Abstract

Nitroxyl (HNO) exhibits unique favorable properties in regulating biological and pharmacological activities. However, currently, there is only one Cu-based HNO sensor that can be recycled for reusable detection, which is a Cu cyclam derivative with a mixed thia/aza ligand. To elucidate the missing mechanistic origin of its high HNO reactivity and subsequent favorable conformation change toward a stable CuI product that is critical to be oxidized back by the physiological O2 level for HNO detection again, a density functional theory (DFT) computational study was performed. It not only reproduced experimental structural and reaction properties but also, more importantly, revealed an unknown role of the coordination atom in high reactivity. Its conformation change mechanism was found to not follow the previously proposed one but involve a novel favorable rotation pathway. Several newly designed complexes incorporating beneficial effects of coordination atoms and substituents to further enhance HNO reactivity while maintaining or even improving favorable conformation changes for reusable HNO detection were computationally validated. These novel results will facilitate the future development of reusable HNO sensors for true spatiotemporal resolution and repeated detection.
Short abstract
HNO possesses significant biomedical functions. Unknown roles of coordination atoms and substituents on excellent reusable HNO sensors were discovered with several new designs to aid the future development.
Introduction
Nitric oxide (NO) is a well-known signaling molecule in many physiological processes.1 Among biological nitrogen oxide species, nitroxyl (HNO), the one-electron-reduced derivative of NO, has gained much attention for its diverse physiological and pharmacological effects.2−5 For instance, HNO is more effective than NO in protecting against ischemia-reperfusion injury.6 Another remarkable activity is its potent inhibition of platelet aggregation in patients with sickle cell disease.7,8 HNO was also used to prevent alcoholism by inhibiting aldehyde dehydrogenase9 and to treat cancer through reactions with thiols and thiol proteins.10 The extensive therapeutic effects of HNO ignite considerable scientific interest in studying its chemical and biological profiles.
However, experimental HNO investigation is difficult due to its reactive nature. It is highly desirable to develop efficient detection methods for recurrent in vivo use to better understand and examine its physiological and pharmacological functions. Conventional HNO detection methods use analytic instruments to quantify products of HNO dimerization/dehydration or reactions with thiols, phosphine, and heme proteins.11−17 However, most of them are indirect, time-consuming, and relatively unreliable. To overcome the limitations, direct HNO detection methods based on mass spectrometry and electrochemistry were developed.18−21 For detections inside living cells, invivo HNO sensors are accessible with bioimaging functions using metal complexes,22−34 phosphines,34−52 thiols,53,54 or esters.55
Despite these developments, there are no reusable fluorescent sensors for biological HNO. Recently, a Cu complex with a mixed thia/aza cyclam derivative 14-N2S2 (Scheme 1) was reported as the first Cu-based reusable HNO sensor.56 Reusability is a strongly desired feature in HNO sensor development since only reusable sensors can provide true spatiotemporal resolution57 and repeated detection58 of sudden bursts of in vivo analyte production. It is also cost-effective compared to one-time sensors. Using soft sulfur atoms is intended to stabilize the reduced Cu center upon reaction with HNO to prevent demetalation in prior one-time-use Cu-based HNO sensors.56 The stable CuI complex can then be oxidized back to the starting CuII system by the physiological level of O2 for the next HNO detection. The 14-N2S2 ligand also offers excellent HNO reactivity compared to the inactive aza-only cyclam (14-N4, Scheme 1).
Scheme 1. (A) Schematic Pathways of HNO Reactions with Cu Macrocyclic Complexes Studied in This Work (a Suffix of N, S, Se, Ts, or Bn Is Added, Respectively, for a Species of a Specific Ligand Listed Above; Coordination Atom Numberings Are Shown in I-1) and (B) trans-I, trans-III, and Tetrahedral Conformations of [Cu(14-N2S2)] (Bn = Benzyl; Ts = Tosyl; Atom Color Scheme (H Omitted for Clarity): C, Green; N, Blue; S, Yellow; Cu, Brown).

However, the mechanistic origins of its greater HNO reactivity and favorable conformation changes to retain Cu are unknown. So, we report a density functional theory (DFT) study of these missing mechanistic details, which were based on the previously found proton-coupled electron transfer (PCET) mechanism and some favorable substituent effects in other metal complexes and metalloproteins.59−61
To help the future development of reusable HNO sensors, we examined new ligand designs (Scheme 1): 14-N2Se2 with modifications on coordination atoms and substituted ligands with both an electron-donating group (EDG) of benzyl (14-N2S2-Bn) and an electron-withdrawing group (EWG) of tosyl (14-N2S2-Ts). Bn/Ts groups were installed on similar Cu cyclams and thus are feasible for future experimental synthesis.30,56 These new studies provide interesting aspects to further enhance HNO reactivities and/or conformation change efficiencies to facilitate the future development of potential reusable fluorescent sensors for biological HNO.
Computational Details
Geometry optimization and subsequent frequency calculations were performed using the mPW1PW9162 method with the LanL2DZ63−65 basis set for the Cu, the 6-311++G(2d,2p) basis set for HNO and Cu-coordinated N/S/Se atoms in the ligand, and the 6-31G(d) basis set for the rest of the atoms. The experimental aqueous solution effect was included using the PCM approach.66 All calculations were performed using Gaussian 0967 and Gaussian 16.68 In each case, the frequency analysis was used to verify the nature of the stationary points on respective potential energy surfaces and to provide zero-point energy-corrected electronic energies (EZPE’s), enthalpies (H’s), and Gibbs free energies (G′s) at 298.15 K and 1 atm. This approach was selected based on a methodological and basis set study69 of several DFT methods, including the more recently (compared to mPW1PW91) developed M0670 and dispersion-corrected ωB97XD,71 as well as comparisons with all-electron basis sets 6-31+G(2d) and 6-311++G(2d,2p), on HNO reactions. This approach has also been recently shown to give accurate descriptions of HNO reactivities in diverse metalloproteins and model systems including quantitative errors well below 1 kcal/mol as compared to experimental reaction barriers.59,60,72,73 The atomic charges and spin densities are from the natural population analysis (NPA) and Mulliken schemes implemented in Gaussian 09/16. Intrinsic reaction coordinate calculations in Gaussian 09/16 were performed to prove the connection between each transition state and its corresponding reactant and product.
To further evaluate the above method of choice, additional methodological studies of both methods and basis sets were conducted. Among several common basis set schemes for the first-row transition metal Cu, namely, 6-311++G(2d,2p), triple-ζ Def2TZVP,74,75 and Wachters’ basis sets,76 the all-electron 6-311++G(2d,2p) basis set has the greatest number of basis functions, so it was used for comparative analysis with the LanL2DZ basis set below. The latest hybrid DFT method MN1577 in Gaussian 16 has the best combined performance compared to 12 selected functionals including the previously studied M0670 and ωB97XD.71,77 Therefore, we compared MN15 with mPW1PW9162 using 6-311++G(2d,2p) and LanL2DZ63−65 basis sets for Cu and the same basis set for other atoms as described above. Geometric comparison of optimized Cu(II) and Cu(I) mixed thia/aza complexes with the X-ray structures in aqueous solution showed no significant differences between the three computational approaches: mPW1PW91/LanL2DZ, MN15/LanL2DZ, and MN15/6-311++G(2d,2p). Compared to the mPW1PW91/LanL2DZ data, results from MN15/6-311++G(2d,2p) only have a slightly better average mean percentage deviation (MPD) of 0.4% and average mean absolute deviation (MAD) of 0.01 Å, while MN15/LanL2DZ results have a mildly worse MPD of 0.8% and MAD of 0.02 Å; see Table S2. The computed trends of the most favorable conformations for Cu(II) and Cu(I) mixed thia/aza complexes also exhibited identical results to experimental data for all studied methods (Table S3). Regarding reaction barriers of the HNO to NO conversion pathways for the mixed thia/aza and pure aza complexes [Cu(14-N2S2)] and [Cu(14-N4)], compared to those from using the mPW1PW91/LanL2DZ method, ΔG data from MN15/LanL2DZ and MN15/6-311++G(2d,2p) are only with MADs of 0.44 and 0.67 kcal/mol; see Table S11. Additional results in Tables S12–S14 show that compared to the current method of mPW1PW91/LanL2DZ, the combination of MN15 with the largest basis studied here, i.e., 6-311++G(2d,2p), has average MADs of only ∼0.03 Å, ∼0.02 e, and ∼0.02 e, for bond lengths, atomic charges, and spin densities respectively, which are basically within computational errors. These methodological results show that the current method is more efficient with similar accuracy to the more advanced and time-consuming method, so the following discussion is based on data from the current approach.
Results and Discussion
Favorable Conformations of CuII and CuI Complexes
Due to the importance of conformation changes here, we first studied the conformation effects of experimental CuII/CuI complexes. The used computational method enabled accurate descriptions of HNO reactivities in various metalloproteins and model systems59−61,69,73,78 and accurate structural predictions of related Cu cyclams.59 The optimized gas phase structures of [CuII(14-N2S2)]2+ and [CuI(14-N2S2)]+ showed excellent agreement with the crystal structures, with an average mean absolute deviation (MAD) of 0.045 Å for all coordination bond lengths and mean percentage deviation (MPD) of only 2.1% (Table S1). As experimental reactions occurred in aqueous solution, optimizations using water as solvent were done, which shows similar structures with a better MAD of 0.040 Å and MPD of 1.8% (Table S2). These data show excellent prediction accuracy of the current computational method.
Because CuII cyclams have two common square planar conformations,79−84trans-III and trans-I (Scheme 1B, with coordination atom’s angles of C–N–C or C–X–C featuring two pointing above/below the plane and all four pointing above the plane; here, C represents neighboring carbons), and the CuI complex of this reusable HNO sensor has a tetrahedral conformation, these three conformations were studied for [CuII(14-N2S2)]2+ and [CuI(14-N2S2)]+ as well as [CuII(14-N4)]2+ and [CuI(14-N4)]+. As shown in Table S3, the experimental trans-III form of [CuII(14-N2S2)]2+ and the tetrahedral conformation of [CuI(14-N2S2)]2+(56) were all found to be most stable, which further validates our calculations. In contrast, for the aza-only complexes, trans-III conformations are preferred for both oxidized and reduced Cu states, with the tetrahedral conformations being of >10 kcal/mol higher energies. Clearly, these results support the experimental use of sulfurs to selectively stabilize the tetrahedral conformation of [CuI(14-N2S2)]2+, critical to prevent demetalation and thus enable recycling it back for reuse.
Cu reduction results in coordination bond length increase; see Table S4. The average Cu bond length increases are 0.119 and 0.118 Å in the trans-III conformation, respectively, for the 14-N2S2 and 14-N4 ligands, which are basically the same. However, the use of a tetrahedral conformation for the reduced complex minimizes structural changes since the average Cu coordination bond length variations are now significantly decreased to 0.024 and 0.063 Å, respectively, for the mixed and aza-only systems. The latter system has 2.6-fold larger geometric changes, indicating that the mixed thia/aza ligand dramatically decreases structural change and its associated energy cost. Larger sulfurs in 14-N2S2 vs nitrogens leads to an enlarged coordination sphere with an average Cu coordination bond length of 2.198 vs 2.039 Å in 14-N4, which leaves a larger space to accommodate the Cu size increase after reduction to an average Cu bond length of 2.222 Å (only 0.024 Å expansion for 14-N2S2). These data provide the first theoretical insights into the preferential stabilization of the tetrahedral conformation for [CuI(14-N2S2)]+.
HNO Reactivity
We then studied HNO reaction pathways for Cu complexes with the aza-only cyclam and its mixed thia/aza derivative to understand their significant experimental reactivity difference. For the starting CuII complexes with both ligands, the trans-III conformations are dominant with ∼98–99% based on their energy differences from other conformers (see Table S3), which was used in their reaction pathway calculations.
As shown in Scheme 1A, the basic pathway is like those with other Cu complexes and metalloproteins:59−61,69 it starts with the HNO-bound intermediate I-1-S (its formation is experimentally facilitated by large excess of the HNO donor56), which undergoes a PCET transition state (TS-S) to generate another intermediate I-2-S, where a proton transfers from HNO to one coordinated nitrogen and an electron from HNO transmits to Cu. It can be seen by the spin density data of the NO fragment (ρNOαβ) for the 14-N2S2 systems increasing from ∼0 e in I-1-S to 0.122 e in TS-S and then to ∼1.0 e in I-2-S and accordingly Cu (ρCuαβ) decreasing from 0.395 e in I-1-S to 0.381 e in TS-S and then to ∼0 e in I-2-S for the electron transfer part. The proton transfer process is illustrated by the H–N3 distance shortening from 3.661 Å in I-1-S to 2.027 Å in TS-S and then to 1.020 Å in I-2-S; see Table 1. The TS-S structure is shown in Figure 1B, and the structures of all species are shown in Figure S5. Another significant change is the Cu–N3 bond cleavage due to proton transfer to N3; see the Cu–N3 bond lengths of 2.075, 3.153, and 3.416 Å in I-1-S, TS-S, and I-2-S, respectively. So, its elongation of ∼1 Å in TS-S is much more significant than the 0.001 Å elongation of the H–N bond (Table S23, also see key changes in Figure 1B). This suggests that breaking the Cu–N3 bond is the most energy-demanding part. Then, NO dissociates to yield the reduced Cu complex I-3-S. This reaction’s Gibbs free energy of activation (ΔG‡) of ∼21 kcal/mol is close to some room-temperature reactions,85−87 and the final reaction energy from I-1 to I-3+NO (ΔG°) of −18.53 kcal/mol is very negative, which shows a kinetically feasible and thermodynamically favorable reaction to support its experimentally observed facile HNO reactivity.56
Table 1. Relative Energies, Key Geometric Parameters, Spin Densities, and Charges in the HNO to NO Conversion Pathways.
| ligand | species | ΔE (kcal/mol) | ΔEzpe (kcal/mol) | ΔH (kcal/mol) | ΔG (kcal/mol) | RCuN (Å) | RHN3 (Å) | RCuN3 (Å) | QCu (e) | QN3 (e) | ραβCu (e) | ραβNO (e) |
|---|---|---|---|---|---|---|---|---|---|---|---|---|
| 14-N2S2 | I-1-S | 0.00 | 0.00 | 0.00 | 0.00 | 2.552 | 3.661 | 2.075 | 0.604 | –0.745 | 0.395 | 0.003 |
| TS-S | 20.61 | 19.94 | 19.31 | 21.10 | 2.012 | 2.027 | 3.153 | 0.584 | –0.764 | 0.381 | 0.122 | |
| I-2-S | –7.74 | –7.24 | –6.85 | –8.40 | 2.019 | 1.020 | 3.416 | 0.244 | –0.604 | 0.017 | 0.987 | |
| I-3-S + NO | –8.67 | –8.94 | –8.84 | –18.53 | 1.030 | 3.080 | 0.354 | –0.635 | ||||
| 14-N4 | I-1-N | 0.00 | 0.00 | 0.00 | 0.00 | 2.692 | 4.104 | 2.048 | 0.966 | –0.753 | 0.501 | 0.001 |
| TS-N | 27.62 | 26.06 | 25.58 | 26.86 | 2.052 | 1.853 | 3.085 | 0.921 | –0.771 | 0.484 | 0.106 | |
| I-2-N | 4.96 | 5.16 | 5.28 | 4.96 | 2.027 | 1.028 | 3.205 | 0.698 | –0.625 | –0.204 | 1.328 | |
| I-3-N + NO | 0.27 | 0.22 | 0.24 | –0.22 | 1.031 | 2.895 | 0.621 | –0.614 | ||||
| 14-N2Se2 | I-1-Se | 0.00 | 0.00 | 0.00 | 0.00 | 2.576 | 4.197 | 2.090 | 0.531 | –0.744 | 0.369 | 0.009 |
| TS-Se | 18.26 | 17.88 | 17.24 | 19.14 | 1.990 | 2.059 | 3.291 | 0.533 | –0.762 | 0.370 | 0.111 | |
| I-2-Se | –10.38 | –9.85 | –9.53 | –10.64 | 2.028 | 1.030 | 3.228 | 0.137 | –0.637 | 0.015 | 0.962 | |
| I-3-Se + NO | –12.26 | –12.36 | –12.31 | –21.62 | 1.036 | 3.138 | 0.286 | –0.646 | ||||
| 14-N2S2-Ts | I-1-Ts | 0.00 | 0.00 | 0.00 | 0.00 | 2.334 | 4.533 | 2.099 | 0.570 | –0.737 | 0.391 | 0.020 |
| TS-Ts | 20.81 | 19.70 | 19.10 | 20.92 | 2.021 | 1.815 | 3.094 | 0.546 | –0.775 | 0.324 | 0.253 | |
| I-2-Ts | –19.48 | –18.54 | –18.05 | –19.66 | 2.181 | 1.027 | 3.280 | 0.226 | –0.631 | 0.038 | 0.943 | |
| I-3-Ts + NO | –16.06 | –16.11 | –15.83 | –26.14 | 1.030 | 3.157 | 0.367 | –0.625 | ||||
| 14-N2S2-Bn | I-1-Bn | 0.00 | 0.00 | 0.00 | 0.00 | 2.581 | 4.318 | 2.106 | 0.613 | –0.579 | 0.408 | 0.003 |
| TS-Bn | 16.94 | 15.93 | 15.42 | 16.57 | 2.026 | 1.952 | 3.208 | 0.628 | –0.601 | 0.349 | 0.145 | |
| I-2-Bn | –17.68 | –16.91 | –16.50 | –18.78 | 2.229 | 1.029 | 3.293 | 0.236 | –0.505 | 0.032 | 0.950 | |
| I-3-Bn + NO | –15.37 | –18.22 | –15.33 | –25.20 | 1.032 | 3.126 | 0.379 | –0.506 |
Figure 1.

(A) Gibbs free energies of the HNO to NO conversion pathways with all studied Cu macrocyclic complexes. (B) Optimized transition-state structures with key geometric changes of RCuN, RCuN3, and RHN from I-1 to TS. Atom color scheme: C, cyan; N, blue; O, red; S, yellow; Se, orange; H, gray; Cu, coral.
The alternative pathway with a proton transferred to another coordinated nitrogen N1 was also studied. It has a much higher ΔG‡ by 6.96 kcal/mol. This is because N3 points toward HNO, while N1 points away from HNO (see Figures S1 and S5), and thus HNO’s H is much closer to N3 in I-1-S than to N1 in I-1′-S by 0.233 Å to shorten the proton transfer distance for the N3 site (Table S16). So, only the favorable N3 pathway was used in the subsequent study.
Experimentally, the use of the 14-N4 ligand does not exhibit any reactivity toward HNO.56 As seen from Table 1 and Figure 1A, this aza-only ligated system has a much higher barrier of ∼27 kcal/mol, a positive ΔG for I-2-N formation, and a basically net-zero reaction energy (thus higher than that for the 14-N2S2 system by 18.31 kcal/mol). So, its reaction has no kinetic feasibility and thermodynamic favorability. These data also point out interesting reaction feature differences because of coordination atom changes not reported before:59,61,69 (1) HNO binds more favorably to the Cu center in I-1-S than in I-1-N, as indicated by the 0.14 Å shorter Cu–N distance in Table 1. This could be a result of less steric hindrance for HNO binding due to the enlarged coordinated sphere for the 14-N2S2 ligand (S is significantly larger than N). This makes HNO’s H closer to N3 by 0.443 Å in I-1-S than in I-1-N to facilitate the proton transfer (Table 1); (2) the enlarged coordination sphere makes the Cu–N3 bond longer by ∼0.03 Å in I-1-S and thus weaker for lower cleavage cost; (3) the less electronegative sulfur vs nitrogen renders a significantly less positively charged Cu (0.604 e vs 0.966 e for Cu in I-1-S and I-1-N, respectively, Table 1), which reduces the attraction between Cu and the negatively charged N3 to reduce the energy cost to break the Cu–N3 bond (the largest energy consumption part in TS).
Based on the effects of S vs N, we hypothesized that using the even larger and softer Se atom to replace S could further enhance the HNO reactivity and evaluated it. Similar to 14-N2S2, [CuII(14-N2Se2)]2+ also exists predominantly in the trans-III conformation (95% based on the energy difference with other common conformations, Table S6), which was used in the pathway calculations. As shown in Table 1 and Figure 1A (blue energy levels), this system indeed reduces ΔG‡ and ΔG° further by 2–3 kcal/mol compared to 14-N2S2. The even larger and softer Se atom leads to an even longer Cu–N3 bond (by 0.015 Å) and less positively charged Cu (by 0.073 e) to further weaken the Cu–N3 bond to cut energy cost. These data provide additional evidence of the coordination atom effect on HNO reactivity found here to facilitate future HNO sensor design.
Since a recent computational work59 revealed beneficial effects of either EDG or EWG substituents on the aza-only 14-N4 ligand in experimental Cu cyclams,30,56 we examined such effects on the mixed thia/aza ligand, i.e., 14-N2S2-Ts and 14-N2S2-Bn (Scheme 1). Bn/Ts atoms were installed on similar Cu cyclams and thus are feasible for future experimental synthesis.30,56 The reason to have only one substituent is to minimize the steric hindrance, the primary effect on HNO reactivity found recently.59
We first optimized their starting CuII complexes in the commonly found trans-III and trans-I conformations to determine the more favorable one for subsequent pathway studies, which show <1 kcal/mol energy differences (Table S6), i.e., within typical computational errors. This suggests coexistence. We then calculated the pathways for both conformations and found that the trans-I conformation has more favorable ΔG‡ and ΔG° by 3–4 kcal/mol (Table S10), which was used in the following discussion. Among the two nitrogen sites on these two ligands, the nitrogen with the Bn substitution for 14-N2S2-Bn and that opposite to the Ts substituent for 14-N2S2-Ts were used as proton acceptors based on their more favorable protonation energies from a prior work on similarly substituted cyclams.59
As shown in Table 1 and Figure 1A (green energy levels), the Ts group on the mixed thia/aza framework further decreases ΔG° by 7.61 kcal/mol to make it much more thermodynamically favorable, although its effect on ΔG‡ is mild (0.18 kcal/mol). This EWG helps attract the electron-rich HNO to facilitate its binding, leading to the shortest Cu–N distance for all I-1 species (Table 1). Its Cu–N3 bond is longer than that in the parent I-1-S system by 0.024 Å, indicating a more weakened Cu–N3 bond to facilitate this reaction. The electron-withdrawing nature of Ts may help enhance Cu’s redox potential to facilitate the electron transfer part of this PCET reaction. Indeed, ρCuαβ change from I-1-Ts to TS-Ts due to electron transfer is the largest here; see Table 1.
In contrast, Bn further reduces ΔG‡ by ∼4 kcal/mol while having a similar ΔG° to Ts; see Table 1 and violet energy levels in Figure 1A. This EDG on N3 not only helps improve this nitrogen’s proton affinity to promote the proton transfer in this reaction but also contributes to a steric effect on the Cu–N3 bond to elongate it, which results in the longest Cu–N3 bond among all I-1 species. Therefore, it has the weakest Cu–N3 bond to save energy cost most in TS, which makes its ΔG‡ the lowest here.
Overall, all three designed complexes show promising improved HNO reactivities with favorable coordination atom and substituent effects. Results also highlight the importance of reducing Cu–N3 bond cleavage energy cost on HNO reactivity. In fact, for all five complexes here, the linear correlation coefficients for barriers vs RCuN3’s and barriers vs |QCu × QN3| data at I-1 are 0.92 and 0.96, respectively (see Figure 2A,B). These excellent correlations show that a weak Cu–N3 bond with a long Cu–N3 bond length and a small electrostatic attraction between Cu and N3, as indicated by a small value of |QCu × QN3|, is associated with a small barrier. These excellent quantitative relationships may help future evaluation of new HNO sensors.
Figure 2.
Linear correlation plots of (A) reaction barriers vs RCuN3 at I-1 and (B) reaction barriers vs |QCu × QN3| at I-1.
Deprotonation
The previously unknown mechanism of spontaneous changes from trans-III I-3-S to the experimental tetrahedral CuI product P–S, essential to prevent the demetalation problem of prior Cu-based HNO sensors,56 was then investigated together with the three new ligands.
Because I-3-S has an extra proton compared to P–S, deprotonation on N3 is needed via a base (A–) in the solution; see Scheme 2. We first performed calculations to see whether the trans-III to tetrahedral conformation change should occur before or after deprotonation. As shown in Table S19, I-3-S is preferred in trans-III by 2 kcal/mol, which corresponds to 97% population. The deprotonated I-5-S, however, has a thermodynamic trend to undergo the conformation change to be tetrahedral, which is more stable by ∼4 kcal/mol. These results suggest that I-3-S is to be deprotonated first to trigger the conformation change.
Scheme 2. Deprotonation Step.
Color sphere represents the transferred proton.
According to pKa data of the potential proton abstractors in the experimental buffered pH 7 solution,56 such as the buffer PIPES (piperazine-N,N′-bis(2-ethanesulfonic acid); pKa, 6.76), the counterion OTf– (pKa, −14), water (pKa, −1.7), and OH– (pKa, 15.74),88 PIPES is the most possible one for use in calculations.
As shown in Scheme 2 with optimized structures in Figure S6, PIPES first forms an adduct I-4-S. In TS-1-S, the proton that is more outward is abstracted by PIPES (the inward one is inaccessible due to steric hindrance of PIPES) to form I-5-S. The increase of the distance between this proton and N3 and the concomitant decrease of the distance between this proton and N(PIPES), the nitrogen atom in PIPES that accepts it, can be clearly seen in Figure 3C.
Figure 3.

(A) Gibbs free energies of deprotonation pathways for 14-N2S2, 14-N2Se2, and 14-N2S2-Ts. (B) Gibbs free energies of the first substituent rotation and the deprotonation pathway for 14-N2S2-Bn. (C) Optimized transition-state structures with key geometric changes of RHN3 and RHN(PIPES) from I-4 to TS-1 (TS-2 for 14-N2S2-Bn). Atom color scheme: C, cyan; N, blue; O, red; S, yellow; Se, orange; H, gray; Cu, coral.
Energy results from I-3-S to I-5-S in Figure 3A show that deprotonation is kinetically and thermodynamically favorable with ΔG‡ and ΔG° of 4.38 and −7.30 kcal/mol, respectively. These species all have negative overall energies with respect to I-1-S. So, deprotonation is spontaneous in this experimental condition.
14-N2Se2 and 14-N2S2-Ts follow the same mechanism and have similarly favorable reaction barriers and energies: ΔG‡ of ∼5–6 kcal/mol and ΔG° of −5 to −8 kcal/mol; see Figure 3A. With the more favorable I-3-Ts and I-3-Se over I-3-S, their deprotonated I-5 species are still more favorable.
However, due to steric hindrance, it is not possible for PIPES to reach the inward proton (transferred from HNO) in 14-N2S2-Bn for deprotonation at the Bn-substituted N3 site. So, Bn first needs to be rotated away in I-3-Bn via TS-1-Bn to expose the N3 proton (see the optimized structures in Figure S21), which forms I-4-Bn with a low energy cost of 2.79 kcal/mol. This rotation can be seen by a large bond angle increase for ∠CuN3CBn (benzyl carbon bonded to N3) from 118.2° in I-3-Bn to 154.3° in I-4-Bn (Table S66). It then binds with PIPES to form adduct I-5-Bn with subsequent TS-2-Bn to take the proton away to yield I-6-Bn; see Figure 3B for energies and Figure S22 for optimized structures. Compared with other ligands, the deprotonation ΔG‡ of 13.98 kcal/mol from I-3-Bn to TS-2-Bn (Figure 3B) is the highest due to the extra Bn rotation and the more difficult proton abstraction at an EDG-substituted site with an enhanced proton affinity. However, it is still lower than the HNO to NO conversion barrier. The energy of TS-2-Bn from I-1-Bn (−10.91 kcal/mol) is still negative. This step is thermodynamically favorable; see Figure 3B. These results show that all four mixed ligands have spontaneous deprotonations.
Proton or Substituent Rotation
The Cu–N3 bond in I-5-S is still broken (∼3 Å, Table S27) because the remaining proton points toward Cu to intrude the Cu–N3 bonding. Since it is reformed (2.1 Å, Tables S1 and S2) in the final product, a proton rotation is needed to position N3 in a closer distance and have its lone pair in the right orientation to rebond with Cu; see Scheme 3 and the optimized structures in Figure S7. After rotation, I-6-S has Cu–N3 rebonded (2.131 Å, Table S30), which restores the symmetry to the Cu–N1 bond like in a typical trans-III structure. The proton rotation can be seen by the increased ∠CuN3H from 51.7° in I-5-S via 80.5° in TS-2-S (see the change in Figure 4C) to 102.9° in I-6-S and the change of ∠S2CN3H from −53.5° in I-5-S to −110.5° in I-6-S. As shown in Table S29, from I-5-S, the proton rotation barrier is 1.16 kcal/mol and its ΔG° toward I-6-S is −4.97 kcal/mol, indicating a fast step to make it ready for following conformational changes.89−92
Scheme 3. Proton Rotation Step.
Green box represents the rotated proton.
Figure 4.

(A) Gibbs free energies of proton rotation pathways for 14-N2S2, 14-N2Se2, and 14-N2S2-Ts. (B) Gibbs free energies of the second substituent rotation pathway for 14-N2S2-Bn. (C) Optimized transition-state structures with the geometric change of ∠CuN3H for proton rotation from I-5 to TS-2 and ∠CuN3CBn for substituent rotation from I-6 to TS-3 with 14-N2S2-Bn. Atom color scheme: C, cyan; N, blue; O, red; S, yellow; Se, orange; H, gray; Cu, coral.
Additional calculations were carried out to see whether the conformation change could proceed without proton rotation. However, as shown in Table S20 for the first conformation change transition state TS-3-S (which is sulfur inversion; see more discussion of the conformation change pathway later), without this proton rotation step, the barrier becomes much higher by 5 kcal/mol, supporting the necessity of proton rotation.
We then examined this step for 14-N2Se2 and 14-N2S2-Ts (optimized structures in Figures S12 and S17, respectively). As shown in Tables S42 and S55, they have similar barriers (∼1 kcal/mol) and similar or more favorable reaction energies (∼− 4 or −8 kcal/mol). Overall, these two new systems have more favorable proton rotation than the 14-N2S2 ligand; see lower energy levels in Figure 4A.
Due to Bn substitution on N3, there is no proton attached to N3 at I-6-Bn and thus no proton rotation step. However, the Bn substituent is still bent away. So, I-6-Bn then undergoes a Bn rotation via TS-3-Bn to restore its normal stable conformation in I-7-Bn with a significant ΔG° drop of ∼11 kcal/mol and relatively small ΔG‡ of 2.25 kcal/mol (Table S71), which also reforms the Cu–N3 bond with a length of 2.237 Å (Table S72). This is the second substituent rotation, which can be seen by the decreased ∠CuN3CBn (CBn is the benzylic C) from 152.3° in I-6-Bn via 147.8° in TS-3-Bn (see the change in Figure 4C) to 106.9° in I-7-Bn.
Overall, this proton or substituent rotation step for all of these ligands is more favorable than the previous proton abstraction step and new ligands have more negative energies than 14-N2S2.
Conformation Change via Deprotonation
The conformation change mechanisms toward tetrahedral for CuI cyclams and S-containing derivatives89,91,92 were reported before, involving sequential inversion of certain coordinated atoms depending on the starting conformation. While S inversion can occur directly, the inversion of a nitrogen center requires deprotonation and then reprotonation before/after the inversion. However, there are no computational studies of this mechanism. So, we first investigated it for 14-N2S2.
As shown in Scheme 1B, in trans-III conformation, one pair of N and S points up and the other pair points down, while both nitrogens point up and both sulfurs point down in the tetrahedral conformation. Thus, one N and one S inversions are needed for this conformation change. For 14-N2S2, specifically, N1 and S4 are to be inverted; see Scheme 4.
Scheme 4. Nitrogen Inversion via Deprotonation.
Color spheres represent the involved proton.
Based on the proposed nitrogen inversion mechanism for metal cyclams and derivatives,89,91,92 N1’s proton in I-6-S is deprotonated by a base via a transition state TS-3*-S to form I-7*-S. Afterward, N1 inversion occurs when a proton is added to the opposite side to generate I-8*-S to complete the needed nitrogen site structural change, which then undergoes a sulfur inversion to form the final P–S. Since the experimental reaction occurred at pH 7, PIPES was again used as a deprotonation agent. However, the reaction energy to form I-7*-S from I-6-S is 33 kcal/mol (Table S21), which is too high to be thermodynamically favorable. Because a reaction barrier is always higher than the reaction energy, it suggests that this deprotonation is also kinetically not feasible. In fact, all efforts to locate TS-3*-S failed. So, these results suggest that this mechanism does not operate here.
Conformation Change via Rotation
We then proposed a rotation pathway for the needed N1 and S4 inversions. To determine the inversion sequence, we calculated I-7-S and I-7′-S with sulfur and nitrogen inverted first, respectively. The S first-inverted intermediate I-7-S has a lower energy by −10.85 kcal/mol (Table S21) than the N first-inverted I-7′-S and is the only thermodynamically favorable one. So, our proposed pathway in Scheme 5 has S inversion first. For S4 inversion (the cyan box part in TS-3-S, Scheme 5), the ΔG‡ for TS-3-S from I-6-S is 13.81 kcal/mol (Figure 5A and Table S32), which is below barriers of some room-temperature reactions85−87 and thus supports its kinetical feasibility. This S4 inversion can be seen by the dihedral angle of ∠N1CuS4C1 (see atom labels in I-6-SFigure 5B) from positive 25.6° (C1 below the N1–Cu–S4 plane) to 6.9° (C1 almost in that plane) to negative −6.8° (C1 above that plane); see the data in Table S33.
Scheme 5. Conformation Change Pathway via Rotation.
Color boxes represent the rotation site.
Figure 5.

(A) Gibbs free energies of the rotational conformation change pathway with Cu macrocyclic complexes: 14-N2S2 and 14-N2Se2. (B) Optimized structures of the species involved in the conformation change pathway in the 14-N2S2 system (the 14-N2Se2 system follows the same reaction pathway; see Figure S13). Atom color scheme: C, cyan; N, blue; O, red; S, yellow; H, gray; Cu, coral. Key atoms’ labels (here, H is the N1’s proton that points down) are shown in the optimized structure of I-6-S.
I-7-S then undergoes a small cost (2.56 kcal/mol) carbon bridge rotation to form I-8-S, where the carbon chain between N1 and S4 (the green box part in I-7-S, Scheme 5; carbons C1 and C2 labeled in I-6-SFigure 5B) has changed to the same conformation in P–S; see the optimized structures in Figure 5B. This C1–C2 orientation change is shown by the dihedral angle of ∠N1CuC1C2 from a negative value of −27.8° in I-7-S to a positive number of 12.5° in I-8-S; see Table S33 and Figure 5B.
Then, N1 inversion starts. Because N1’s H points down in I-6-S and should point up like N3′s in P–S, N1 inversion requires simultaneous H rotation, which takes two steps. First, this H (the pink box part in TS-4-S, Scheme 5) rotates from pointing away from Cu in I-8-S to toward Cu in I-9-S via TS-4-S. This is indicated by the gradual decrease of ∠HN1Cu from 103.0° via 71.7 to 61.3° in this step; see Table S33 and Figure 5B. It has a small barrier of 6.46 kcal/mol (Figure 5A). This proton rotation disrupts the Cu–N1 bond: the bond length changes from 2.177 via 2.778 to 3.295 Å from I-8-S via TS-4-S to I-9-S (Table S33), which allows more flexibility for the following concerted nitrogen and hydrogen inversion from pointing down to up via TS-5-S to I-10-S. The N1 inversion can be seen by a ∠C2C3CuN1 change from +31.8° in I-9-S to −31.7° in I-10-S, corresponding to the N1 change from below the C2–C3–Cu plane to above; see Table S33 and Figure 5B. Accordingly, its proton inversion is shown by a positive ∠CuC2C3H value of 67.2° in I-9-S to a negative value of −68.6° in I-10-S, reflecting the H position from below the Cu–C2–C3 plane to above. The Cu–N1 bond reforms (2.114 Å) in I-10-S. Now, the only difference between I-10-S and P–S is the carbon bridge between N1 and S2: the relative heights for the three carbons from N1 to S2 are high–low–high (see the blue box in I-10-S in Scheme 5), which are changed to low–high–low in P–S. This step is energetically downhill with −2.76 kcal/mol.
As shown in Figure 5A, in this rotation pathway from I-6-S to P–S, the rate-determining step (RDS) is the first S rotation (TS-3-S) with a small barrier of ∼14 kcal/mol. This conformation change pathway is also thermodynamically favorable, with a reaction energy of −7.37 kcal/mol. With the starting HNO to NO conversion reaction, the overall reaction energy is −38.17 kcal/mol (Figure 5A). So, this mechanism is the first plausible one to explain the experimentally observed facile conformation change for 14-N2S2.
As 14-N2Se2 has the same starting conformation as 14-N2S2, it follows the same pathway; see the blue energy levels in Figure 5A and optimized structures in Figure S13. Compared to 14-N2S2, the ΔG‡ from I-6-Se (19.18 kcal/mol) is higher but still within the range of room-temperature reactions,85−87 while the reaction energy (−9.16 kcal/mol) is lower. This may be a result of the larger size of Se, which increases the strain (Cu–N1/N3 bond lengths elongated by ca. 0.05 Å, Tables S33 and S46) in TS-3-Se for Se inversion but allows more space for subsequent N inversion to make them more favorable; see the lower (blue vs red) energy levels in Figure 5A.
Pathways for 14-N2S2-Ts and 14-N2S2-Bn are different because their Cu complexes are in trans-I conformation, not trans-III for 14-N2S2 and 14-N2Se2. Structural differences between trans-I and tetrahedral lie in both sulfurs, which point up and down, respectively (Scheme 1B). Hence, the conformation change involves two S inversions here, which take fewer steps (a favorable reaction feature) than the nonsubstituted systems; see Figures 6 and 7. Since both S sites need to be inverted and they are symmetrically distributed on two sides of the substituent, we chose the sulfur with a relatively longer Cu–S bond length (slightly more flexible for structural changes) to be first inverted. Barriers are similar (<1–2 kcal/mol differences) for two S sites due to symmetry; see Figures 6 and 7.
Figure 6.

(A) Gibbs free energies of the rotational conformation change pathway in 14-N2S2-Ts. (B) Optimized structures of species involved in the conformation change pathway in the 14-N2S2-Ts system. Atom color scheme: C, cyan; N, blue; O, red; S, yellow; H, gray; Cu, coral.
Figure 7.

(A) Gibbs free energies of the rotational conformation change pathway in 14-N2S2-Bn. (B) Optimized structures of species involved in the conformation change pathway in the 14-N2S2-Bn system. Atom color scheme: C, cyan; N, blue; O, red; S, yellow; H, gray; Cu, coral.
For 14-N2S2-Ts, the RDS has a ΔG‡ of 15.56 kcal/mol from I-6-Ts; see Table S58. The EWG substituent on N1 destabilized the Cu–N1 bond, which is longer than the Cu–N3 bond by ca. 0.5 Å in I-6-Ts. The S inversion further weakens it to be broken in TS-3-Ts and restored in P-Ts; see the optimized structures in Figure S18 and geometric data in Table S59. The two S inversions can be seen by the sign changes of corresponding dihedral angles of ∠C5C6CuS2 and ∠C10C1CuS4 (see atom labels in Figure 6), respectively, to reflect that S2 and S4 change from above their respective planes to below; see Table S59. The conformation change alone has a negative ΔG° of −1.39 kcal/mol. The 14-N2S2-Ts system overall has a reaction energy of −43.39 kcal/mol, which is ca. 5 kcal/mol lower than that of 14-N2S2.
For 14-N2S2-Bn, the starting trans-I structure for conformation change is I-7-Bn with a normal Bn conformation. The two S inversions can be seen by the sign changes of corresponding dihedral angles of ∠C5C6CuS2 and ∠C10C1CuS4 (see atom labels in Figure 7), respectively, to reflect that S2 and S4 change from above their respective planes to below; see Table S75. As seen from Table S74 (optimized structures in Figure S24), the RDS has a ΔG‡ of 19.62 kcal/mol, which is higher than other ligands but still in the range for room-temperature reactions. Its conformation change ΔG° of −1.77 kcal/mol is like that for 14-N2S2-Ts. Overall, the final tetrahedral product formation energy of −38.82 kcal/mol is slightly lower than that of 14-N2S2. So, the new mechanism is favorable for all four ligands.
Comparisons with Other Systems
The HNO reaction with the oxidized metal center to form a reduced metal and a proton transfer to the system is a common feature of the metal-based HNO reaction sensors or probes.22−34 We have studied a number of such reactions including both Cu-based systems and heme proteins,59−61 which all exhibit the PCET mechanism. Table 2 presents the values of ΔG‡’s, ΔG°’s, and ΔRM···N’s (change of the distance between the metal and the coordination nitrogen that accepts HNO’s proton from the transition state to its preceding intermediate, i.e., TS vs I-1) of this reaction from the HNO-bound complex to the NO-released stage, i.e., from I-1 to I-3 and NO.
Table 2. Comparisons of Key Energies and Geometric Parameters in the HNO to NO Conversion Pathways.
| system | ΔG‡ (kcal/mol) | ΔG° (kcal/mol) | ΔRM···Na (Å) | refsb |
|---|---|---|---|---|
| [Cu(14-N4)] | 26.86 | –0.22 | 1.037 | t.w. |
| [Cu(14-N2S2)] | 21.10 | –18.53 | 1.078 | t.w. |
| [Cu(14-N2Se2)] | 19.14 | –21.62 | 1.201 | t.w. |
| [Cu(14-N2S2-Ts)] | 20.92 | –26.14 | 0.995 | t.w. |
| [Cu(14-N2S2-Bn)] | 16.57 | –25.20 | 1.102 | t.w. |
| CuDHX1 | 20.59 | –17.22 | 1.025 | (59) |
| Cu-Fl | 18.84 | –11.21 | 1.050 | (59) |
| Cu-Ts | 23.53 | –11.76 | 0.932 | (59) |
| Cu-Me4 | 19.02 | –14.29 | 0.971 | (59) |
| CuZnSOD | 10.98 | –42.66 | 0.374 | (61) |
| CuZnSOD model | 12.12 | –12.80 | 1.076 | (69) |
| myoglobin | –1.92 | –10.07c | N.A. | (60) |
| catalase | –1.18 | –11.13c | N.A. | (60) |
N.A. means not applicable.
Reference number. t.w. means this work.
The reaction energy is up to the NO-bound reduced metal, not the NO-released system.
It can be seen that the Cu cyclam derivative systems with coordination S/Se atoms studied in this work have a relatively lower reaction barrier range (17–21 kcal/mol) and lower reaction energy range of −19 to −26 kcal/mol than Cu cyclams (19–27 kcal/mol and ∼0 to −17 kcal/mol, respectively) reported recently.59 These data clearly show that the mixed S/Se–N coordination systems have better HNO reactivities, which were found to originate from at least three favorable factors described in detail in the HNO Reactivity Section under Results and Discussion Section: (1) less steric hindrance due to the enlarged coordinated sphere facilitates HNO binding; (2) the enlarged coordination sphere weakens the Cu–N3 bond for lower cleavage cost; and (3) the less electronegative S/Se vs N renders a significantly less positively charged Cu to reduce its attraction with the negatively charged N3 to decrease the bond cleavage cost.
However, the CuZnSOD (SOD, superoxide dismutase) protein has the lowest barrier and reaction energy among all Cu systems (see Table 2), which was found to be a result of a more flexible nonplanar coordination environment in the protein to lower the Cu–N bond-breaking cost and a stronger proton acceptor due to a negatively charged coordination histidine ligand.61 The previously developed CuZnSOD model for HNO detection with a similar coordination environment but with a neutral ligand has higher ΔG‡ and ΔG°,69 see Table 2.
In contrast with these Cu systems, the heme proteins (myoglobin and catalase) have basically barrierless (ΔG‡’s are in Table 2; ΔE‡’s are 0.25 and 0.83 kcal/mol, respectively) HNO to NO conversion because HNO’s proton is transferred to a protein residue that has strong hydrogen bonding with it before the reaction.60 Thus, the heme systems have no metal-coordination-N bond breaking to save the energy cost involved in the Cu systems. However, the strong binding of NO to ferrous heme centers makes the final systems remain nitrosylated and thus difficult to recycle for reuse.
Overall, the introduction of sulfur (and selenium) as coordination atoms in Cu cyclam derivatives leads to stable tetrahedral structures of the reduced species, a unique feature the all-nitrogen coordination system does not have (e.g., [Cu(14-N4)] still prefers the trans-III conformation in the reduced form, Table S3). This stable reduced species is also free of anything from HNO after the reaction, which makes it clean to be recycled back for repeated uses. Therefore, only the Cu systems with the mixed S/Se coordination atoms have subsequent conformation change pathways, which were found to be kinetically feasible and thermodynamically favorable from the above-mentioned studies.
Conclusions
Quantum chemical calculations accurately (∼2% error) predicted experimental structures of CuII/CuI complexes of the mixed thia/aza system and offered the first theoretical insights into the selective stabilization of the tetrahedral conformation for the reduced system. Results also reproduced the experimental facile HNO reactivity of 14-N2S2 vs inactivity for 14-N4 and the spontaneous conformation change to the final product for 14-N2S2. The complete pathway of HNO to NO conversion, deprotonation, proton or substituent rotation, and conformation change via rotation was uncovered for this reusable HNO sensor. A novel conformation change mechanism was discovered, which is more favorable than the previously proposed deprotonation–inversion–protonation pathway for the room-temperature, neutral-pH condition of the experimental or physiological systems. The whole reaction pathway was then examined for three new designs, which were found to have improved reactivities due to the introduction of softer and larger coordinate atoms and electron-withdrawing or electron-donating substituents. The favorable results with substituted mixed thia/aza systems are particularly interesting, which paves new venues to incorporate fluorophores into reusable HNO sensors that have not been discovered before. Overall, this first and systematic mechanistic study of a reusable HNO sensor uncovered numerous insightful results to facilitate HNO sensor development for broad biochemical and biomedical uses.
Acknowledgments
This work was supported by an NIH grant GM085774 to Y.Z.
Supporting Information Available
The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.inorgchem.3c04506.
Details of all the studied structures and pathways with various geometric and electronic parameters and additional figures, together with absolute energies and coordinates of favorable species (PDF)
Author Contributions
Y.Z. conceived the idea and designed the research. J.M.C., D.B., and V.N. conducted the computational studies. All authors participated in the data analyses and preparation of data tables and figures. J.M.C. and Y.Z. wrote the manuscript together with input from all authors. Y.Z. supervised the research.
The authors declare no competing financial interest.
Supplementary Material
References
- Fukuto J. M. A recent history of nitroxyl chemistry, pharmacology and therapeutic potential. Br. J. Pharmacol. 2019, 176 (2), 135–146. 10.1111/bph.14384. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Fukuto J. M.; Bartberger M. D.; Dutton A. S.; Paolocci N.; Wink D. A.; Houk K. N. The Physiological Chemistry and Biological Activity of Nitroxyl (HNO): The Neglected, Misunderstood, and Enigmatic Nitrogen Oxide. Chem. Res. Toxicol. 2005, 18, 790–801. 10.1021/tx0496800. [DOI] [PubMed] [Google Scholar]
- Bartberger M. D.; Fukuto J. M.; Houk K. N. On the acidity and reactivity of HNO in aqueous solution and biological systems. Proc. Natl. Acad. Sci. U.S.A. 2001, 98 (5), 2194–2198. 10.1073/pnas.041481598. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Irvine J. C.; Ritchie R. H.; Favaloro J. L.; Andrews K. L.; Widdop R. E.; Kemp-Harper B. K. Nitroxyl (HNO): the Cinderella of the nitric oxide story. Trends Pharmacol. Sci. 2008, 29 (12), 601–608. 10.1016/j.tips.2008.08.005. [DOI] [PubMed] [Google Scholar]
- Fukuto J. M.; Dutton A. S.; Houk K. N. The chemistry and biology of nitroxyl (HNO): A chemically unique species with novel and important biological activity. ChemBioChem 2005, 6, 612–619. 10.1002/cbic.200400271. [DOI] [PubMed] [Google Scholar]
- Pagliaro P.; Mancardi D.; Rastaldo R.; Penna C.; Gattullo D.; Miranda K. M.; Feelisch M.; Wink D. A.; Kass D. A.; Paolocci N. Nitroxyl affords thiol-sensitive myocardial protective effects akin to early preconditioning. Free Radical Biol. Med. 2003, 34 (1), 33–43. 10.1016/S0891-5849(02)01179-6. [DOI] [PubMed] [Google Scholar]
- Bullen M. L.; Miller A. A.; Dharmarajah J.; Drummond G. R.; Sobey C. G.; Kemp-Harper B. K. Vasorelaxant and antiaggregatory actions of the nitroxyl donor isopropylamine NONOate are maintained in hypercholesterolemia. Am. J. Physiol.: Heart Circ. Physiol. 2011, 301 (4), H1405–H1414. 10.1152/ajpheart.00489.2011. [DOI] [PubMed] [Google Scholar]
- Bullen M. L.; Miller A. A.; Andrews K. L.; Irvine J. C.; Ritchie R. H.; Sobey C. G.; Kemp-Harper B. K. Nitroxyl (HNO) as a Vasoprotective Signaling Molecule. Antioxid. Redox Signalling 2011, 14 (9), 1675–1686. 10.1089/ars.2010.3327. [DOI] [PubMed] [Google Scholar]
- Fukuto J. M.; Bianco C. L.; Chavez T. A. Nitroxyl (HNO) signaling. Free Radical Biol. Med. 2009, 47 (9), 1318–1324. 10.1016/j.freeradbiomed.2009.06.014. [DOI] [PubMed] [Google Scholar]
- Shoman M. E.; Aly O. M. Nitroxyl (HNO): A Possible Strategy for Fighting Cancer. Curr. Top. Med. Chem. 2016, 16 (22), 2464–2470. 10.2174/1568026616666160212123006. [DOI] [PubMed] [Google Scholar]
- Doctorovich F.; Bikiel D.; Pellerino J.; Suarez S. A.; Larsen A.; Marti M. A. Nitroxyl (azanone) trapping by metalloporphyrins. Coord. Chem. Rev. 2011, 255, 2764–2784. 10.1016/j.ccr.2011.04.012. [DOI] [Google Scholar]
- Martí M. A.; Bari S. E.; Estrin D. A.; Doctorovich F. Discrimination of Nitroxyl and nitric oxide by water-soluble Mn(III) porphyrins. J. Am. Chem. Soc. 2005, 127, 4680–4684. 10.1021/ja044632n. [DOI] [PubMed] [Google Scholar]
- Suarez S. A.; Muñoz M.; Alvarez L.; Venâncio M. F.; Rocha W. R.; Bikiel D. E.; Marti M. A.; Doctorovich F. HNO Is Produced by the Reaction of NO with Thiols. J. Am. Chem. Soc. 2017, 139 (41), 14483–14487. 10.1021/jacs.7b06968. [DOI] [PubMed] [Google Scholar]
- Donzelli S.; Espey M. G.; Thomas D. D.; Mancardi D.; Tocchetti C. G.; Ridnour L. A.; Paolocci N.; King S. B.; Miranda K. M.; Lazzarino G.; et al. Discriminating formation of HNO from other reactive nitrogen oxide species. Free Radical Biol. Med. 2006, 40 (6), 1056–1066. 10.1016/j.freeradbiomed.2005.10.058. [DOI] [PubMed] [Google Scholar]
- Donzelli S.; Espey M. G.; Flores-Santana W.; Switzer C. H.; Yeh G. C.; Huang J.; Stuehr D. J.; King S. B.; Miranda K. M.; Wink D. A. Generation of nitroxyl by heme protein-mediated peroxidation of hydroxylamine but not N-hydroxy-L-arginine. Free Radical Biol. Med. 2008, 45, 578–584. 10.1016/j.freeradbiomed.2008.04.036. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Reisz J. A.; Klorig E. B.; Wright M. W.; King S. B. Reductive Phosphine-Mediated Ligation of Nitroxyl (HNO). Org. Lett. 2009, 11 (13), 2719–2721. 10.1021/ol900914s. [DOI] [PubMed] [Google Scholar]
- Reisz J. A.; Zink C. N.; King S. B. Rapid and Selective Nitroxyl (HNO) Trapping by Phosphines: Kinetics and New Aqueous Ligations for HNO Detection and Quantitation. J. Am. Chem. Soc. 2011, 133, 11675–11685. 10.1021/ja203652z. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Cline M. R.; Tu C.; Silverman D. N.; Toscano J. P. Detection of nitroxyl (HNO) by membrane inlet mass spectrometry. Free Radical Biol. Med. 2011, 50, 1274–1279. 10.1016/j.freeradbiomed.2011.02.008. [DOI] [PubMed] [Google Scholar]
- Suárez S. A.; Fonticelli M. H.; Rubert A. A.; de la Llave E.; Scherlis D.; Salvarezza R. C.; Martí M. A.; Doctorovich F. A Surface Effect Allows HNO/NO Discrimination by a Cobalt Porphyrin Bound to Gold. Inorg. Chem. 2010, 49 (15), 6955–6966. 10.1021/ic1007022. [DOI] [PubMed] [Google Scholar]
- Suárez S. A.; Bikiel D. E.; Wetzler D. E.; Martí M. A.; Doctorovich F. Time-Resolved Electrochemical Quantification of Azanone (HNO) at Low Nanomolar Level. Anal. Chem. 2013, 85 (21), 10262–10269. 10.1021/ac402134b. [DOI] [PubMed] [Google Scholar]
- Gallego C. M.; Mazzeo A.; Vargas P.; Suárez S.; Pellegrino J.; Doctorovich F. Azanone (HNO): generation, stabilization and detection. Chem. Sci. 2021, 12 (31), 10410–10425. 10.1039/D1SC02236A. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Rosenthal J.; Lippard S. J. Direct detection of nitroxyl in aqueous solution using a tripodal copper (II) BODIPY complex. J. Am. Chem. Soc. 2010, 132 (16), 5536–5537. 10.1021/ja909148v. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Islam A. S. M.; Sasmal M.; Maiti D.; Dutta A.; Ganguly S.; Katarkar A.; Gangopadhyay S.; Ali M. Phenazine-Embedded Copper(II) Complex as a Fluorescent Probe for the Detection of NO and HNO with a Bioimaging Application. ACS Appl. Bio Mater. 2019, 2 (5), 1944–1955. 10.1021/acsabm.9b00010. [DOI] [PubMed] [Google Scholar]
- Zhao X.; Gao C.; Li N.; Liu F.; Huo S.; Li J.; Guan X.; Yan N. BODIPY based fluorescent turn-on sensor for highly selective detection of HNO and the application in living cells. Tetrahedron Lett. 2019, 60 (21), 1452–1456. 10.1016/j.tetlet.2019.04.049. [DOI] [Google Scholar]
- Maiti D.; Islam A. S. M.; Dutta A.; Sasmal M.; Prodhan C.; Ali M. Dansyl-appended CuII-complex-based nitroxyl (HNO) sensing with living cell imaging application and DFT studies. Dalton Trans. 2019, 48 (8), 2760–2771. 10.1039/C8DT04564J. [DOI] [PubMed] [Google Scholar]
- Palanisamy S.; Wang Y.-L.; Chen Y.-J.; Chen C.-Y.; Tsai F.-T.; Liaw W.-F.; Wang Y.-M. In Vitro and in Vivo Imaging of Nitroxyl with Copper Fluorescent Probe in Living Cells and Zebrafish. Molecules 2018, 23 (10), 2551 10.3390/molecules23102551. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Dutta A.; Alam R.; Islam A. S. M.; Dutta A.; Ali M. A dual response fluorescent sensor for HNO and S2– ions using a Cu(ii) complex based probe assisted by detailed DFT studies. Dalton Trans. 2018, 47 (33), 11563–11571. 10.1039/C8DT02784F. [DOI] [PubMed] [Google Scholar]
- Sun X.; Kim G.; Xu Y.; Yoon J.; James T. D. A Water-Soluble Copper(II) Complex for the Selective Fluorescence Detection of Nitric Oxide/Nitroxyl and Imaging in Living Cells. ChemPlusChem 2016, 81 (1), 30–34. 10.1002/cplu.201500436. [DOI] [PubMed] [Google Scholar]
- Loas A.; Radford R. J.; Liang A. D.; Lippard S. J. Solid-phase synthesis provides a modular, lysine-based platform for fluorescent discrimination of nitroxyl and biological thiols. Chem. Sci. 2015, 6 (7), 4131–4140. 10.1039/C5SC00880H. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wrobel A. T.; Johnstone T. C.; Liang A. D.; Lippard S. J.; Rivera-Fuentes P. A Fast and Selective Near-Infrared Fluorescent Sensor for Multicolor Imaging of Biological Nitroxyl (HNO). J. Am. Chem. Soc. 2014, 136, 4697–4705. 10.1021/ja500315x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhou Y.; Yao Y.-W.; Li J.-Y.; Yao C.; Lin B.-P. Nitroxyl induced fluorescence enhancement via reduction of a copper(II) coumarin-ester complex: Its application for bioimaging in vivo. Sens. Actuators, B 2012, 174, 414–420. 10.1016/j.snb.2012.08.067. [DOI] [Google Scholar]
- Zhou Y.; Liu K.; Li J.-Y.; Fang Y.; Zhao T.-C.; Yao C. Visualization of Nitroxyl in Living Cells by a Chelated Copper(II) Coumarin Complex. Org. Lett. 2011, 13 (6), 1290–1293. 10.1021/ol103077q. [DOI] [PubMed] [Google Scholar]
- Smulik-Izydorczyk R.; Dębowska K.; Pięta J.; Michalski R.; Marcinek A.; Sikora A. Fluorescent probes for the detection of nitroxyl (HNO). Free Radical Biol. Med. 2018, 128, 69–83. 10.1016/j.freeradbiomed.2018.04.564. [DOI] [PubMed] [Google Scholar]
- Dong B.; Kong X.; Lin W. Reaction-Based Fluorescent Probes for the Imaging of Nitroxyl (HNO) in Biological Systems. ACS Chem. Biol. 2018, 13 (7), 1714–1720. 10.1021/acschembio.7b00901. [DOI] [PubMed] [Google Scholar]
- Xu J.; Bai Y.; Ma Q.; Sun J.; Tian M.; Li L.; Zhu N.; Liu S. Ratiometric Determination of Nitroxyl Utilizing a Novel Fluorescence Resonance Energy Transfer-Based Fluorescent Probe Based on a Coumarin-Rhodol Derivative. ACS Omega 2022, 7 (6), 5264–5273. 10.1021/acsomega.1c06403. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Niu H.; Mi X.; Hua X.; Zhang Y.; Zhai Y.; Qin F.; Ye Y.; Zhao Y. A bifunctional fluorescent probe based on “AND logic” for the simultaneous recognition of H2S/HNO and its bioimaging applications. Anal. Chim. Acta 2022, 1192, 339341 10.1016/j.aca.2021.339341. [DOI] [PubMed] [Google Scholar]
- Zhang C.; He F.; Huang F.; Xu J.; Hu Y. A fluorescent probe with low-background for visualization of nitroxyl in living cells and zebrafish. Dyes Pigm. 2021, 185, 108889 10.1016/j.dyepig.2020.108889. [DOI] [Google Scholar]
- Chai Z.; Liu D.; Li X.; Zhao Y.; Shi W.; Li X.; Ma H. A tumor-targeted near-infrared fluorescent probe for HNO and its application to the real-time monitoring of HNO release in vivo. Chem. Commun. 2021, 57 (41), 5063–5066. 10.1039/D1CC01462E. [DOI] [PubMed] [Google Scholar]
- Li H.; Wang C.; Cai L.; Yu X.; Wu L.; Yuan N.; Zhu Y.; Jia N.; James T. D.; Huang C. Versatile Ratiometric Fluorescent Probe Based on the Two-Isophorone Fluorophore for Sensing Nitroxyl. Ind. Eng. Chem. Res. 2021, 60 (44), 15913–15920. 10.1021/acs.iecr.1c02020. [DOI] [Google Scholar]
- Li Z.; Li J.; Zhang D.; Zhu X.; Ye Y.; Zhao Y. A novel dual-channel fluorescent probe for nitroxyl detection and its application in HepG-2 cells. Sens. Actuators, B 2020, 312, 127944 10.1016/j.snb.2020.127944. [DOI] [Google Scholar]
- Liu Z.; Sun Q. A near-infrared fluorescent probe for imaging of nitroxyl in living cells. Spectrochim. Acta, Part A 2020, 241, 118680 10.1016/j.saa.2020.118680. [DOI] [PubMed] [Google Scholar]
- Palanisamy S.; Chen L.-F.; Tzou S.-C.; Wang Y.-M. Near-infrared templated fluorescent probe for nitroxyl: Selective and sensitive turn-on detection in living cells. Sens. Actuators, B 2020, 310, 127839 10.1016/j.snb.2020.127839. [DOI] [Google Scholar]
- Peng S.; Li Z.; Zhang Y.; Cao W.; Liu J.; Zhu W.; Ye Y. A two-photon fluorescent probe for HNO rapid visualization in endoplasmic reticulum. Sens. Actuators, B 2020, 317, 128211 10.1016/j.snb.2020.128211. [DOI] [Google Scholar]
- Zhang H.; Qiao Z.; Wei N.; Zhang Y.; Wang K. A rapid-response and near-infrared fluorescent probe for imaging of nitroxyl in living cells. Talanta 2020, 206, 120196 10.1016/j.talanta.2019.120196. [DOI] [PubMed] [Google Scholar]
- Wei C.; Wang X.; Li X.; Jia X.; Hao X.; Zhang J.; Zhang P.; Li X. An isophorone-fused near-infrared fluorescent probe with a large Stokes shift for imaging endogenous nitroxyl in living cells and zebrafish. Spectrochim. Acta, Part A 2020, 227, 117765 10.1016/j.saa.2019.117765. [DOI] [PubMed] [Google Scholar]
- Wang T.; Chai Y.; Chen S.; Yang G.; Lu C.; Nie J.; Ma C.; Chen Z.; Sun Q.; Zhang Y.; Ren J.; Wang F.; Zhu W.-H. Near-infrared fluorescent probe for imaging nitroxyl in living cells and zebrafish model. Dyes Pigm. 2019, 166, 260–265. 10.1016/j.dyepig.2019.03.013. [DOI] [Google Scholar]
- An W.; Ryan L. S.; Reeves A. G.; Bruemmer K. J.; Mouhaffel L.; Gerberich J. L.; Winters A.; Mason R. P.; Lippert A. R. A Chemiluminescent Probe for HNO Quantification and Real-Time Monitoring in Living Cells. Angew. Chem., Int. Ed. 2019, 58 (5), 1361–1365. 10.1002/anie.201811257. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zheng K.; Chen H.; Fang S.; Wang Y. A ratiometric fluorescent probe based on a Bodipy-Coumarin conjugate for sensing of nitroxyl in living cells. Sens. Actuators, B 2016, 233, 193–198. 10.1016/j.snb.2016.04.053. [DOI] [Google Scholar]
- Liu P.; Jing X.; Yu F.; Lv C.; Chen L. A near-infrared fluorescent probe for the selective detection of HNO in living cells and in vivo. Analyst 2015, 140 (13), 4576–4583. 10.1039/C5AN00759C. [DOI] [PubMed] [Google Scholar]
- Miao Z.; Reisz J. A.; Mitroka S. M.; Pan J.; Xian M.; King S. B. A selective phosphine-based fluorescent probe for nitroxyl in living cells. Bioorg. Med. Chem. Lett. 2015, 25 (1), 16–19. 10.1016/j.bmcl.2014.11.041. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhang H.; Liu R.; Tan Y.; Xie W. H.; Lei H.; Cheung H.-Y.; Sun H. A FRET-based Ratiometric Fluorescent Probe for Nitroxyl Detection in Living Cells. ACS Appl. Mater. Interfaces 2015, 7 (9), 5438–5443. 10.1021/am508987v. [DOI] [PubMed] [Google Scholar]
- Jing X.; Yu F.; Chen L. Visualization of nitroxyl (HNO) in vivo via a lysosome-targetable near-infrared fluorescent probe. Chem. Commun. 2014, 50 (91), 14253–14256. 10.1039/C4CC07561G. [DOI] [PubMed] [Google Scholar]
- Pino N. W.; Davis J.; Yu Z.; Chan J. NitroxylFluor: A Thiol-Based Fluorescent Probe for Live-Cell Imaging of Nitroxyl. J. Am. Chem. Soc. 2017, 139 (51), 18476–18479. 10.1021/jacs.7b11471. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Johnson G. M.; Chozinski T. J.; Salmon D. J.; Moghaddam A. D.; Chen H. C.; Miranda K. M. Quantitative Detection of Nitroxyl Upon Trapping with Glutathione and Labeling with a Specific Fluorogenic Reagent. Free Radical Biol. Med. 2013, 63, 476–484. 10.1016/j.freeradbiomed.2013.05.011. [DOI] [PubMed] [Google Scholar]
- Kong Y.; Wan X.; Liu Z.; Chen F.; Wu F.; Qin G.; Cao D.; Cui Y. A novel dansyl-naphthalimide fluorescent probe for visualizing nitroxyl (HNO) in biological systems. Sens. Actuators, B 2022, 350, 130852 10.1016/j.snb.2021.130852. [DOI] [Google Scholar]
- Kim S.; Minier M. A.; Loas A.; Becker S.; Wang F.; Lippard S. J. Achieving Reversible Sensing of Nitroxyl by Tuning the Ligand Environment of Azamacrocyclic Copper(II) Complexes. J. Am. Chem. Soc. 2016, 138, 1804–1807. 10.1021/jacs.5b12825. [DOI] [PubMed] [Google Scholar]
- Tonzetich Z. J.; McQuade L. E.; Lippard S. J. Detecting and Understanding the Roles of Nitric Oxide in Biology. Inorg. Chem. 2010, 49 (14), 6338–6348. 10.1021/ic9022757. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kaur A.; Kolanowski J. L.; New E. J. Reversible Fluorescent Probes for Biological Redox States. Angew. Chem., Int. Ed. 2016, 55 (5), 1602–1613. 10.1002/anie.201506353. [DOI] [PubMed] [Google Scholar]
- Shi Y.; Stella G.; Chu J.-M.; Zhang Y. Mechanistic Origin of Favorable Substituent Effects in Excellent Cu Cyclam Based HNO Sensors. Angew. Chem., Int. Ed. 2022, 61 (45), e202211450 10.1002/anie.202211450. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Shi Y.; Zhang Y. Mechanisms of HNO Reactions with Ferric Heme Proteins. Angew. Chem., Int. Ed. 2018, 57, 16654–16658. 10.1002/anie.201807699. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Shi Y.; Michael M. A.; Zhang Y. HNO to NO Conversion Mechanism with Copper Zinc Superoxide Dismutase, Comparison with Heme Protein Mediated Conversions, and the Origin of Questionable Reversibility. Chem. - Eur. J. 2021, 27 (15), 5019–5027. 10.1002/chem.202100015. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Adamo C.; Barone V. Exchange functionals with improved long-range behavior and adiabatic connection methods without adjustable parameters: The mPW and mPW1PW models. J. Chem. Phys. 1998, 108 (2), 664–675. 10.1063/1.475428. [DOI] [Google Scholar]
- Hay P. J.; Wadt W. R. Ab initio effective core potentials for molecular calculations. Potentials for the transition metal atoms scandium to mercury. J. Chem. Phys. 1985, 82 (1), 270–283. 10.1063/1.448799. [DOI] [Google Scholar]
- Wadt W. R.; Hay P. J. Ab initio effective core potentials for molecular calculations. Potentials for main group elements Na to Bi. J. Chem. Phys. 1985, 82 (1), 284–298. 10.1063/1.448800. [DOI] [Google Scholar]
- Hay P. J.; Wadt W. R. Ab initio effective core potentials for molecular calculations. Potentials for K to Au including the outermost core orbitals. J. Chem. Phys. 1985, 82 (1), 299–310. 10.1063/1.448975. [DOI] [Google Scholar]
- Mennucci B.; Tomasi J. Continuum solvation models: A new approach to the problem of solute’s charge distribution and cavity boundaries. J. Chem. Phys. 1997, 106 (12), 5151–5158. 10.1063/1.473558. [DOI] [Google Scholar]
- Frisch M. J.; Trucks G. W.; Schlegel H. B.; Scuseria G. E.; Robb M. A.; Cheeseman J. R.; Scalmani G.; Barone V.; Mennucci B.; Petersson G. A.; Nakatsuji H.; Caricato M.; Li X.; Hratchian H. P.; Izmaylov A. F.; Bloino J.; Zheng G.; Sonnenberg J. L.; Hada M.; Ehara M.; Toyota K.; Fukuda R.; Hasegawa J.; Ishida M.; Nakajima T.; Honda Y.; Kitao O.; Nakai H.; Vreven T.; Montgomery J. A. Jr.; Peralta J. E.; Ogliaro F.; Bearpark M.; Heyd J. J.; Brothers E.; Kudin K. N.; Staroverov V. N.; Keith T.; Kobayashi R.; Normand J.; Raghavachari K.; Rendell A.; Burant J. C.; Iyengar S. S.; Tomasi J.; Cossi M.; Rega N.; Millam J. M.; Klene M.; Knox J. E.; Cross J. B.; Bakken V.; Adamo C.; Jaramillo J.; Gomperts R.; Stratmann R. E.; Yazyev O.; Austin A. J.; Cammi R.; Pomelli C.; Ochterski J. W.; Martin R. L.; Morokuma K.; Zakrzewski V. G.; Voth G. A.; Salvador P.; Dannenberg J. J.; Dapprich S.; Daniels A. D.; Farkas O.; Foresman J. B.; Ortiz J. V.; Cioslowski J.; Fox D. J.. Gaussian Program, Gaussian 09, Revision D.01; Gaussian, Inc.: Wallingford CT, 2013.
- Frisch M. J.; Trucks G. W.; Schlegel H. B.; Scuseria G. E.; Robb M. A.; Cheeseman J. R.; Scalmani G.; Barone V.; Petersson G. A.; Nakatsuji H.; Li X.; Caricato M.; Marenich A. V.; Bloino J.; Janesko B. G.; Gomperts R.; Mennucci B.; Hratchian H. P.; Ortiz J. V.; Izmaylov A. F.; Sonnenberg J. L.; Williams-Young D.; Ding F.; Lipparini F.; Egidi F.; Goings J.; Peng B.; Petrone A.; Henderson T.; Ranasinghe D.; Zakrzewski V. G.; Gao J.; Rega N.; Zheng G.; Liang W.; Hada M.; Ehara M.; Toyota K.; Fukuda R.; Hasegawa J.; Ishida M.; Nakajima T.; Honda Y.; Kitao O.; Nakai H.; Vreven T.; Throssell K.; Montgomery J. A. Jr.; Peralta J. E.; Ogliaro F.; Bearpark M. J.; Heyd J. J.; Brothers E. N.; Kudin K. N.; Staroverov V. N.; Keith T. A.; Kobayashi R.; Normand J.; Raghavachari K.; Rendell A. P.; Burant J. C.; Iyengar S. S.; Tomasi J.; Cossi M.; Millam J. M.; Klene M.; Adamo C.; Cammi R.; Ochterski J. W.; Martin R. L.; Morokuma K.; Farkas O.; Foresman J. B.; Fox D. J.. Gaussian Program, Gaussian 16, Revision C.01.; Gaussian, Inc.: Wallingford CT, 2019.
- Michael M. A.; Pizzella G.; Yang L.; Shi Y.; Evangelou T.; Burke D. T.; Zhang Y. HNO/NO Conversion Mechanisms of Cu-based HNO Probes with Implications for Cu,Zn-SOD. J. Phys. Chem. Lett. 2014, 5, 1022–1026. 10.1021/jz5002902. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Truhlar D. G.; Zhao Y. The M06 suite of density functionals for main group thermochemistry, thermochemical kinetics, noncovalent interactions, excited states, and transition elements: two new functionals and systematic testing of four M06-class functionals and 12 other functionals. Theor. Chem. Acc. 2008, 120, 215–241. 10.1007/s00214-007-0310-x. [DOI] [Google Scholar]
- Chai J.-D.; Head-Gordon M. Long-range corrected hybrid density functionals with damped atom-atom dispersion corrections. Phys. Chem. Chem. Phys. 2008, 10, 6615–6620. 10.1039/b810189b. [DOI] [PubMed] [Google Scholar]
- Abucayon E. G.; Khade R. L.; Powell D. R.; Zhang Y.; Richter-Addo G. B. Hydride Attack on a Coordinated Ferric Nitrosyl: Experimental and DFT Evidence for the Formation of a Heme Model-HNO Derivative. J. Am. Chem. Soc. 2016, 138 (1), 104–107. 10.1021/jacs.5b12008. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Abucayon E. G.; Khade R. L.; Powell D. R.; Shaw M. J.; Zhang Y.; Richter-Addo G. B. Over or under: Hydride attack at the metal versus the coordinated nitrosyl ligand in ferric nitrosyl porphyrins. Dalton Trans. 2016, 45, 18259–18266. 10.1039/C6DT03860C. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Weigend F.; Ahlrichs R. Balanced basis sets of split valence, triple zeta valence and quadruple zeta valence quality for H to Rn: Design and assessment of accuracy. Phys. Chem. Chem. Phys. 2005, 7 (18), 3297–3305. 10.1039/b508541a. [DOI] [PubMed] [Google Scholar]
- Weigend F. Accurate Coulomb-fitting basis sets for H to Rn. Phys. Chem. Chem. Phys. 2006, 8 (9), 1057–1065. 10.1039/b515623h. [DOI] [PubMed] [Google Scholar]
- Wachters A. J. H. Gaussian Basis Set for Molecular Wavefunctions Containing Third-Row Atoms. J. Chem. Phys. 1970, 52 (3), 1033–1036. 10.1063/1.1673095. [DOI] [Google Scholar]
- Yu H. S.; He X.; Li S. L.; Truhlar D. G. MN15: A Kohn–Sham global-hybrid exchange–correlation density functional with broad accuracy for multi-reference and single-reference systems and noncovalent interactions. Chem. Sci. 2016, 7 (8), 5032–5051. 10.1039/C6SC00705H. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Abucayon E. G.; Khade R. L.; Powell D. R.; Zhang Y.; Richter-Addo G. B. Hydride Attack on a Coordinated Ferric Nitrosyl: Experimental and DFT Evidence for the Formation of a Heme Model–HNO Derivative. J. Am. Chem. Soc. 2016, 138 (1), 104–107. 10.1021/jacs.5b12008. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ikeda R.; Soneta Y.; Miyamura K. First X-ray structure of a trans-IV of Ni complex of tetra-azamacrocycles with pendent groups, 1, 8-bis (N, N-dimethylcarbamoylethyl)-1, 5, 8, 12-tetramethyl-1, 4, 8, 11-tetraazacyclotetra-decane. Inorg. Chem. Commun. 2007, 10 (5), 590–592. 10.1016/j.inoche.2007.02.007. [DOI] [Google Scholar]
- Hart S. L.; Haines R. I.; Decken A.; Wagner B. D. Isolation of the trans-I and trans-II isomers of Cu II (cyclam) via complexation with the macrocyclic host cucurbit [8] uril. Inorg. Chim. Acta 2009, 362 (11), 4145–4151. 10.1016/j.ica.2009.06.011. [DOI] [Google Scholar]
- Lu T. H.; Shui W. Z.; Tung S. F.; Chi T. Y.; Liao F. L.; Chung C. S. (Isothiocyanato-κN)(1, 4, 8, 11-tetramethyl-1, 4, 8, 11-tetraazacyclotetradecane-κ4N) copper (II) Perchlorate. Acta Crystallogr., Sect. C: Cryst. Struct. Commun. 1998, 54 (8), 1071–1072. 10.1107/S0108270198001711. [DOI] [Google Scholar]
- Maimon E.; Zilbermann I.; Golub G.; Ellern A.; Shames A. I.; Cohen H.; Meyerstein D. Comproportionation and redox catalyzed isomerization of Cu (II)(1R, 4S, 8R, 11S-1, 4, 8, 11-tetramethyl-1, 4, 8, 11-tetraaza-cyclotetradecane) 2+ in aqueous solutions. Inorg. Chim. Acta 2001, 324 (1), 65–72. 10.1016/S0020-1693(01)00534-5. [DOI] [Google Scholar]
- Lee T. J.; Lee T. Y.; Hong C. Y.; Wu D. T.; Chung C. S. Structure of aqua [(1SR, 4RS, 8SR, 11RS)-1, 4, 8, 11-tetramethyl-1, 4, 8, 11-tetraazacyclotetradecane] copper (II) perchlorate monohydrate. Acta Crystallogr., Sect. C: Cryst. Struct. Commun. 1986, 42 (8), 999–1001. 10.1107/S0108270186093721. [DOI] [Google Scholar]
- Derrick J. S.; Kim Y.; Tak H.; Park K.; Cho J.; Kim S. H.; Lim M. H. Stereochemistry of metal tetramethylcyclam complexes directed by an unexpected anion effect. Dalton Trans. 2017, 46 (39), 13166–13170. 10.1039/C7DT01489A. [DOI] [PubMed] [Google Scholar]
- Dzik W. I.; Xu X.; Zhang X. P.; Reek J. N. H.; de Bruin B. ‘Carbene Radicals’ in CoII(por)-Catalyzed Olefin Cyclopropanation. J. Am. Chem. Soc. 2010, 132, 10891–10902. 10.1021/ja103768r. [DOI] [PubMed] [Google Scholar]
- Kazemi M.; Åqvist J. Chemical reaction mechanisms in solution from brute force computational Arrhenius plots. Nat. Commun. 2015, 6, 7293 10.1038/ncomms8293. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ryu H.; Park J.; Kim H. K.; Park J. Y.; Kim S.-T.; Baik M.-H. Pitfalls in Computational Modeling of Chemical Reactions and How To Avoid Them. Organometallics 2018, 37 (19), 3228–3239. 10.1021/acs.organomet.8b00456. [DOI] [Google Scholar]
- Silverstein T. P.; Heller S. T. pKa Values in the Undergraduate Curriculum: What Is the Real pKa of Water?. J. Chem. Educ. 2017, 94 (6), 690–695. 10.1021/acs.jchemed.6b00623. [DOI] [Google Scholar]
- Kulatilleke C. P.; Goldie S. N.; Heeg M. J.; Ochrymowycz L. A.; Rorabacher D. B. Formation and Dissociation Kinetics and Crystal Structures of Nickel(II)–Macrocyclic Tetrathiaether Complexes in Acetonitrile. Comparison to Nickel(II)–Macrocyclic Tetramines. Inorg. Chem. 2000, 39 (7), 1444–1453. 10.1021/ic991019+. [DOI] [PubMed] [Google Scholar]
- Wijetunge P.; Kulatilleke C. P.; Dressel L. T.; Heeg M. J.; Ochrymowycz L. A.; Rorabacher D. B. Effect of Conformational Constraints on Gated Electron-Transfer Kinetics. 3. Copper(II/I) Complexes with cis- and trans-Cyclopentanediyl-1,4,8,11-tetrathiacyclotetradecane1. Inorg. Chem. 2000, 39 (13), 2897–2905. 10.1021/ic0000909. [DOI] [PubMed] [Google Scholar]
- Galijasevic S.; Krylova K.; Koenigbauer M. J.; Jaeger G. S.; Bushendorf J. D.; Heeg M. J.; Ochrymowycz L. A.; Taschner M. J.; Rorabacher D. B. Chelate ring sequence effects on thermodynamic, kinetic and electron-transfer properties of copper(ii/i) systems involving macrocyclic ligands with S4 and NS3 donor sets. Dalton Trans. 2003, (8), 1577–1586. 10.1039/b300602f. [DOI] [Google Scholar]
- Galijasevic S. Thermodynamic and electron-transfer properties of copper(II/I) polyaminothiaether complexes. Transition Met. Chem. 2013, 38 (8), 883–890. 10.1007/s11243-013-9763-z. [DOI] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.





