026. [DOI] [PubMed] [Google Scholar]
- 44.Carver TE, Brantley RE, Singleton EW, Arduini RM, Quillin ML, Phillips GN, Olson JS. J Biol Chem. 1992;267:14443–14450. [PubMed] [Google Scholar]
- 45.Wondimagegn T, Ghosh A. J Am Chem Soc. 2001;123:5680–5683. doi: 10.1021/ja004314y. [DOI] [PubMed] [Google Scholar]
- 46.Ellison MK, Scheidt WR. J Am Chem Soc. 1999;121:5210–5219. [Google Scholar]
- 47.Ellison MK, Schulz CE, Scheidt WR. J Am Chem Soc. 2002;124:13833–13841. doi: 10.1021/ja0207145. [DOI] [PubMed] [Google Scholar]
- 48.Rich AM, Armstrong RS, Ellis PJ, Lay PA. J Am Chem Soc. 1998;120:10827–10836. [Google Scholar]
- 49.Rauhut G, Pulay P. J Phys Chem. 1995;99:3093–3100. [Google Scholar]
- 50.Rauhut G, Jarzecki AA, Pulay P. J Comput Chem. 1997;18:489–500. [Google Scholar]
- 51.Merrick JP, Moran D, Radom L. J Phys Chem A. 2007;111:11683–11700. doi: 10.1021/jp073974n. [DOI] [PubMed] [Google Scholar]
- 52.Sigfridsson E, Ryde U. J Biol Inorg Chem. 1999;4:99–110. doi: 10.1007/s007750050293. [DOI] [PubMed] [Google Scholar]
- 53.Rovira C, Schulze B, Eichinger M, Evanseck JD, Parrinello M. Biophys J. 2001;81:435–445. doi: 10.1016/S0006-3495(01)75711-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 54.Sigfridsson E, Ryde U. J Inorg Biochem. 2002;91:101–115. doi: 10.1016/s0162-0134(02)00426-9. [DOI] [PubMed] [Google Scholar]
- 55.Xu C, Spiro TG. J Biol Inorg Chem. 2008;13:613–621. doi: 10.1007/s00775-008-0349-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 56.Tangen E, Svadberg A, Ghosh A. Inorg Chem. 2005;44:7802–7805. doi: 10.1021/ic050486q. [DOI] [PubMed] [Google Scholar]
- 57.Kozlowski PM, Vogel KM, Zgierski MZ, Spiro TG. J Porphyrins Phthalocyanines. 2001;5:312–322. [Google Scholar]
- 58.Linder DP, Rodgers KR. J Biol Inorg Chem. 2007;12:721–731. doi: 10.1007/s00775-007-0223-0. [DOI] [PubMed] [Google Scholar]
- 59.Ellison MK, Schulz CE, Scheidt WR. Inorg Chem. 2000;39:5102–5110. doi: 10.1021/ic000789e. [DOI] [PubMed] [Google Scholar]
- 60.Weichsel A, Andersen JF, Roberts SA, Montfort WR. Nature Struct Biol. 2000;7:551–554. doi: 10.1038/76769. [DOI] [PubMed] [Google Scholar]
- 61.Roberts SA, Weichsel A, Qiu Y, Shelnutt JA, Walker FA, Montfort WR. Biochemistry. 2001;40:11327–11337. doi: 10.1021/bi0109257. [DOI] [PubMed] [Google Scholar]
- 62.Kondrashov DA, Roberts SA, Weichsel A, Montfort WR. Biochemistry. 2004;43:13637–13647. doi: 10.1021/bi0483155. [DOI] [PubMed] [Google Scholar]
- 63.Lad L, Wang J, Li H, Friedman J, Bhaskar B, Ortiz de Montellano PR, Poulos TL. J Mol Biol. 2003;330:527–538. doi: 10.1016/s0022-2836(03)00578-3. [DOI] [PubMed] [Google Scholar]
- 64.Wasser IM, De Vries S, Moenne-Loccoz P, Schroder I, Karlin KD. Chem Rev. 2002;102:1201–1234. doi: 10.1021/cr0006627. [DOI] [PubMed] [Google Scholar]
- 65.Zumft WG. Microbiol Mol Biol Rev. 1997;61:533–616. doi: 10.1128/mmbr.61.4.533-616.1997. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 66.Moenne-Loccoz P, Richter OMH, Huang H, Wasser IM, Ghiladi RA, Karlin KD, de Vries S. J Am Chem Soc. 2000;122:9344–9345. [Google Scholar]
- 67.Scott RA. Ann Rev Biophys Biophys Chem. 1989;18:137–158. doi: 10.1146/annurev.bb.18.060189.001033. [DOI] [PubMed] [Google Scholar]
- 68.Powers L, Chance B, Ching Y, Angiolillo P. Biophys J. 1981;34:465–498. doi: 10.1016/S0006-3495(81)84863-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 69.Thomson AJ, Greenwood C, Gadsby PMA, Peterson J, Eglinton DG, Hill BC, Nicholls P. J Inorg Biochem. 1985;23:187–197. doi: 10.1016/0162-0134(85)85025-x. [DOI] [PubMed] [Google Scholar]
- 70.Collman JP, Yang Y, Dey A, Decreau RA, Ghosh S, Ohta T, Solomon EI. Proc Natl Acad Sci USA. 2008;105:15660–15665. doi: 10.1073/pnas.0808606105. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 71.Wang JL, Takahashi S, Hosler JP, Mitchell DM, Ferguson-Miller S, Gennis RB, Rousseau DL. Biochemistry. 1995;34:9819–9825. doi: 10.1021/bi00031a001. [DOI] [PubMed] [Google Scholar]
- 72.Das TK, Tomson FL, Gennis RB, Gordon M, Rousseau DL. Biophys J. 2001;80:2039–2045. doi: 10.1016/S0006-3495(01)76177-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 73.Foley EW. PhD Dissertation. Rice University; Houston, TX: 2006. [Google Scholar]
- 74.Brucker EA, Olson JS, Ikeda-Saito M, Phillips JGN. Proteins Struct Funct Genet. 1998;30:352–356. [PubMed] [Google Scholar]
- 75.Eich RF, Li T, Lemon DD, Doherty DH, Curry SR, Aitken JF, Mathews AJ, Johnson KA, Smith RD, Phillips GN, Olson JS. Biochemistry. 1996;35:6976–6983. doi: 10.1021/bi960442g. [DOI] [PubMed] [Google Scholar]
- 76.Hoshino M, Maeda M, Konishi R, Seki H, Ford PC. J Am Chem Soc. 1996;118:5702–5707. [Google Scholar]
- 77.Fernandez BO, Lorkovic IM, Ford PC. Inorg Chem. 2004;43:5393–5402. doi: 10.1021/ic049532x. [DOI] [PubMed] [Google Scholar]
- 78.Ford PC, Fernandez BO, Lim MD. Chem Rev. 2005;105:2439–2455. doi: 10.1021/cr0307289. [DOI] [PubMed] [Google Scholar]
- 79.Collman JP, Dey A, Yang Y, Decreau RA, Ohta T, Solomon EI. J Am Chem Soc. 2008;130:16498–16499. doi: 10.1021/ja807700n. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 80.Mingos DMP. Inorg Chem. 1973;12:1209–1211. [Google Scholar]
- 81.Pierpont CG, Eisenberg R. J Am Chem Soc. 1971;93:4905–4907. [Google Scholar]
- 82.Hawkins TW, Hall MB. Inorg Chem. 1980;19:1735–1739. [Google Scholar]
- 83.Hoffmann R, Chen MML, Thorn DL. Inorg Chem. 1976;16:503–511. [Google Scholar]
- 84.Xu N, Powell DR, Cheng L, Richter-Addo GB. Chem Commun. 2006:2030–2032. doi: 10.1039/b602611g. [DOI] [PubMed] [Google Scholar]
- 85.Pant K, Crane BR. Biochemistry. 2006;45:2537–2544. doi: 10.1021/bi0518848. [DOI] [PubMed] [Google Scholar]
- 86.Paulat F, Lehnert N. Inorg Chem. 2007;46:1547–1549. doi: 10.1021/ic070023f. [DOI] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.