Open Access Article
This Open Access Article is licensed under a Creative Commons Attribution-Non Commercial 3.0 Unported Licence

17O NMR spectroscopy-assisted in vitro bioactivity studies of the intermediates formed via Na2S and RSNO cross-linking reactions

Xingyu Zhu and Yin Gao*
Key Laboratory for Molecular Enzymology and Engineering of Ministry of Education, School of Life Sciences, Jilin University, Changchun 130012, China. E-mail: yin.gao@queensu.ca

Received 8th June 2020 , Accepted 22nd October 2020

First published on 29th October 2020


Abstract

The cross-linking reaction between sulfide and S-nitrosothiol moieties has been intensively investigated and thionitrite/thionitrous acid (SNO/HSNO) as well as nitrosopersulfide (SSNO) were reported to be the intermediates that could serve as reservoirs for nitric oxide (NO). However, debate still exists regarding the stability and biological activity of SNO/HSNO and SSNO. In order to investigate the chemical properties and biological activity of SNO and SSNO, we set out to re-characterize the reaction intermediates using UV-Vis and 15N NMR spectroscopy techniques, as well as a new 17O NMR approach. The effects of SNO and SSNO on cellular NO and cGMP levels were assessed via cell culture experiments, and also the effects of SNO and SSNO on cell proliferation, migration, and capillary-like structure formation were evaluated with human umbilical vein endothelial cells (HUVEC). Through this work, the characteristic peaks and half-lives of SNO and SSNO were elucidated under various preparation conditions. The biological assays demonstrated that SSNO increased the cellular NO and cGMP levels and also facilitated cell proliferation, migration and stimulated angiogenesis, while in contrast SNO did not exhibit these effects.


Introduction

Although hydrogen sulfide (H2S) is widely known as a noxious gas, it has also emerged as an important intracellular signal transducer along with nitric oxide (NO) and carbon monoxide (CO), which are involved in many physiological processes.1 H2S exhibits similar biological effects to those of NO in terms of vascular tone regulation and control of blood pressure.2 Both H2S and NO are endogenously synthesized in biological systems through precise enzymatic mechanisms.3 These two gases usually exert similar and partially interdependent biological effects, which can lead to different chemical and biological reactions that attenuate or enhance each other.4 For example, H2S acts as an enhancer of NO to promote vasodilation, but can also reverse the effects of NO to induce vasoconstriction.5,6 The cross-linking reaction between H2S and NO has been well established. Recent studies have shown that polysulfide or thiol species formed from the reaction between H2S and NO were possible H2S-derived signaling molecules.7,8 The interaction of H2S with NO or NO donors can activate neuroendocrine signaling pathways to regulate vasodilation and control meningeal blood flow.9,10

The reaction between sulfide and S-nitrosothiols has been extensively investigated, and thionitrite/thionitrous acid (SNO/HSNO) as well as nitrosopersulfide (SSNO) have been reported to be the intermediates that serve as reservoirs for NO.11 However, the stability and bioactivity of SNO/HSNO and SSNO have been subjects of lively debate11–14 since the characterization of HSNO by electrospray ionization time-of-flight mass spectrometry (ESI-TOF-MS) and 15N NMR spectroscopy under physiological conditions was first reported in 2012.15 In the studies by Filipovic and coworkers, the smallest reported S-nitrosothiol HSNO was prepared via the reaction between S-nitrosoglutathione GSNO/GS15NO and Na2S (at a 1[thin space (1/6-em)]:[thin space (1/6-em)]1 molar ratio) in phosphate buffer at pH = 7.4. The m/z peak of 64 Da that was obtained via positive mode ESI-TOF-MS analysis was attributed to the [HSNO + H+] species. Meanwhile, the 15N NMR spectrum showed a peak at 322 ppm which was reported to be the chemical shift of HSNO/SNO.15 However, Cortese-Krott and coworkers were unable to reproduce the aforementioned MS data nor the 15N NMR spectrum in their lab.12 They found that SNO was very unstable and was immediately replaced by a more stable species, SSNO, which they had reported to actually account for the sustained bioactivity of NO.16–19 However, Filipovic and coworkers stated that the SSNO species was a rather unstable intermediate that was sensitive to light, water and acid, which also could further react with H2S to generate HSNO. They claimed that the HSNO species rather than SSNO, was the one that would actually induce cell signaling events.13,20

Up to now the chemical properties and bioactivity of SNO and SSNO still have not been fully elucidated. Although the absorption spectroscopy of the sulfide and S-nitrosothiols cross-linking reaction intermediates has been intensively studied, these studies are still under debate due to the lack of chemical evidence, and opposing opinions have been raised regarding the absorbance peaks of SNO and SSNO.13,17–19,21,22 In the current study, we endeavored to re-characterize the SNO and SSNO intermediates using UV-Vis and 15N NMR spectroscopy techniques as well as a new 17O NMR approach. In addition, the bioactivity of SNO and SSNO was assessed by using cultured human umbilical vein endothelial cells (HUVEC).

Results and discussion

Fig. 1 shows the UV-Vis spectroscopy spectra of the intermediates that were formed by cross-linking reactions between S-nitroso-N-acetylpenicillamine SNAP and Na2S under aqueous and non-aqueous conditions. As shown in Fig. 1a, immediately after SNAP and Na2S (at a 1[thin space (1/6-em)]:[thin space (1/6-em)]1 ratio) had been mixed in DMSO, the absorption at 340 nm was the only peak that was observed in the initial spectrum (red line). As time progressed, the absorbance of the peak at 340 nm diminished while that of a new peak at 450 nm gradually increased. We assumed that these two peaks respectively represented the intermediates SNO and SSNO, which had formed via the cross-linking reaction based on earlier studies by other researchers.17 To support our hypothesis, we have previously proposed a reaction pathway for the generation of [Fe(CN)5N(O)S]4− and [Fe(CN)5N(O)SS]4− during the Gmelin reaction, in which [Fe(CN)5N(O)S]4− was the first intermediate formed after the mixing of sodium nitroprusside ([Fe(CN)5NO]2−) and Na2S.23 [Fe(CN)5N(O)S]4− subsequently decomposed relatively quickly (with a half-life of 1.5 min), and was replaced by [Fe(CN)5N(O)SS]4− as a relatively stable intermediate.23 For this reason, we proposed that the absorption signal at 340 nm represented SNO, while SSNO exhibited an absorption band at 450 nm.
image file: d0ra05054g-f1.tif
Fig. 1 Reaction of SNAP and Na2S. UV-Vis spectra were recorded every min after the mixing of SNAP with Na2S. The line in black was SNAP, the line in red was the first spectrum that was recorded immediately after mixing (0 min), which was followed in sequence at 1 min intervals by the lines in blue (1 min), pink (2 min), green (3 min), and orange (4 min). (a) Mixing of 1.5 mM SNAP and 1 molar equivalent of Na2S in DMSO. The signal at λmax = 340 nm can be attributed to SNO, while that at λmax = 450 nm corresponds to SSNO. (b) Reaction of 1.5 mM SNAP and 0.5 molar equivalents (0.75 mM) of Na2S in DMSO. (c) Reaction of 1.5 mM SNAP and 1 molar equivalent of Na2S in DMF. (d) Reaction of 1.5 mM SNAP and 0.5 molar equivalents of Na2S in DMF. (e) Reaction of 1.5 mM SNAP and 1 molar equivalent of Na2S in 0.5 M phosphate buffer at pH 7.4. The signal at λmax = 329 nm represents SNO, while that at λmax = 412 nm is attributed to SSNO. The intensity of the SSNO band increased during the first 2 min and subsequently diminished. (f) Reaction of 1.5 mM SNAP and 2 molar equivalents Na2S in 0.5 M phosphate buffer at pH 7.4.

Based on this assumption, in the cases in which equimolar amounts of SNAP and Na2S were mixed in DMSO, all of the SNAP would have been converted to SNO initially, and would subsequently react with HSS to form SSNO. However, once the SNAP[thin space (1/6-em)]:[thin space (1/6-em)]Na2S molar ratio was changed to 1[thin space (1/6-em)]:[thin space (1/6-em)]0.5, peaks at both 340 and 450 nm were observed immediately after mixing, and grew weaker as time progressed. This could be due to the excess of SNAP that might have reacted with both HS and HSS to generate SNO and SSNO simultaneously, or at least the two peaks were visible at the same time (Fig. 1b). In comparison, the reaction that was performed in DMF exhibited similar spectra with those recorded in DMSO, as shown in Fig. 1c and d. In contrast, when SNAP and Na2S were mixed in phosphate buffer at pH 7.4, the absorption peak representing SNO shifted from 340 to 329 nm, while that corresponding to SSNO had shifted from 450 to 412 nm (Fig. 1e). Additionally, a higher molar equivalent of Na2S was required to convert all of the SNAP to SNO than was needed for the corresponding reactions performed in DMSO and DMF, since HS behaves as a stronger nucleophile in a non-aqueous solvent than in aqueous environments.11,24 Therefore, the spectra observed during the formation of SNO and SSNO via the mixing of equimolar amounts of SNAP and Na2S in phosphate buffer were similar to those recorded during the formation of SNO and SSNO (1[thin space (1/6-em)]:[thin space (1/6-em)]0.5 molar ratio, SNAP[thin space (1/6-em)]:[thin space (1/6-em)]Na2S) in non-aqueous solvents (Fig. 1b and d).

When 2 molar equivalents of Na2S were mixed with one equivalent of SNAP in phosphate buffer, a weak absorption at 329 nm (SNO) and a strong absorption at 412 nm (SSNO) were observed in the first recorded spectrum (Fig. 1f). SNO was then quickly replaced by SSNO, thus causing the peak at 412 nm to grow in the second minute (blue line). However, SSNO also eventually decomposed, suggesting that both SNO and SSNO were less stable in aqueous media than in non-aqueous solvents.

Although UV-vis spectroscopy is a convenient and relatively fast characterization method, it only provides relatively little structural application cannot be used alone to directly confirm the formation of a particular species. Meanwhile, 15N nuclei usually have very long spin-lattice relaxation times, so that a long acquisition time is required to obtain a 15N NMR spectrum. However, UV-Vis analysis showed that one of the intermediates was very unstable and only lasted briefly in phosphate buffer, which would limit the usefulness of 15N NMR spectroscopy for the detection of this intermediate. With these considerations in mind, we decided to explore the utility of 17O (I = 5/2) NMR spectroscopy, which offers short acquisition times due to the rapid relaxation of this nucleus. As 17O has a very low natural abundance (0.037%), we first prepared 17O-labeled SNAP and GSNO and then allowed these species to react with Na2S to generate 17O-labeled SNO and SSNO. GSNO is water soluble, and 17O-labeled GSNO exhibited a rather broad 17O NMR signal at δ = 1211.6 ppm with a full-width at the half height (FWHH) of 3552 Hz (Fig. 2a). Meanwhile, the 17O-labeled SNAP exhibits a narrower 17O NMR signal at δ = 1296.0 ppm with a FWHH of 1560 kHz (Fig. 2b). These 17O chemical shifts are comparable to that exhibited by the S-nitrosothiols RSNO (R = –CH(COO)(CH2)–COO) that was prepared with 2-mercaptosuccinic acid (MSA) and NaNO2 at δ = 1200 ppm and [Fe(CN)5N(O)SR]3− (R = –CH(CO O)(CH2)–COO) at δ = 1035 ppm (FWHH = 4600 kHz) (Scheme 1).25 These broad 17O NMR signals are due to the relatively large size of the RSNO molecules, which induces a very rapid 17O nuclear quadrupole relaxation. Upon the addition of one molar equivalent of Na2S to the GSNO in phosphate buffer, only one sharp 17O NMR signal was detected at δ = 896.5 ppm (FWHH = 513 Hz) (Fig. 2c). As seen in Fig. 1f, SNO decomposed faster than SSNO, and SNO had decomposed completely during the acquisition of the spectra. Therefore, this 17O NMR signal observed in Fig. 2c can be attributed to SSNO, which has better stability.


image file: d0ra05054g-f2.tif
Fig. 2 17O NMR spectra of (a) 10 mM 17O-labeled GSN O in phosphate buffer pH 7.4, (b) 1.5 mM 17O-labeled SNAP in phosphate buffer pH 7.4, (c) freshly prepared 17O-labeled SSNO that was obtained by mixing 10 mM 17O-labeled GSNO with 1 molar equivalent of Na2S in 1 M phosphate buffer at pH 7.4, (d) 100 mM 17O-labeled SNAP in DMF, (e) freshly prepared 17O-labeled SSNO that was obtained by mixing 100 mM 17O-labeled SNAP with 1 molar equivalent of Na2S in DMF containing 10% D2O, and (f) freshly prepared 17O-labeled SNO that was obtained by adding 6 molar equivalents of TPH to SSNO. TPH: triphenylphosphine.

image file: d0ra05054g-s1.tif
Scheme 1 Structures of S-nitrosothiols. GSNO[thin space (1/6-em)]:[thin space (1/6-em)]S-nitrosoglutathione, SNAP[thin space (1/6-em)]:[thin space (1/6-em)]S-nitroso-N-acetylpenicillamine, and S-nitroso-MSA[thin space (1/6-em)]:[thin space (1/6-em)]S-nitroso-mercaptosuccinic acid.

Since both of these intermediates were less stable in aqueous media, the reaction was also performed in DMF by mixing SNAP with Na2S (at a 1[thin space (1/6-em)]:[thin space (1/6-em)]1 molar ratio). SNAP is relatively hydrophobic and poorly soluble in water, but is soluble in DMF. The signal of SNAP in DMF was detected at δ = 1311.6 ppm (FWHH = 1152 Hz) (Fig. 2d). After mixing SNAP with Na2S (at a 1[thin space (1/6-em)]:[thin space (1/6-em)]1 molar ratio) in DMF only one sharp signal was observed at δ = 987.1 ppm (FWHH = 268 Hz) (Fig. 2e), NMR signals (Fig. 2d and e) were narrower in DMF and the chemical shifts had moved down-field, evidently due to the hydrophobicity of the solvent. After many attempts, we were still unable to obtain the 17O NMR signal of SNO in these reaction mixtures. Therefore, we used triphenylphosphine (TPH) to reduce SSNO to SNO with the product of triphenylphosphine sulfide (TPH-S). By adding excess of TPH to SSNO, a sharp signal was obtained at δ = 1209.6 ppm (FWHH = 176 Hz), which can be attributed to SNO (Fig. 2f). This finding was consistent with our previous 17O NMR data obtained for [Fe(CN)5N(O)S]4− and [Fe(CN)5N(O)SS]4− intermediates that were formed in the Gmelin reaction, in which [Fe(CN)5N(O)S]4− exhibited a 17O NMR signal at δ = 1027 ppm and [Fe(CN)5N (O)SS]4− exhibited a 17O NMR signal at δ = 938 ppm.23

17O-Labelled SSNO and SNO were prepared in the same manner in DMSO, and similar chemical shifts were detected regardless of the solvent effect (Fig. 3a–c). However, upon the addition of D2O, the NMR signals shifted to lower chemical shifts due to the electron shielding effect, which was due to the more hydrophilic surrounding environment (Fig. 3d and e). Filipovic and coworkers have recorded the 15N NMR signal of HSNO/SNO under conditions, but we found that SNO is very sensitive in the presence of water, and decomposed very rapidly.


image file: d0ra05054g-f3.tif
Fig. 3 17O NMR spectra of (a) 100 mM 17O-labeled SNAP in DMSO, (b) freshly prepared 17O-labeled SSNO that was obtained by mixing 100 mM 17O-labeled SNAP with 1 molar equivalent of Na2S in DMSO containing 10% D2O, (c) freshly prepared 17O-labeled SNO that was obtained by adding 2 molar equivalents of TPH to SSNO, (d) a spectrum recorded after water was added to SNO to yield 30% H2O in DMSO, and (e) a spectrum recorded after water was added to SNO to yield 50% H2O in DMSO.

In particular, it was difficult to detect the 17O NMR signal of SNO when the percentage of water exceeded 50% in DMSO. Since We also attempted to obtain the 15N NMR signals of the SSNO and SNO species in DMSO and in water–DMSO mixtures. Firstly, 15N-labeled SNAP was prepared and then allowed to react with Na2S to generate SNO and SSNO. As seen in Fig. 4a, 15N-labeled SNAP exhibited a 15N NMR signal at δ = 836.7 ppm (relative to liquid NH3). This 15N chemical shift is comparable to that exhibited by the RSNO (R = –CH(COO)(CH2)–COO) at δ = 761 ppm in aqueous solution,25 as well as other RSNO (R = a variety of groups) compounds exhibiting 15N NMR signals at δ = 765–830 ppm (Scheme 1).26 Upon the addition of one molar equivalent of Na2S to the SNAP in DMSO, only one sharp 15N NMR signal was detected at δ = 710.7 ppm (Fig. 4b), which we assigned to SSNO. The addition of water to this solution caused the signals to shift to lower resonances, and no signals were visible once the percentage of water exceeded 50% (Fig. 4c and d). SNO was again obtained by adding excess TPH into a freshly prepared SSNO formulation, and the 15N signal was detected at δ = 897.9 ppm. This signal was found to be less stable than that of SSNO in the presence of water, since no signals were visible once the percentage of water exceeded 30% (Fig. 4e and f). The lack of a signal can be attributed to the long spin-lattice relaxation times of 15N, which limit the suitability of this technique for the detection of short-lived intermediates. This finding was consistent with our previous findings obtained via the 15N NMR characterization of intermediates [Fe(CN)5N(O)S]4− and [Fe(CN)5N(O)SS]4− obtained from Gmelin reaction in which [Fe(CN)5N(O)S]4− exhibited a 15N NMR signal at δ = 700 ppm and [Fe (CN)5N(O)SS]4− exhibited 15N NMR signal at δ = 630 ppm.23


image file: d0ra05054g-f4.tif
Fig. 4 15N NMR spectra of (a) 100 mM 15N-labeled SNAP in DMSO, (b) freshly prepared 15N-labeled SSNO that was obtained by mixing 100 mM 15N-labeled SNAP with 1 molar equivalent of Na2S in DMSO containing 10% D2O, (c) water was added to SSNO (the formulation described in (b)) to obtain 30% H2O in DMSO, (d) water was added to SSNO to obtain 50% H2O in DMSO, (e) freshly prepared 15N-labeled SNO that was obtained by adding 3 molar equivalents of TPH to SSNO, and (f) water was added to SNO (the formulation described in (e)) to get 30% H2O in DMSO.

After having established the 17O NMR signature of SNO and SSNO, we turned our attention toward stability (i.e., the half-life) measurements. To this end, we recorded the 17O NMR spectra and plotted the signal intensity as a function of time for SNO and SSNO in different conditions. As reported in previous studies by Feelisch and coworkers, oxygen consumption occurs during the course of this nitrosothiol/sulfide cross-linking reaction.17 SNO and SSNO were therefore prepared under anaerobic conditions to assess their stability, and the preparation methods are described in the caption of Fig. 5. As shown in Fig. 5a, SNO was rather stable with a half-life of 21 h in the absence of oxygen (O2) in DMSO, but once water was added to the preparations, the half-life of SNO had decreased to 87 and 11 min in 30% H2O/DMSO and 50% H2O/DMSO, respectively (Fig. 5b and c). In comparison, the half-life of SNO in DMF was found to be 32 min (Fig. 5d), which was much shorter than the value determined in DMSO. This could be due to the presence of trace amounts of O2 in DMF. These results indicated that SNO was unstable in both water and air. Compared with the half-life of SNO, it was evident that SSNO was more stable in DMSO under the anaerobic conditions with a half-life of 58 h (Fig. 5e), while in the presence of O2 its half-life is reduced to 14 h in DMSO (Fig. 5f). Moreover, the SSNO species that was prepared by mixing GSNO with 2 molar equivalents of Na2S in phosphate buffer at pH 7.4 exhibited a half-life of only 39 min, and that prepared by mixing GSNO[thin space (1/6-em)]:[thin space (1/6-em)]Na2S at a 1[thin space (1/6-em)]:[thin space (1/6-em)]1 molar ratio exhibited an even shorter half-life of 18 min. These collective decomposed quickly under physiological conditions, SSNO is still much more stable than SNO and could be used for bioactivity evaluation in cellular assays.18,19


image file: d0ra05054g-f5.tif
Fig. 5 Decomposition of SNO and SSNO over time as monitored by 17O NMR spectroscopy (with signals at 1216 and 988 ppm, respectively). In (a), SNO was prepared under anaerobic conditions (in a glove box) by adding 1 molar equivalent of Na2S (in H2O) to 100 mM 17O-labeled SNAP in DMSO, and subsequently adding to that solution another DMSO solution containing 3 molar equivalents of TPH. In (b), SNO was obtained as described in (a) except that H2O was added to provide a 30% H2O/DMSO mixture. In (c), SNO was obtained as described in (a) except that H2O was added to provide a 50% H2O/DMSO mixture. In (d), SNO was obtained under anaerobic conditions by adding 1 molar equivalent of Na2S (in H2O) to 100 mM 17O-labeled SNAP in DMF, and subsequently adding 6 molar equivalents of TPH in DMF. In (e), SSNO was obtained under anaerobic conditions by adding 1 molar equivalent of Na2S (in H2O) to 100 mM 17O-labeled SNAP in DMSO. In (f), SSNO was obtained by adding 1 molar equivalent of Na2S (in H2O) to 100 mM 17O-labeled SNAP in DMSO under ambient conditions. In (g), SSNO was obtained under anaerobic conditions by adding 2 molar equivalents of Na2S (in D2O) to 10 mM 17O-labeled SNAP in 1 M D2O phosphate buffer pH 7.4. In (h), SSNO was obtained by adding 1 molar equivalent of Na2S (solid) to 50 mM 17O-labeled GSNO in 1 M D2O phosphate buffer pH 7.4 under ambient conditions.

It has been suggested that the pKa of HSNO exceeds 10.5.15 Unfortunately, we were unable to experimentally measure the pKa value of HSNO/SNO due to its instability in aqueous solution. SSNO was shown to have greater stability under physiological conditions and its chemical shifts that were determined via 17O and 15N NMR were less than the chemical shifts of SNO, suggesting a lower pKa for SSNO. We have recorded the 17O NMR spectra of SSNO at various pH values. Remarkably, the 17O NMR signal does not show any noticeable change over a pH range from 4.69 to 12.10, neither with regard to the signal intensity nor its chemical shift (data are shown in Fig. 6). However, SSNO becomes too unstable below pH 4.69 to be studied by NMR spectroscopy. Nevertheless, the 17O NMR spectra clearly show that SSNO has a pKa value of less than 4.69. Thus, the SSNO anion is the predominate form at physiological pH.


image file: d0ra05054g-f6.tif
Fig. 6 17O NMR spectra of SSNO recorded at various pH values. SSNO was prepared by mixing 10 mM 17O-labeled GSNO with 1 molar equivalent of Na2S in 0.5 M phosphate buffer. Different pH values were obtained by adjusting with acidic resin or NaOH powder. For each spectrum, a total of ca. 21[thin space (1/6-em)]664 transients was recorded with a recycle delay of 20 ms (with a total acquisition time of 27 min).

Since SNO and SSNO are more stable under anaerobic conditions, we attempted to capture the short lived SNO species in a freshly prepared cross-linking reaction mixture by adding one molar equivalent of Na2S to 100 mM 17O-labelled SNAP in DMSO (under anaerobic conditions) and compared the resultant spectrum to that of an otherwise similar sample that was prepared in air. A very weak signal was observed at 1219 ppm in the 17O NMR spectrum of this sample that was prepared under anaerobic conditions (Fig. 7a), which we attributed to SNO. In contrast, this signal at 1219 ppm was not detected in the spectrum of the mixture prepared in air (Fig. 3b).


image file: d0ra05054g-f7.tif
Fig. 7 17O NMR spectra of freshly prepared cross-linking reaction mixtures in DMSO that were obtained via different methods under anaerobic conditions. In (a), the mixture was prepared with 200 μL of 200 mM 17O-SNAP in DMSO with 1 molar equivalent of Na2S (200 μL of DMSO and 40 μL of 1 M Na2S in D2O). In (b), the mixture was prepared by slowly adding 200 μL of 200 mM 17O-labeled SNAP solution to 240 μL of Na2S, via 10 μL additions at 2 min intervals. In (c), the mixture was prepared by slowly adding 240 μL of Na2S solution to 200 μL of 200 mM 17O-labeled SNAP solution, via 10 μL additions at 2 min intervals. In (d) the mixture was prepared with 400 μL 100 mM 17O-labeled SNAP in DMSO with 1 molar equivalent of Na2S (40 μL 1 M Na2S in D2O), with 10% 17O-labeled nicotinamide used as a reference. In (e), the mixture was prepared with 400 μL of 100 mM 17O-labeled SNAP in DMSO with 1 molar equivalent of Na2S2 (40 μL 1 M Na2S2 in D2O), with 10% 17O-labeled nicotinamide used as a reference.

To further support our findings, a comparison between the 17O NMR spectra of samples prepared via different mixing methods was undertaken, as seen in Fig. 7b and c. In Fig. 7b, the mixture was prepared by slowly adding 200 μL of 200 mM SNAP to one molar equivalent of Na2S solution (200 μL DMSO and 40 μL of 1 M Na2S in D2O), via 10 μL additions at 2 min intervals. When this method was employed, the resultant 17O NMR spectrum did not exhibit an SNO signal, suggesting that in the presence of excess HS, SNO was quickly converted to SSNO.

This finding was consistent with our observations from the UV-Vis analysis (Fig. 1f), and in accordance with the previous findings by others via UV-Vis spectroscopy, where it was found that increasing the molar ratio of Na2S caused the intermediates to become converted from SNO to SSNO.17 Conversely, when the method was reversed by slowly adding the Na2S solution to the SNAP solution, neither the SNO nor the SSNO signals were detected, and only the signals of decomposed products such as NO2 and NO3 anions were detected at δ = 683 and 425 ppm, respectively (Fig. 7c). This behavior suggested that mixing SSNO with excess HS would not lead to the formation of SNO, but rather to decomposition to NO2 and NO3 anions. In addition, we compared the amount of intermediate formed by mixing one molar equivalent of Na2S/Na2S2 with SNAP, with 10% 17O-labeled nicotinamide used as an internal reference. As seen in Fig. 7d and e, the signal intensity was much higher if Na2S2 (which provide HSS in the reaction) was added instead of Na2S. This further demonstrated that the 17O signal at 986.9 ppm represents SSNO.

After the SNO and SSNO were successfully characterized via UV-vis and NMR spectroscopy, we further investigated their bioactivity via cultured HUVEC. The research undertaken by Koeppenol et al. indicated that the formation of SSNO was thermodynamically unfavorable due to the low concentrations of reactants in vivo.27 In the current study, we prepared the SNO and SSNO via a chemical reaction in vitro, and evaluate their bioactivity accordingly. Cells were grown in the recommended conditions as described in the Experimental Section and respectively treated with SNAP, Na2S, TPH, and TPH-S alone, as well as with the SSNO-enriched mixtures of SNAP and Na2S that were mixed at different molar ratios (SNAP[thin space (1/6-em)]:[thin space (1/6-em)]Na2S = 1[thin space (1/6-em)]:[thin space (1/6-em)]2 and 1[thin space (1/6-em)]:[thin space (1/6-em)]5),17 or with SNO-enriched mixtures of SNAP, Na2S, and TPH that were mixed at different molar ratio (SNAP[thin space (1/6-em)]:[thin space (1/6-em)]Na2S[thin space (1/6-em)]:[thin space (1/6-em)]TPH = 1[thin space (1/6-em)]:[thin space (1/6-em)]2:2, 1[thin space (1/6-em)]:[thin space (1/6-em)]2:3, 1[thin space (1/6-em)]:[thin space (1/6-em)]5:3 and 1[thin space (1/6-em)]:[thin space (1/6-em)]5:5). The intracellular concentration of nitric oxide (NO) in live cells after these treatments was measured with a NO indicator, 3-amino,4-aminomethyl-2′,7′-difluorescein, diacetate (DAF-FM DA). HUVEC were initially incubated with this NO probe which can freely cross cell membranes. Once inside the cells, the probe can react with NO and generate a strong fluorescence signal, which can then be analyzed via flow cytometry.28 After the treatments were applied to cells, higher levels of intracellular NO would be revealed by increased fluorescence intensities, as shown in Fig. 8, NC groups contained only the NO probe, without receiving any of the treatments listed above, and were employed to measure the endogenous NO that initially resided in the cells. Respective treatments with Na2S, TPH, and TPH-S alone showed little effect on the NO concentration. Conversely, cells treated with SSNO-enriched mixtures of SNAP and Na2S (at both 1[thin space (1/6-em)]:[thin space (1/6-em)]2 and 1[thin space (1/6-em)]:[thin space (1/6-em)]5 molar ratios) showed a substantial elevation of the NO concentrations, suggesting that SSNO increased the intracellular NO levels in HUVEC. In contrast, no significant increase of NO levels was observed in cells that were treated with SNO-enriched mixtures, suggesting that SNO does not exhibit NO releasing behavior in cell cultures. This could be due to the instability of SNO under physiological conditions and this finding was consistent with our 17O NMR measurements of the half-life values of SNO and SSNO (Fig. 5).


image file: d0ra05054g-f8.tif
Fig. 8 (a) The cellular NO levels as determined in HUVEC after had received various treatments, including SNAP, Na2S, TPH, SSNO-enriched mixture, and SNO-enriched mixture. (b) The cGMP levels as observed in HUVEC after these cells had received various treatments. *p < 0.05, **p < 0.01, compared to NC. HUVEC: human umbilical vein endothelial cells, cGMP: cyclic guanosine monophosphate.

Although the DAF-FM DA NO detection assays suggested an elevation of intracellular NO concentration in SSNO treated cell cultures, the method also encountered limitations in the presence of oxidants/antioxidants, which interfere with the results by potentially modifying the steady-state concentration of the NO˙ radical.29 Therefore, the release of NO by SSNO in cells was further demonstrated by determining the changes in cGMP levels, as cGMP is a downstream marker of NO signaling and is involved in the regulation of vascular tone.30 As shown in Fig. 8e, the cGMP concentration did not change significantly with 2 mM Na2S, but treatment with 5 mM Na2S decreased the cGMP concentration in cells. Meanwhile, treatment with SSNO-enriched mixtures increased the cGMP levels in the cells regardless of the negative effect of excess Na2S, indicating that SSNO strongly increased the cGMP level through a NO-dependent signaling pathway. This effect was also proven in previous studies, where elevated cGMP levels were detected after RFL-6 cells had been treated with SSNO-enriched mixtures.17,22 These results showed the potential bioactivity of SSNO in NO-cGMP-dependent vasodilation.

Ondrias and coworkers have previously suggested that the products obtained via cross-linking reactions between RSNO and sulfide moieties could be a potent vasorelaxants.18 To this end, we evaluated the effect of SSNO in promoting angiogenesis, as angiogenesis and vasodilation accompany one another in vivo and NO can mediate angiogenesis.31,32 As seen in Fig. 9a, SSNO-enriched mixtures of SNAP and Na2S promoted HUVEC proliferation regardless of the negative effects that are presented by excess Na2S in the cell culture. Scratched wound healing assays were utilized to assess the migration capabilities of cells that had been treated with SSNO-enriched mixtures. The scratched area was measured via Image J software. In comparison to the initial wound area, cells that had been treated with SNAP or with SSNO-enriched mixtures of SNAP and Na2S (1[thin space (1/6-em)]:[thin space (1/6-em)]2 and 1[thin space (1/6-em)]:[thin space (1/6-em)]5 molar ratio) showed a significant reduction in the wound area after 24 h incubation, while treatment with Na2S (2 and 5 mM) alone resulted in only a slight reduction in the wound area (Fig. 9c). The slower wound area closure rate observed in cells treated with 2 and 5 mM Na2S could be due to the cytotoxicity of Na2S during prolonged incubation, as it was also observed that treatment with Na2S inhibited cell proliferation (Fig. 9a). The migration rate was calculated by the equation: migration rate = (W0Wt)/W0, where W0 is the initial wound area, and Wt is the wound area after 24 h of incubation. As seen in Fig. 9b, the migration rate of HUVEC was significantly reduced in the Na2S treatment groups, but increased in the groups treated with SNAP and the SSNO-enriched mixtures in comparison to the NC group, thus demonstrating the cell migration promoting properties of SSNO.


image file: d0ra05054g-f9.tif
Fig. 9 The proliferation and migration of HUVEC was assessed with the treatment of SNAP, Na2S, and SSNO-enriched mixtures. (a) Cell proliferation rate determined after the above treatments, (b) migration rate comparison among the treatment groups, and (c) photographs taken under 100× magnification. *p < 0.05, **p < 0.01, compared to the NC group.

Moreover, tube forming assays were also employed to evaluate the angiogenesis promoting effect of SSNO-enriched mixtures. HUVEC are the commonly used cell line to study the angiogenesis in vitro.33 The photographs were taken under 400× magnification (Fig. 10a) and the numbers of meshes were counted in order to evaluate the ability of cells to form tubes in 3D culturing conditions (Fig. 10b). Both SNAP and the SSNO-enriched mixture were found to significantly increase the number of meshes in cells, which was not achieved with Na2S.


image file: d0ra05054g-f10.tif
Fig. 10 The proliferation of HUVEC was assessed with the treatment of SNAP, Na2S, and a SSNO-enriched mixture using the endothelial cell tube formation assay. (a) Photographs taken under 400× magnification, and (b) comparison of the numbers of meshes among the treatment groups, *p < 0.05, **p < 0.01, compared to NC.

Conclusions

In summary, we have characterized the intermediates from the cross-linking reaction between H2S and NO via UV-Vis spectroscopy and provided for the first time the 15N and 17O NMR data for previously elusive SNO and SSNO intermediates. Moreover, the 17O NMR data indicated that SSNO has a pKa value of less than 4.7. Thus, the SSNO anion is the predominate species at physiological pH. We have also discovered that the SNO was rapidly converted to SSNO and was less stable than SSNO. In vitro SNO and SSNO bioactivity analysis suggested that SSNO acts as a NO-releasing intermediate in cells, increases cellular cGMP levels, and facilitates angiogenesis in vitro through a NO-cGMP dependent signaling pathway. We hope that our research will contribute toward a better understanding of the H2S and NO cross-linking reaction as well as the chemical and biological characteristics of the reaction intermediates. In addition, we hope to encourage others to consider 17O NMR spectroscopy as a promising new method to probe short-lived reaction intermediates.

Experimental

Materials and methods

Chemicals were obtained from Sigma-Aldrich unless otherwise stated. These included sodium hydroxide (NaOH), sulfuric acid (H2SO4), sodium sulfide (Na2S·9H2O), sodium nitrite (NaNO2), 15N-labeled sodium nitrite (Na15NO2, 98% 15N), 17O-labeled water (H217O, 41.1% 17O, purchased from CortecNet), L-glutathione (GSH), N-acetyl-D-penicillamine (NAP), monosodium phosphate (NaH2PO4), disodium phosphate (Na2HPO4), triphenylphosphine, dimethyl sulfoxide (DMSO), dimethylformamide (DMF), deuterium oxide (D2O, CIL), acetone-d6 (CIL), Matrigel (Corning), 3-(4,5-di-methylthiazol-2-yl)-2,5-diphenyl tetrazolium bromide (MTT), and ion-exchange resin (Amberlite IR-120, strongly acidic form). HUVEC was obtained from the Cell Bank of the Committee on Type Culture Collection of Chinese Academy of Sciences (Shanghai, China). Complete cell growth medium RPMI 1640 (Thermo Fisher Scientific, Shanghai, China) containing 10% fetal bovine serum (FBS, Thermo Fisher Scientific) and 1% penicillin/streptomycin (Thermo Fisher Scientific) were used to grow cells. The levels of NO and cGMP within the cells was determined with the use of DAF-FM DA (NO detection kit, Beyotime Biotechnology) and cGMP ELISA detection kit (GenScript).

Synthesis of isotope-labelled compounds

15N-Labelled SNAP was prepared according to the literature method.34 NAP (191.2 mg) was mixed with 1.1 molar equivalents of Na15NO2 in 0.58 mL of 0.55 M HCl. The reaction mixture was kept on ice for 40 min prior to the addition of 10 μL of concentrated H2SO4. The product was collected by filtration, washed with ice-cold water (5 × 2 mL), and subsequently dried under vacuum to give a green powder (120 mg, 63% yield). The 15N enrichment of the product was 98%. 15N NMR (40.6 MHz, acetone-d6): δ = 835 ppm (ref. to liquid NH3). 15N NMR (50.6 MHz, DMF): δ = 836.7 ppm.

17O–NaNO2 was (0.8 g, 11.6 mmol) was dissolved in 40% 17O-labelled H2O (M wt. 19–20, 1 g, 50 mmol) in acidic condition. The mixture was stirred slowly for 20 min until a solution was obtained. This solution was heated overnight at 75 °C. 17O–NaNO2 signal was recorded at 661.15 ppm. The sample was then lyophilized overnight.

17O-SNAP was prepared in 0.5 mL of 0.55 M HCl (in H217O, 41.1% 17O) by mixing NAP (231.1 mg) with 1.1 molar equivalents of NaN17O2. The work-up procedures were similar to that used in the preparation of the 15N-labeled SNAP. The 17O enrichment of the product was 30%. 17O NMR (54.3 MHz, acetone-d6): δ = 1314 ppm (ref. to water). 17O NMR (67.7 MHz, DMF): δ = 1312 ppm. 17O NMR (67.7 MHz, PB buffer pH 7.41): δ = 1296 ppm. 17O NMR (67.7 MHz, DMSO-d6): δ = 1307 ppm. FWHH = 2053 Hz (line broadening = 20).

15N-GSNO was prepared according to the literature method.35 GSH (150 mg) was mixed with 1.1 molar equivalents of Na15NO2 in 1 mL of 0.48 M HCl. The reaction mixture was kept on ice for 40 min prior to the addition of 3 mL of ice-cold acetone, and subsequently stirred for 10 min. The product was collected by centrifugation at 3000 rpm for 2 min. The supernatant was discarded. The residue was resuspended in acetone. This process was repeated 3 times. The residue was dried under a flow of Ar gas and left under vacuum to give a pink powder (69 mg, 46% yield). 15N NMR (40.6 MHz, D2O): δ = 768 ppm (ref. to liquid NH3).

17O-GSNO was prepared in 0.58 mL of 0.55 M HCl (in H217O, 41.1% 17O) by mixing GSH (150 mg) with 1.1 molar equivalents of NaN17O2. The work-up procedures were the same as that used to prepare the 15N-labeled GSNO. The 17O enrichment of the product was 30%. 17O NMR (54.3 MHz, D2O): δ = 1211 ppm (ref. to water).

Preparation of SSNO and SNO-enriched mixtures. SSNO-enriched mixtures were prepared by mixing SNAP with Na2S at various molar ratios in different solvents under aerobic or anaerobic conditions as indicated in the Fig legends. SNO-enriched mixtures were prepared via the addition of various molar equivalents of TPH into the SSNO-enriched mixtures under different conditions as indicated in the Fig legends. For the cell cultures, a high concentration of Na2S stock solution (0.5 M) was freshly prepared with 0.5 M phosphate buffer at pH 7.4, and PBS was then used to dilute this solution to lower concentrations. These solutions were then mixed with SNAP at various ratios as indicated in the figures.
Measurement of cellular NO level. Cellular NO levels were determined with the DAF-FM DA NO detection kit. Briefly, HUVEC were seeded in 6-well plates at a density of 3 × 105 cells per well and grown to reach 100% confluence. The medium was gently aspirated with a pipette before 1 mL of the probe (1[thin space (1/6-em)]:[thin space (1/6-em)]1000 dilution of the original solution) was added to each well. Plates were placed into a plate shaker and incubated for 20 min. The probe was then aspirated and the wells were washed twice with PBS (pH 7.4) to remove the remaining probe. HUVEC were then treated with SNAP, Na2S, TPH, TPH-S, SNO and SSNO-enriched mixtures in serum free medium accordingly for 20 min. The medium was then aspirated from the wells and washed once with PBS gently. Cells were trypsinized for 3 min and then neutralized by adding complete growth medium. Cell pellets were collected by centrifugation, washed with PBS, and re-suspended in 500 μL of PBS in flow cytometry tubes, covered with aluminum foil, and finally analyzed via flow cytometry.
Measurement of cellular cGMP levels. Cellular cGMP levels were determined via a cGMP ELISA Detection kit according to the manufacturer's instructions. Briefly, HUVEC were seeded into 6-well plates at a density of 3 × 105 cells per well and grown to reach 100% confluence. The growth medium was then replaced with SNAP, Na2S, and SSNO-enriched mixtures in serum-free medium. After 20 min of incubation, the medium was aspirated and 100 μL of RIPA lysis buffer was added and each sample was incubated for 20 min. The cells were scraped from the wells using a cell scraper, centrifuged at 2000 rpm for 10 min, and the supernatants were then collected. To prepare an ELISA plate for cGMP detection, 100 μL of Anti-cGMP pAb was added to each well of a blank ELISA plate. Plates were sealed and incubated at 25 °C for 45 min to achieve an optimal linking. After incubation, wells were washed 4 times with 260 μL of washing solution. Subsequently, the plate was inverted on absorbent paper to remove residual liquid from the wells. Assay buffer A (100 μL) was added to the non-specific binding (NSB) wells; 100 μL of standard cGMP solutions at concentrations of 0, 0.3, 0.8, 2.5, 7.4, 22.2, 66.7 pmol mL−1 were added to the calibration wells, and cell supernatants were added to the remaining wells. After this step, 50 μL of assay buffer B was added to the NSB well, and 50 μL of cGMP-HRP was added to the remaining wells. After 1 h of incubation at 4 °C, plates were washed 4 times with 260 μL of washing solution and the residual liquid in the wells was removed. TMB (100 μL) was then added to all of the wells and further incubated at 25 °C in the dark for 20 min. Finally, 50 μL of stop solution was added to all of the wells to stop the reaction. The absorbance of each well was measured at OD 450 nm via an enzyme-linked immunosorbent detector.
Measurement of cell proliferation. The cell proliferation effect was evaluated with MTT assays. HUVEC were seeded in 96-well plates at a density of 8000 cells per well and incubated for 24 h to reach 70% confluence. The complete growth medium was replaced with serum-free medium containing SNAP, Na2S, and SSNO in the corresponding wells, and further incubated for 24 h. MTT (10 μL, 5 mg mL−1) was then added to each well and the reaction was allowed to proceed for another 4 h. Finally, 150 μL of DMSO was added to each well and the plate was agitated on a plate shaker for 3 min. The cell viability was determined using a microplate reader at wavelengths of 490 and 560 nm.
Measurement cell migration. Cell migration was studied using scratch assays. HUVEC were seeded in 6-well plates at a density of 3 × 105 cells per well and grew to reach 100% confluence. The scratches were made through the monolayer in the middle of each plate using 200 μL pipette tips. Wells were washed 3 times with PBS to remove floating cells. Cells were then treated with SNAP, Na2S, and SSNO-enriched mixtures in serum free medium. After 24 h of incubation, images were obtained with an Olympus BX53 microscope.
Measurement angiogenesis. In vitro angiogenesis was examined using Matrigel-based (a liquid laminin/collagen gel) endothelial cell tube formation assay. Briefly, 20 μL per well of Matrigel were added into 96-well plates. While waiting for Matrigel to solidify, cells were trypsinized and resuspended using fresh growth medium. After the Matrigel had solidified, HUVEC were seeded in 96-well plates containing Matrigel at a density of 3 × 104 cells per well. Subsequently, SNAP, Na2S, and SSNO-enriched mixture were added to the corresponding wells and incubated for 4 h. Cells were observed using a microscope (Olympus BX53).

Conflicts of interest

There are no conflicts to declare.

Acknowledgements

Y. G. thanks the National Natural Science Foundation of China [grant number 31700713]; the Department of Science and Technology of Jilin Province [grant number 20200801066GH] for financial support.

Notes and references

  1. B. Olas, Clin. Chim. Acta, 2015, 439, 212–218 CrossRef CAS.
  2. S. Cacanyiova, A. Berenyiova and F. Kristek, Physiological Research, 2016, 65, S273–S289 CAS.
  3. D. J. Polhemus and D. J. Lefer, Circ. Res., 2014, 114, 730–737 CrossRef CAS.
  4. H. Kimura, Nitric Oxide, 2014, 41, 4–10 CrossRef CAS.
  5. M. Y. Ali, C. Y. Ping, Y. Y. Mok, L. Ling, M. Whiteman, M. Bhatia and P. K. Moore, Br. J. Pharmacol., 2006, 149, 625–634 CrossRef CAS.
  6. G. Yang, L. Wu, B. Jiang, W. Yang, J. Qi, K. Cao, Q. Meng, A. K. Mustafa, W. Mu, S. Zhang, S. H. Snyder and R. Wang, Science, 2008, 322, 587–590 CrossRef CAS.
  7. S. Koike, Y. Ogasawara, N. Shibuya, H. Kimura and K. Ishii, FEBS Lett., 2013, 587, 3548–3555 CrossRef CAS.
  8. Y. Kimura, Y. Mikami, K. Osumi, M. Tsugane, J. Oka and H. Kimura, FASEB J., 2013, 27, 2451–2457 CrossRef CAS.
  9. M. Eberhardt, M. Dux, B. Namer, J. Miljkovic, N. Cordasic, C. Will, T. I. Kichko, J. de la Roche, M. Fischer, S. A. Suarez, D. Bikiel, K. Dorsch, A. Leffler, A. Babes, A. Lampert, J. K. Lennerz, J. Jacobi, M. A. Marti, F. Doctorovich, E. D. Hogestatt, P. M. Zygmunt, I. Ivanovic-Burmazovic, K. Messlinger, P. Reeh and M. R. Filipovic, Nat. Commun., 2014, 5, 4381 CrossRef CAS.
  10. K. Ondrias, A. Stasko, S. Cacanyiova, Z. Sulova, O. Krizanova, F. Kristek, L. Malekova, V. Knezl and A. Breier, Pflügers Arch., 2008, 457, 271–279 CrossRef CAS.
  11. M. M. Cortese-Krott, B. O. Fernandez, M. Kelm, A. R. Butler and M. Feelisch, Nitric Oxide, 2015, 46, 14–24 CrossRef CAS.
  12. M. M. Cortese-Krott, A. R. Butler, J. D. Woollins and M. Feelisch, Dalton Trans., 2016, 45, 5908–5919 RSC.
  13. R. Wedmann, I. Ivanovic-Burmazovic and M. R. Filipovic, Interface Focus, 2017, 7, 20160139 CrossRef.
  14. I. Ivanovic-Burmazovic and M. R. Filipovic, Inorg. Chem., 2019, 58, 4039–4051 CrossRef CAS.
  15. M. R. Filipovic, J. Miljkovic, T. Nauser, M. Royzen, K. Klos, T. Shubina, W. H. Koppenol, S. J. Lippard and I. Ivanovic-Burmazovic, J. Am. Chem. Soc., 2012, 134, 12016–12027 CrossRef CAS.
  16. C. L. Bianco and J. M. Fukuto, Proc. Natl. Acad. Sci. U. S. A., 2015, 112, 10573–10574 CrossRef CAS.
  17. M. M. Cortese-Krott, B. O. Fernandez, J. L. Santos, E. Mergia, M. Grman, P. Nagy, M. Kelm, A. Butler and M. Feelisch, Redox Biol., 2014, 2, 234–244 CrossRef CAS.
  18. A. Berenyiova, M. Grman, A. Mijuskovic, A. Stasko, A. Misak, P. Nagy, E. Ondriasova, S. Cacanyiova, V. Brezova, M. Feelisch and K. Ondrias, Nitric Oxide, 2015, 46, 123–130 CrossRef CAS.
  19. M. M. Cortese-Krott, D. Pullmann and M. Feelisch, Pharmacol. Res., 2016, 113, 490–499 CrossRef CAS.
  20. R. Wedmann, A. Zahl, T. E. Shubina, M. Durr, F. W. Heinemann, B. E. Bugenhagen, P. Burger, I. Ivanovic-Burmazovic and M. R. Filipovic, Inorg. Chem., 2015, 54, 9367–9380 CrossRef CAS.
  21. J. P. Marcolongo, U. N. Morzan, A. Zeida, D. A. Scherlis and J. A. Olabe, Phys. Chem. Chem. Phys., 2016, 18, 30047–30052 RSC.
  22. M. M. Cortese-Krott, G. G. Kuhnle, A. Dyson, B. O. Fernandez, M. Grman, J. F. DuMond, M. P. Barrow, G. McLeod, H. Nakagawa, K. Ondrias, P. Nagy, S. B. King, J. E. Saavedra, L. K. Keefer, M. Singer, M. Kelm, A. R. Butler and M. Feelisch, Proc. Natl. Acad. Sci. U. S. A., 2015, 112, E4651–E4660 CrossRef CAS.
  23. Y. Gao, A. Toubaei, X. Kong and G. Wu, Chemistry, 2015, 21, 17172–17177 CrossRef CAS.
  24. D. Tsikas and A. Bohmer, Nitric Oxide, 2017, 65, 22–36 CrossRef CAS.
  25. Y. Gao, B. Mossing and G. Wu, Dalton Trans., 2015, 44, 20338–20343 RSC.
  26. W. Kun, H. Yongchun, Z. Wei, M. B. Ksebati, X. Ming, J.-P. Cheng and P. G. Wang, Bioorg. Med. Chem. Lett., 1999, 9, 2897–2902 CrossRef.
  27. W. H. Koppenol and P. L. Bounds, Arch. Biochem. Biophys., 2017, 617, 3–8 CrossRef CAS.
  28. C. J. Ku, W. Karunarathne, S. Kenyon, P. Root and D. Spence, Anal. Chem., 2007, 79, 2421–2426 CrossRef CAS.
  29. M. M. Cortese-Krott, A. Rodriguez-Mateos, G. G. Kuhnle, G. Brown, M. Feelisch and M. Kelm, Free Radicals Biol. Med., 2012, 53, 2146–2158 CrossRef CAS.
  30. F. Z. Monica, K. Bian and F. Murad, Adv. Pharmacol., 2016, 77, 1–27 CAS.
  31. M. Ziche and L. Morbidelli, Journal of Neuro-Oncology, 2000, 50, 139–148 CrossRef CAS.
  32. A. Papapetropoulos, G. Garcia-Cardena, J. A. Madri and W. C. Sessa, J. Clin. Invest., 1997, 100, 3131–3139 CrossRef CAS.
  33. J. M. Quesada-Gomez, R. Santiago-Mora, C. Navarro-Valverde, G. Dorado and A. Casado-Diaz, J. Steroid Biochem. Mol. Biol., 2015, 148, 214–218 CrossRef CAS.
  34. D. Tsikas, D. O. Stichtenoth, R. H. Boger, S. M. Bode-Boger and J. C. Frolich, J. Labelled Compd. Radiopharm., 1994, 34, 1055–1062 CrossRef CAS.
  35. T. W. Hart, M. B. Vine and N. R. Walden, Tetrahedron Lett., 1985, 26, 3879–3882 CrossRef CAS.

This journal is © The Royal Society of Chemistry 2020