Next Article in Journal
Hydrogen Sulfide (H2S)/Polysulfides (H2Sn) Signalling and TRPA1 Channels Modification on Sulfur Metabolism
Previous Article in Journal
Progress to Clarify How NOTCH3 Mutations Lead to CADASIL, a Hereditary Cerebral Small Vessel Disease
Previous Article in Special Issue
Mitochondrial Reactive Oxygen Species, Insulin Resistance, and Nrf2-Mediated Oxidative Stress Response—Toward an Actionable Strategy for Anti-Aging
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Mitochondria Play Essential Roles in Intracellular Protection against Oxidative Stress—Which Molecules among the ROS Generated in the Mitochondria Can Escape the Mitochondria and Contribute to Signal Activation in Cytosol?

by
Daisuke Masuda
1,2,†,
Ikuo Nakanishi
3,*,†,
Kei Ohkubo
4,
Hiromu Ito
3,5,
Ken-ichiro Matsumoto
6,
Hiroshi Ichikawa
7,
Moragot Chatatikun
8,9,
Wiyada Kwanhian Klangbud
8,9,
Manas Kotepui
8,
Motoki Imai
10,11,
Fumitaka Kawakami
10,12,13,
Makoto Kubo
10,14,15,
Hirofumi Matsui
16,
Jitbanjong Tangpong
8,17,
Takafumi Ichikawa
10,11,
Toshihiko Ozawa
18,
Hsiu-Chuan Yen
19,20,
Daret K. St Clair
21,
Hiroko P. Indo
5,* and
Hideyuki J. Majima
1,5,8,17,*
1
Department of Space Environmental Medicine, Kagoshima University Graduate School of Medical and Dental Sciences, Kagoshima 890-8544, Kagoshima, Japan
2
Utilization & Engineering Department, Japan Manned Space Systems Corporation, 2-1-6 Tsukuba, Tsukuba 305-0047, Ibaraki, Japan
3
Quantum RedOx Chemistry Team, Institute for Quantum Life Science (iQLS), Quantum Life and Medical Science Directorate (QLMS), National Institutes for Quantum Science and Technology (QST), 4-9-1 Anagawa, Inage-ku, Chiba 263-8555, Japan
4
Institute for Advanced Co-Creation Studies, Open and Transdisciplinary Research Initiatives, Osaka University, Suita 565-0871, Japan
5
Department of Maxillofacial Radiology, Field of Oncology, Graduate School of Medical and Dental Sciences, Kagoshima University, Kagoshima 890-8544, Kagoshima, Japan
6
Quantitative RedOx Sensing Group, Department of Radiation Regulatory Science Research, Institute for Radiological Science (NIRS), Quantum Life and Medical Science Directorate (QLMS), National Institutes for Quantum Science and Technology (QST), 4-9-1 Anagawa, Inage-ku, Chiba 263-8555, Japan
7
Department of Medical Life Systems, Graduate School of Life and Medical Sciences, Doshisha University, Kyoto 610-0394, Kyoto, Japan
8
School of Allied Health Sciences, Walailak University, Thasala, Nakhon Si Thammarat 80161, Thailand
9
Center of Excellence Research for Melioidosis and Microorganisms, Walailak University, Thasala, Nakhon Si Thammarat 80161, Thailand
10
Regenerative Medicine and Cell Design Research Facility, School of Allied Health Sciences, Kitasato University, 1-15-1 Kitasato, Sagamihara 252-0373, Kanagawa, Japan
11
Department of Molecular Diagnostics, School of Allied Health Sciences, Kitasato University, 1-15-1 Kitasato, Sagamihara 252-0373, Kanagawa, Japan
12
Department of Regulation Biochemistry, Kitasato University Graduate School of Medical Sciences, 1-15-1 Kitasato, Sagamihara 252-0373, Kanagawa, Japan
13
Department of Health Administration, School of Allied Health Sciences, Kitasato University, 1-15-1 Kitasato, Sagamihara 252-0373, Kanagawa, Japan
14
Division of Microbiology, Kitasato University School of Allied Health Sciences, 1-15-1 Kitasato, Minami-ku, Sagamihara 252-0373, Kanagawa, Japan
15
Department of Environmental Microbiology, Graduate School of Medical Sciences, Kitasato University, 1-15-1 Kitasato, Minami-ku, Sagamihara 252-0373, Kanagawa, Japan
16
Division of Gastroenterology, Graduate School of Comprehensive Human Science, University of Tsukuba, Tsukuba 305-8575, Ibaraki, Japan
17
Research Excellence Center for Innovation and Health Products (RECIHP), School of Allied Health Sciences, Walailak University, Thasala, Nakhon Si Thammarat 80160, Thailand
18
Nihon Pharmaceutical University, 10281 Komuro, Ina-machi, Kitaadachi-gun, Saitama 362-0806, Saitama, Japan
19
Department of Medical Biotechnology and Laboratory Science, College of Medicine, Chang Gung University, Taoyuan 33302, Taiwan
20
Department of Nephrology, Chang Gung Memorial Hospital at Linkou, Taoyuan 33305, Taiwan
21
Department of Toxicology and Cancer Biology, University of Kentucky College of Medicine, Lexington, KY 40536, USA
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Biomolecules 2024, 14(1), 128; https://doi.org/10.3390/biom14010128
Submission received: 8 November 2023 / Revised: 28 December 2023 / Accepted: 29 December 2023 / Published: 19 January 2024

Abstract

:
Questions about which reactive oxygen species (ROS) or reactive nitrogen species (RNS) can escape from the mitochondria and activate signals must be addressed. In this study, two parameters, the calculated dipole moment (debye, D) and permeability coefficient (Pm) (cm s−1), are listed for hydrogen peroxide (H2O2), hydroxyl radical (•OH), superoxide (O2•−), hydroperoxyl radical (HO2•), nitric oxide (•NO), nitrogen dioxide (•NO2), peroxynitrite (ONOO), and peroxynitrous acid (ONOOH) in comparison to those for water (H2O). O2•− is generated from the mitochondrial electron transport chain (ETC), and several other ROS and RNS can be generated subsequently. The candidates which pass through the mitochondrial membrane include ROS with a small number of dipoles, i.e., H2O2, HO2•, ONOOH, •OH, and •NO. The results show that the dipole moment of •NO2 is 0.35 D, indicating permeability; however, •NO2 can be eliminated quickly. The dipole moments of •OH (1.67 D) and ONOOH (1.77 D) indicate that they might be permeable. This study also suggests that the mitochondria play a central role in protecting against further oxidative stress in cells. The amounts, the long half-life, the diffusion distance, the Pm, the one-electron reduction potential, the pKa, and the rate constants for the reaction with ascorbate and glutathione are listed for various ROS/RNS, •OH, singlet oxygen (1O2), H2O2, O2•−, HO2•, •NO, •NO2, ONOO, and ONOOH, and compared with those for H2O and oxygen (O2). Molecules with negative electrical charges cannot directly diffuse through the phospholipid bilayer of the mitochondrial membranes. Short-lived molecules, such as •OH, would be difficult to contribute to intracellular signaling. Finally, HO2• and ONOOH were selected as candidates for the ROS/RNS that pass through the mitochondrial membrane.

1. Introduction

Reactive oxygen species (ROS) and reactive nitrogen species (RNS) consist of both radical and nonradical molecules and are reactive species that have different degrees of oxidizing potential in biological systems [1]. Many chronic diseases, such as cancer, alcoholic liver disease, Crohn’s disease, rheumatoid arthritis, diabetes, muscular dystrophy, cystic fibrosis, septic shock, premature babies, atherosclerosis, infertility, cataracts, aging, hepatitis, ARDS, ischemia, neuronal degeneration, etc., are recognized as oxidative-stress-related diseases (OSDs) [2]. A major source of ROS in cells is the mitochondria [3]. The electron transport chain (ETC) consists of Complexes I, II, III, and IV. Oxidative phosphorylation is the process of the coupling between the ETC and ATP production in Complex V. Mitochondrial DNA (mtDNA) encodes 13 proteins inside the mitochondrial matrix, and those proteins are parts of Complexes I, III, IV, and V. [4]. Overall, 2~3% of electrons leak from the ETC and oxygen captures them, resulting in the production of superoxide anions (O2•−). It is well known that mitochondria are the major site of ATP production, but they also produce O2•−, which mainly leaks from Complexes I and III [2]. Impairment of the ETC caused by chemicals or mtDNA damage can cause an increase in the generation of O2•− and subsequent ROS [3]. These impairments are closely related to the cause of OSDs [4,5]. Hydroperoxyl radical (HO2•) is the protonated form of O2•−, but whether its amount could be affected by the pH gradient across the mitochondrial inner membrane is uncertain [6]. There is evidence of nitic oxide (•NO) formation in the mitochondria, although whether mitochondrial nitric oxide synthase (NOS) exists is still controversial [7]. Singlet oxygen (1O2) can be generated endogenously through different mechanisms [8], but its formation in the mitochondria has only been addressed in one study [9].
In mammalian cells, there are three superoxide dismutase (SOD) isoenzymes: copper–zinc SOD (CuZnSOD), or SOD1 [10]; manganese SOD (MnSOD), or SOD2 [11]; and extracellular SOD (ECSOD), or SOD3 [12]. SOD catalyzes the dismutation of two superoxide radicals into hydrogen peroxide and oxygen. MnSOD is an enzyme localized in the mitochondrial matrix. Okado-Matsumoto and Fridovich showed that CuZnSOD is localized in the intermembrane space of the mitochondria [13]. It has been recognized that increases in the generation of ROS from the mitochondria can cause lipid oxidation and apoptosis. MnSOD could protect against these processes [14].
How do antioxidant systems, which are intracellular defense systems, work? MnSOD generates one hydrogen peroxide (H2O2) from two superoxide radicals (O2•−). MnSOD may also reduce the formation of hydroxyl radicals (•OH) from superoxide (O2•−) and hydrogen peroxide (H2O2) through the Haber–Weiss reaction under the catalysis of iron ions [15,16,17]. However, H2O2 from MnSOD could be quickly detoxified by mitochondrial glutathione peroxidase (mtGPx) by reducing it to water [14,18]. This reaction could be accompanied by glutathione, of which the level for most cells is ~5 mM, an excess amount for the reaction [14,18]. Furthermore, GPx4 knockout (KO) is known to cause acute renal failure and death [19,20], suggesting that GPx4 plays an essential role as an antioxidant in mitochondria. Due to the emergence of the role of nitric oxide (•NO) in OSDs, reactive nitrogen cascades are sometimes included in reactive oxygen cascades. O2•− and •NO can be easily bound and produce peroxynitrite (ONOO) with k = 5 × 109 M−1 s−1; however, in the opposite reaction, k = 0.023 s−1 [21]. ONOOH produces •NO2 and •OH with k = 0.35 s−1, indicating that the decomposition of ONOO and ONOOH is not straightforward [21]. Kissner et al. (2003) suggested that, regarding peroxynitrite formation under physiological conditions, when 10 nM •NO and 10 µM SOD, ONOO formation/O2•− dismutation is 1/125, while with 2 µM •NO and 2 µM SOD, ONOO formation/ O2•− dismutation is 8/1 [22], suggesting that ONOO formation is dependent on intracellular •NO concentration.
Mitochondrial ROS (mtROS) might be related to an increase in signal transduction and may control anti-oxidative-stress-related molecular defense mechanisms. Redox states could thus represent essential pathways to maintain homeostasis. The importance of this subject, the mitochondrial ROS come out from mitochondria and initiate the signal transduction inside cells, has been hypothesized by many researchers [23,24,25,26,27,28,29,30,31,32,33]. The role of mitochondrial ROS in initiating signal transductions in the cell cytosol has been the subject of discussion [34]. Indo et al. showed that manganese superoxide dismutase (MnSOD) transfection decreases the expression levels of GATA 1, 3, 4, and 5, which are nuclear factor kappa-light-chain-enhancer of activated B cells (NF-κB) regulating genes [34]. The results showed that MnSOD transfected cells revealed a decrease in expression compared to those in the control. We previously demonstrated that mtROS causes intracellular signaling, and we published a paper entitled “Evidence of Nrf2/Keap1 Signaling Regulation by Mitochondrial-Generated Oxygen Species in RGK1 cells” in a Special Issue of Biomolecules entitled “The Physiological and Pathological New Function of Mitochondrial ROS and Intraorganellar Cross-Talks” in 2023 (https://www.mdpi.com/journal/biomolecules/special_issues/0XTJ2MAYET, accessed on 7 November 2023) [35]. They transfected MnSOD gene-contained vectors in a gastric mucosal tumorized cell line, RGK1 cells. They examined the expression levels of NF-E2-related factor 2 (Nrf2), Kelch-like ECH-associated protein1 (Keap1), heme oxygenase-1 (HO-1) and 2, MnSOD, glutamate-cysteine ligase (GCL), glutathione S-transferase (GST), and NAD(P)H Quinone oxidoreductase 1 (NQO1), which are all Nrf2-Keap1 regulating gens. The results of immunocytochemistry staining showed a decrease in those expressions in the MnSOD transfected RGK1 cells compared to those in the control. The transfected MnSOD gene should decrease the mitochondrial ROS levels, so after MnSOD transfection, all decreased expression was shown, suggesting mtROS levels control the levels of Nrf2-Keap1 regulating genes. However, the question of which ROS go out from mitochondria and contribute to intracellular signaling remains unclear.
The plasma membrane consists of both lipids and proteins. The fundamental structure of the membrane is the phospholipid bilayer, which forms a stable barrier between two aqueous compartments [36]. Most biologically important solutes require protein carriers to cross cell membranes, via a process of either passive or active transport. Active transport requires the cell to expend energy to move the materials, while passive transport can be performed without using cellular energy [37]. Certain substances easily pass through the membrane through passive diffusion, such as O2 and CO2, along with small relatively hydrophobic molecules, fatty acids, and alcohols [37]. Mitochondria possess double membranes, and the inner membrane contains cardiolipin. Cardiolipin is not the main lipid that forms a phospholipid bilayer but fulfills other functions (e.g., stabilization of protein complexes), because it contains four fatty acid residues, and is a non-bilayer forming phospholipid [38,39]. It is known that cardiolipin is oxidized in mitochondria by X-irradiation [40]. If the ROS are related to cell defense signal transduction, ROS must pass through the membranes and exist in the cytosol to activate signal transduction. In this study, in the mitochondria, we study which ROS can pass through the mitochondrial membrane.
In this paper, we try to clarify which ROS are responsible for signal activation in cytosol through calculations and examination of the literature: •OH, singlet oxygen (1O2), HO2•, •NO, •NO2, ONOO, ONOOH. The dipole moments of ROS and RNS are calculated using density functional theory (DFT) calculations. Possible candidates of ROS which pass through the mitochondrial membrane and enter the cytosol to activate the signal transduction pathway are estimated using the calculated dipole moment and experimental permeability coefficient. In addition, the lifetime of each molecule is listed, and ROS that escape from the mitochondria and act as initiators to activate signal transduction in the cytosol are taken into consideration.

2. Materials and Methods

2.1. Theoretical Calculations of Dipole Moments for ROS and RNS

The dipole moments [41] were calculated according to the dipole information (Table 1). The DFT calculations were performed using Gaussian 09 (Revision A.02, Gaussian, Inc., Wallingford, CT, USA) [42]. The calculations were performed on a 32-processor QuantumCubeTM (Parallel Quantum Solutions, Fayetteville, AR, USA) at the B3LYP/6-311++G(3df,3pd) level of theory [43,44,45] with a keyword “polar” to output the molecular polarity (electric dipole moment in D (debye)) [46]. Graphical outputs of the computational results were generated using the GaussView software program (ver. 3.09) developed by Semichem, Inc., Shawnee, KS, USA [47]. The dipole moments were calculated for various ROS and RNS; we calculated the dipole moments of major ROS and RNS that exist in the mitochondria (Table 1): hydroxyl radical (•OH), superoxide (O2•−), hydroperoxyl radical (HO2•), nitric oxide (•NO), nitrogen dioxide (•NO2), peroxynitrite (ONOO), and peroxynitrous acid (ONOOH). We also listed the number of molecules of water (H2O) and hydrogen peroxide (H2O2).

2.2. Predictive Performance of Mitochondria-Originating Reactive Oxygen Species

The predictive performance of mitochondria-originating reactive oxygen species included the following parameters: the intracellular amount (amount/cell); the half-life; the diffusion distance (µm); permeability coefficients (Pm) (in cm s−1); the one-electron reduction potential (Eo) (in V vs. NHE, NHE: normal hydrogen electrode) at pH 7.4; pKa; and the rate constants for the reaction with ascorbate (AscH) (k (AscH)/M−1 s−1) and glutathione (GSH) (k (GSH)/M−1 s−1) for various ROS and RNS. We focused on ROS generated from the mitochondrial electron transport chain (mtETC). The ROS studied included •OH, singlet oxygen (1O2), O2•−, HO2•, •NO, •NO2, ONOO, ONOOH, alkoxyl radicals (RO•), and peroxyl radicals (ROO•). The H2O, oxygen (O2), and H2O2 were also listed. Those radicals were initiated from O2•−, starting from electron leakage from the ETC and then binding with O2. Then, the O2•− changed form to become other ROS, such as •OH, singlet oxygen (1O2), HO2•, •NO, •NO2, ONOO, and ONOOH, in the mitochondria [2]. •OH and •NO2 are constructed by the binding of O2•− and •NO. This information was collected from the literature listed in the Table 2 references. To exit the mitochondrial membrane into the cytosol, the ROS should be present in an appropriate amount and have a long half-life, long diffusion distance, large Pm and Eo, and relatively small rate constants for the k (AscH)/M−1 s−1) and k (GSH)/M−1 s−1).

3. Results

The results for the calculated dipole moment (in D) and experimental permeability coefficient (in cm s−1) are listed in Table 1.
Table 2 shows the predictive performance of the mitochondria-originating ROS. The intracellular amount (amount/cell); the half-life; the diffusion distance (µm); permeability coefficients (Pm; cm s−1); Eo, the one-electron reduction potential (V vs. NHE) at pH 7.4; pKa; and the rate constants for the reaction with ascorbate (AscH) (k (AscH)/M−1 s−1) and glutathione (GSH) (k (GSH)/M−1 s−1) were examined. For considerations of reactions of ROS in the mitochondria, we used AscH and GSH. Finally, we detected the ONOOH and HO2• for the responsible ROS, which crossed the mitochondrial membrane and initiated the intracellular signaling in cytosol (Figure 1).

4. Discussion

Majima et al. were the first to report that reactive oxygen species (ROS) generated from the mitochondria promote apoptosis [106], while Itoh et al. described the function of the Nrf2-Keap1 intercellular signal for the first time [107,108]. A recent study described that ROS generated from the mitochondria initiates cellular transduction in the cytosol [34,35]. The further roles of ROS and the subsequent intracellular signals, proteins, and molecule transport change need to be clarified. The establishment of cellular signaling and metabolism change based on mitochondrial ROS augmentation is in demand. Thus, studies on the physiological and pathological functions of mitochondrial ROS will be necessary.
This paper aims to consider the roles of mitochondrial ROS in the activation of intracellular signals. The dipole potential (represented by Ψd) is shown as the potential difference that arises due to the nonrandom orientation of dipolar residues of the lipids and associated water molecules within the membrane [109,110]. ROS with a positive or negative charge cannot escape mitochondria by passive diffusion through phospholipid bilayer due to their large number of dipoles. The results of the dipole moments (Table 1) show that H2O2 is permeable (the dipole moment is 0.00 D). The dipole moment of •NO2 was 0.35 D, indicating permeability. Although the dipole moment of O2•− is 0.00 D, the negative charge in O2•− precludes its penetration into the membrane. ONOO is non-permeable. H2O (with a dipole moment of 1.89 D), •OH (with a dipole moment of 1.67 D), ONOOH (with a dipole moment of 1.77 D), and HO2• (with a dipole moment of 2.23 D) might be permeable. The candidates that can escape from the mitochondria include ROS with small dipole moments, i.e., H2O2, •NO, •NO2, HO2•, ONOOH, •OH, and H2O. It is well known that •NO2 reacts with urate, ascorbate, and GSH at 107 M−1 s−1 [96]. Therefore, the reaction of •NO2 with specific targets in the cytoplasm, where GSH is present at µM~mM levels [111,112], likely occurs with very low frequency [113]. The candidates that can escape from the mitochondria thus include ROS with small dipole moments, i.e., H2O2, HO2•, ONOOH, •OH, and •NO.
The reactivity of ROS/RNS should be essential. However, if the molecules disappear in a short period, there is less chance of the reaction occurring. A greater amount, a long half-life, a greater diffusion distance, a greater Pm, a greater I, a greater one-electron reduction potential, a smaller pKa, and greater rate constants for the reaction with ascorbate and GSH would be preferable for the studied ROS/RNS. Molecules with electrical charges cannot pass the phospholipid bilayers of mitochondrial membranes [36]. Short-lived molecules, such as •OH, are difficult to contribute to intracellular signaling due to the characteristics of the short-lived molecule (Table 2). For signal activation inside the cytosol, again, H2O2, HO2•, ONOOH, •OH, and •NO can be selected as candidates (Table 2).
It is also essential to consider the conditions that ROS/RNS must overcome to pass through the mitochondrial membrane to become signaling molecules in the cytosol. The plasma membrane consists of both lipids and proteins. The fundamental structure of the membrane is the phospholipid bilayer, which forms a stable barrier between two aqueous compartments. [36]. Most biologically important solutes require protein carriers to cross cell membranes via a process of either passive or active transport. Active transport requires the cell to expend energy to move the materials, while passive transport can be achieved without using cellular energy [37]. Certain substances easily pass through the membrane via passive diffusion, such as O2 and CO2, along with small relatively hydrophobic molecules, fatty acids, and alcohols [37]. In this study, in the mitochondria, we study which ROS can pass through the mitochondrial membrane.
The ROS produced in the mitochondrial matrix can pass through the two membranes in the mitochondria and enter into the cytosol in order to initiate intracellular signals. Lynch and Fridovich (1978) addressed the question of whether superoxide permeates membranes [114]. The pH of the intermembrane space is lower than that in the matrix due to proton pumping into the intermembrane space; in the intermembrane space (IMS), the concentration of protons is about ten times higher than in the matrix [115]. The pH values obtained were 6.88 ± 0.09 in the IMS, 7.78 ± 0.17 in the matrix, and 7.59 ± 0.01 in the cytosol using a human endothelial cell line, ECV304. [103]. HO2• and O2•− are of considerable importance in oxidation processes, and the pKa of HO2•/O2•− is 4.8 [62,90]. Therefore, at the physiological pH, HO2• hardly exists. In addition to covalent, there is also ionic bonding. There are almost 10 times more protons in the IMS compared to in the matrix. Thus, it may be possible for H+ to bind anion molecules, leading to protonation [116]. ROS produced in the mitochondria, HOON- and O2•−, can be easily protonated in the IMS through ionic bonding. Whereas O2•− generated in the mitochondrial matrix may be easily and completely detoxified by mitochondrial SOD, any O2•− generated on the outside of the inner membrane will have a longer lifetime and, due to the more acidic environment there than in the matrix, it is likely that O2•− will be protonated to HO2• and react with a phospholipid in the membrane [117]. Which radicals can penetrate through the mitochondrial membrane? Gus’kova et al. (1984) determined the permeability of the liposomal membrane for O2•− and HO2•, being P’O2•− = (7.6 + 0.3) × 10−8 cm s−1 and P’HO2• = 4.9 × 10−4 cm s−1, respectively [51]. Cordeiro (2014) described simulations that showed that molecular oxygen (O2) accumulated at the interior membrane. Superoxide (O2•−) radicals and hydrogen peroxide (H2O2) remained in the aqueous phase and could not enter the membrane. Both hydroxyl (•OH) and hydroperoxyl (HO2•) radicals were able to penetrate deep into the lipid headgroup region in the membrane [118]. ROS are produced in the mitochondria, and to establish which ROS can pass through the membrane, we needed to establish the interactions between ROS and the lipid membrane. Cordeiro evaluated HO2, O2•−, •OH, and H2O2 in terms of the residence times in the phospholipid headgroup region, reported in units of ns [118]. The results show that HO2 and O2•− have residence times of 17.3 and 12.4 ns, respectively, while •OH and H2O2 have residence times of 3.8 and 1.5 ns, respectively. A longer residence time suggests a higher affinity for the ROS and phospholipids, and a shorter residence time suggests a lower affinity for the ROS and phospholipids. O2•− in the mitochondrial intermembrane space can penetrate the outer membrane mitochondrial membranes through voltage-dependent anion channels (VDACs) [114,119]. However, how much O2•− can penetrate through VDACs is unknown. It may be possible for HO2• to pass through the membrane without difficulty.

5. Conclusions

As a result, HO2• and ONOOH were found to be the top candidates to initiate intracellular signaling among the mitochondrial ROS from Table 1 and Table 2. Figure 1 shows the possible ROS that can initiate signal transduction in cells, which are HO2• and ONOOH. Further experiments to prove that HO2• and ONOOH go out of mitochondria and initiate signals inside cells will be necessary.

Author Contributions

Conceptualization and formal analysis, D.M. and H.I. (Hiromu Ito); research planning, discussion of the results and writing, and original draft preparation, I.N., K.O., K.-i.M., H.I. (Hiroshi Ichikawa), M.C., W.K.K., M.K. (Manas Kotepui), M.I., F.K., M.K. (Makoto Kubo), H.M., T.I. and T.O.; all aspects of this study, including writing, review, and editing, H.-C.Y., D.K.S.C., H.P.I., J.T. and H.J.M. All authors have read and agreed to the published version of the manuscript.

Funding

We thank the JSPS (the Japan Society for the Promotion of Science) Core-to-Core Program for supporting our research.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All data are shown in this paper.

Acknowledgments

The authors thank Tomoaki Sato, Kazuo Tomita, Shosei Kishida, and Shigeaki Suenaga of Kagoshima University and William St. Clair of the University of Kentucky for their help and encouragement in completing this study.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Halliwell, B.; Whiteman, M. Measuring reactive species and oxidative damage in vivo and in cell culture: How should you do it and what do the results mean? Br. J. Pharmacol. 2004, 142, 231–255. [Google Scholar] [CrossRef]
  2. Majima, H.J.; Indo, H.P.; Suenaga, S.; Kaneko, T.; Matsui, H.; Yen, H.-C.; Ozawa, T. Mitochondria as source of free radicals. In Free Radical Biology in Digestive Diseases; Naito, Y., Suematsu, M., Yoshikawa, T., Eds.; Karger Publisher: Basel, Switzerland, 2011; Volume 29, pp. 12–22. [Google Scholar] [CrossRef]
  3. Indo, H.P.; Davidson, M.; Yen, H.-C.; Suenaga, S.; Tomita, K.; Nishii, T.; Higuchi, M.; Koga, Y.; Ozawa, T.; Majima, H.J. Evidence of ROS generation by mitochondria in cells with impaired electron transport chain and mitochondrial DNA damage. Mitochondrion 2007, 7, 106–118. [Google Scholar] [CrossRef]
  4. DiMauro, S.; Schon, E.A. Mitochondrial respiratory-chain diseases. N. Engl. J. Med. 2003, 348, 2656–2668. [Google Scholar] [CrossRef]
  5. DiMauro, S.; Schon, E.A.; Carelli, V.; Hirano, M. The clinical maze of mitochondrial neurology. Nat. Rev. Neurol. 2013, 9, 429–444. [Google Scholar] [CrossRef]
  6. Salvador, A.; Sousa, J.; Pinto, R.E. Hydroperoxyl, superoxide and pH gradients in the mitochondrial matrix: A theoretical assessment. Free Radic. Biol. Med. 2001, 31, 1208–1215. [Google Scholar] [CrossRef]
  7. Tengan, C.H.; Moraes, C.T. NO control of mitochondrial function in normal and transformed cells. Biochim. Biophys. Acta Bioenerg. 2017, 1858, 573–581. [Google Scholar] [CrossRef]
  8. Onyango, A.N. Endogenous Generation of singlet oxygen and ozone in human and animal tissues: Mechanisms, Biological significance, and influence of dietary components. Oxid. Med. Cell Longev. 2016, 2016, 2398573. [Google Scholar] [CrossRef]
  9. Kerver, E.D.; Vogels, I.M.; Bosch, K.S.; Vreeling-Sindelárová, H.; Van den Munckhof, R.J.; Frederiks, W.M. In situ detection of spontaneous superoxide anion and singlet oxygen production by mitochondria in rat liver and small intestine. Histochem. J. 1997, 29, 229–237. [Google Scholar] [CrossRef]
  10. McCord, J.M.; Fridovich, I. Superoxide dismutase. An enzymic function for erythrocuprein (hemocuprein). J. Biol. Chem. 1969, 244, 6049–6055. [Google Scholar] [CrossRef]
  11. Weisiger, R.A.; Fridovich, I. Mitochondrial superoxide dismutase. J. Biol. Chem. 1973, 248, 4793–4796. [Google Scholar] [CrossRef]
  12. Fukai, T.; Siegfried, M.R.; Ushio-Fukai, M.; Cheng, Y.; Kojda, G.; Harrison, D.G. Regulation of the vascular extracellular superoxide dismutase by nitric oxide and exercise training. J. Clin. Investig. 2000, 105, 1631–1639. [Google Scholar] [CrossRef]
  13. Okado-Matsumoto, A.; Fridovich, I. Subcellular distribution of superoxide dismutases (SOD) in rat liver Cu,Zn-SOD in mitochondria. J. Biol. Chem. 2001, 276, 38388–38393. [Google Scholar] [CrossRef]
  14. Indo, H.P.; Yen, H.-C.; Nakanishi, I.; Matsumoto, K.I.; Tamura, M.; Nagano, Y.; Matsui, H.; Gusev, O.; Cornette, R.; Okuda, T.; et al. A mitochondrial superoxide theory for oxidative stress diseases and aging. J. Clin. Biochem. Nutr. 2015, 56, 1–7. [Google Scholar] [CrossRef]
  15. Liochev, S.I.; Fridovich, I. Superoxide and iron: Partners in crime. IUBMB Life 1999, 48, 157–161. [Google Scholar] [CrossRef]
  16. Liochev, S.I.; Fridovich, I. The Haber-Weiss cycle—70 years later: An alternative view. Redox Rep. 2002, 7, 55–57; author reply 59–60. [Google Scholar] [CrossRef]
  17. Liochev, S.I.; Fridovich, I. The effects of superoxide dismutase on H2O2 formation. Free Radic. Biol. Med. 2007, 42, 1465–1469. [Google Scholar] [CrossRef]
  18. Davies, K.J. Oxidative stress, antioxidant defenses, and damage removal, repair, and replacement systems. IUBMB Life 2000, 50, 279–289. [Google Scholar] [CrossRef]
  19. Friedmann Angeli, J.P.; Schneider, M.; Proneth, B.; Tyurina, Y.Y.; Tyurin, V.A.; Hammond, V.J.; Herbach, N.; Aichler, M.; Walch, A.; Eggenhofer, E.; et al. Inactivation of the ferroptosis regulator Gpx4 triggers acute renal failure in mice. Nat. Cell Biol. 2014, 16, 1180–1191. [Google Scholar] [CrossRef]
  20. Azuma, K.; Koumura, T.; Iwamoto, R.; Matsuoka, M.; Terauchi, R.; Yasuda, S.; Shiraya, T.; Watanabe, S.; Aihara, M.; Imai, H.; et al. Mitochondrial glutathione peroxidase 4 is indispensable for photoreceptor development and survival in mice. J. Biol. Chem. 2022, 298, 101824. [Google Scholar] [CrossRef]
  21. Goldstein, S.; Rabai, J. Mechanism of nitrite formation by nitrate photolysis in aqueous solutions: The role of peroxynitrite, nitrogen dioxide, and hydroxyl radical. J. Am. Chem. Soc. 2007, 129, 10597–10601. [Google Scholar] [CrossRef]
  22. Kissner, R.; Nauser, T.; Kurz, C.; Koppenol, W.H. Peroxynitrous acid--where is the hydroxyl radical? IUBMB Life 2003, 55, 567–572. [Google Scholar] [CrossRef]
  23. Finkel, T. Signal transduction by mitochondrial oxidants. J. Biol. Chem. 2012, 287, 4434–4440. [Google Scholar] [CrossRef]
  24. Zhang, D.X.; Gutterman, D.D. Mitochondrial reactive oxygen species-mediated in endothelial cells. Am. J. Physiol. Heart Circ. Physiol. 2007, 292, H2023–H2031. [Google Scholar] [CrossRef]
  25. Cosentino-Goes, D.; Rocco-Machado, N.; Meyer-Fernandes, J.R. Cell signaling through protein kinase C oxidation and activation. Int. J. Mol. Sci. 2012, 13, 10697–10721. [Google Scholar] [CrossRef]
  26. Holmström, K.M.; Finkel, T. Cellular mechanisms and physiological consequences of redox-dependent signalling. Nat. Rev. Mol. Cell Biol. 2014, 15, 411–421. [Google Scholar] [CrossRef]
  27. Chandel, N.S. Evolution of Mitochondria as Signaling Organelles. Cell Metab. 2015, 22, 204–206. [Google Scholar] [CrossRef]
  28. Shadel, G.S.; Horvath, T.L. Mitochondrial ROS signaling in organismal homeostasis. Cell 2015, 163, 560–569. [Google Scholar] [CrossRef]
  29. Itoh, K.; Ye, P.; Matsumiya, T.; Tanji, K.; Ozaki, T. Emerging functional cross-talk between the Keap1-Nrf2 system and mitochondria. J. Clin. Biochem. Nutr. 2015, 56, 91–97. [Google Scholar] [CrossRef]
  30. Brand, M.D. Mitochondrial generation of superoxide and hydrogen peroxide as the source of mitochondrial redox signaling. Free Radic. Biol. Med. 2016, 100, 14–31. [Google Scholar] [CrossRef]
  31. Bouchez, C.; Devin, A. Mitochondrial biogenesis and mitochondrial reactive oxygen species (ROS): A complex relationship regulated by the cAMP/PKA signaling pathway. Cells 2019, 8, 287. [Google Scholar] [CrossRef]
  32. Brillo, V.; Chieregato, L.; Leanza, L.; Muccioli, S.; Costa, R. Mitochondrial dynamics, ROS, and cell signaling: A blended overview. Life 2021, 11, 332. [Google Scholar] [CrossRef]
  33. Kasai, S.; Kokubu, D.; Mizukami, H.; Itoh, K. Mitochondrial reactive oxygen species, insulin resistance, and Nrf2-mediated oxidative stress response-toward an actionable strategy for anti-aging. Biomolecules 2023, 13, 1544. [Google Scholar] [CrossRef]
  34. Indo, H.P.; Hawkins, C.L.; Nakanishi, I.; Matsumoto, K.-I.; Matsui, H.; Suenaga, S.; Davies, M.J.; St Clair, D.K.; Ozawa, T.; Majima, H.J. Role of Mitochondrial reactive oxygen species in the activation of cellular signals, molecules, and function. Handb. Exp. Pharmacol. 2017, 240, 439–456. [Google Scholar] [CrossRef]
  35. Indo, H.P.; Masuda, D.; Sriburee, S.; Ito, H.; Nakanishi, I.; Matsumoto, K.-I.; Mankhetkorn, S.; Chatatikun, M.; Surinkaew, S.; Udomwech, L.; et al. Evidence of Nrf2/Keap1 Signaling regulation by mitochondria-generated reactive oxygen species in RGK1 cells. Biomolecules 2023, 13, 445. [Google Scholar] [CrossRef]
  36. Cooper, G. Structure of the plasma membrane. In The Cell: A Molecular Approach, 2nd ed.; ASM Press: Washington, DC, USA, 2000; ISBN 0-87893-106-6. Available online: https://www.ncbi.nlm.nih.gov/books/NBK9898/ (accessed on 14 April 2023).
  37. Stein, W.D.; Lieb, W.R. Transport and Diffusion across Cell Membranes; Academic Press: New York, NY, USA, 1986. [Google Scholar] [CrossRef]
  38. Serricchio, M.; Vissa, A.; Kim, P.K.; Yip, C.M.; McQuibban, G.A. Cardiolipin synthesizing enzymes form a complex that interacts with cardiolipin-dependent membrane organizing proteins. Biochim. Biophys. Acta Mol. Cell Biol. Lipids 2018, 1863, 447–457. [Google Scholar] [CrossRef]
  39. Gasanoff, E.S.; Yaguzhinsky, L.S.; Garab, G. Cardiolipin, non-bilayer structures and mitochondrial bioenergetics: Relevance to cardiovascular disease. Cells 2021, 10, 1721. [Google Scholar] [CrossRef]
  40. Indo, H.P.; Inanami, O.; Koumura, T.; Suenaga, S.; Yen, H.C.; Kakinuma, S.; Matsumoto, K.; Nakanishi, I.; St Clair, W.; St Clair, D.K.; et al. Roles of mitochondria-generated reactive oxygen species on X-ray-induced apoptosis in a human hepatocellular carcinoma cell line, HLE. Free Radic. Res. 2012, 46, 1029–1043. [Google Scholar] [CrossRef]
  41. Clough, S.A.; Beers, Y.; Klein, G.P.; Rothman, L.S. Dipole moment of water from Stark measurements of H2O, HDO, and D2O. Chem. Phys. 1973, 59, 2254–2259. [Google Scholar] [CrossRef]
  42. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Nakatsuji, H.; Li, X.; et al. Gaussian 09, Revision A.02; Gaussian, Inc.: Wallingford, CT, USA, 2009; Available online: https://gaussian.com/g09citation/ (accessed on 30 April 2021).
  43. Becke, A.D. Density-functional thermochemistry. III. The role of exact exchange. J. Chem. Phys. 1993, 98, 5648–5652. [Google Scholar] [CrossRef]
  44. Lee, C.; Yang, W.; Parr, R.G. Development of the Colle-Salavati correlation-energy formula into a functional of the electron density. Phys. Rev. B Condens. Matter 1988, 37, 785–789. [Google Scholar] [CrossRef]
  45. Hay, J.J.; Wadt, R.W. Ab initio effective core potentials for molecular calculations. Potentials for the transition metal atoms Sc to Hg. J. Chem. Phys. 1985, 82, 270–283. [Google Scholar] [CrossRef]
  46. Olsen, J.; Jørgensen, P.P. Linear and nonlinear response functions for an exact state and for an MCSCF state. J. Chem. Phys. 1985, 82, 3235–3264. [Google Scholar] [CrossRef]
  47. Dennington, R.; Keith, T.; Millam, J. Gauss View, Version 6.1.1; Semichem, Inc.: Wallingford, CT, USA, 2003. [Google Scholar]
  48. Finkelstein, A.; Cass, A. Effect of cholesterol on the water permeability of thin lipid membranes. Nature 1967, 216, 717–718. [Google Scholar] [CrossRef]
  49. Chance, B.; Sies, H.; Boveris, A. Hydroperoxide metabolism in mammalian organs. Physiol. Rev. 1979, 59, 527–605. [Google Scholar] [CrossRef]
  50. Takahashi, M.; Asada, K. Superoxide anion permeability of phospholipid membranes and chloroplast thylakoids. Arch. Biochem. Biophys. 1983, 226, 558–566. [Google Scholar] [CrossRef]
  51. Gus’kova, R.A.; Ivanov, I.I.; Kol’tover, V.K.; Akhobadze, V.V.; Rubin, A.B. Permeability of bilayer lipid membranes for superoxide (O2) radicals. Biochim. Biophys. Acta 1984, 778, 579–585. [Google Scholar] [CrossRef]
  52. Subczynski, W.K.; Lomnicka, M.; Hyde, J.S. Permeability of nitric oxide through lipid bilayer membranes. Free Radical. Res. 1996, 24, 343–349. [Google Scholar] [CrossRef]
  53. Mendiara, S.N.; Perissinotti, L.J. Dissociation equilibrium of dinitrogen tetroxide in organic solvents: An electron paramagnetic resonance measurement. Appl. Magn. Reson. 2003, 25, 323–346. [Google Scholar] [CrossRef]
  54. Denicola, A.; Souza, J.M.; Radi, R. Diffusion of peroxynitrite across erythrocyte membranes. Proc. Natl. Acad. Sci. USA 1998, 95, 3566–3571. [Google Scholar] [CrossRef]
  55. Marla, S.S.; Lee, J.; Groves, J.T. Peroxynitrite rapidly permeates phospholipid membranes. Proc. Natl. Acad. Sci. USA 1997, 94, 14243–14248. [Google Scholar] [CrossRef]
  56. Khairutdinov, R.F.; Coddington, J.W.; Hurst, J.K. Permeation of phospholipid membranes by peroxynitrite. Biochemistry 2000, 39, 14238–14249. [Google Scholar] [CrossRef] [PubMed]
  57. Trujillo, M.; Piacenza, L.; Radi, R. Reactivity of mitochondrial peroxiredoxins with biological hydroperoxides. Redox Biochem. Chem. 2023, 5–6, 100017. [Google Scholar] [CrossRef]
  58. Möller, M.N.; Lancaster, J.R., Jr.; Denicola, A. Chapter 2 The interaction of reactive oxygen and nitrogen species with membranes. In Current Topics in Membranes; Matalon, S., Ed.; Academic Press Inc.: New York, NY, USA, 2008; Volume 61, pp. 23–42. [Google Scholar] [CrossRef]
  59. Möller, M.N.; Cuevasanta, E.; Orrico, F.; Lopez, A.C.; Thomson, L.; Denicola, A. Diffusion and Transport of Reactive Species Across Cell Membranes. Adv. Exp. Med. Biol. 2019, 1127, 3–19. [Google Scholar] [CrossRef]
  60. Armstrong, D.A.; Huie, R.E.; Lymar, S.; Koppenol, W.H.; Merényi, G.; Neta, P.; Stanbury, D.M.; Steenken, S.; Wardman, P. Standard electrode potentials involving radicals in aqueous solution: Inorganic radicals. BioInorg. React. Mech. 2013, 9, 59–61. [Google Scholar] [CrossRef]
  61. Buettner, G.R.; Jurkiewicz, B.A. Catalytic metals, ascorbate and free radicals: Combinations to avoid. Radiat. Res. 1996, 145, 532–541. [Google Scholar] [CrossRef] [PubMed]
  62. Warren, J.J.; Tronic, T.A.; Mayer, J.M. Thermochemistry of proton-coupled electron transfer reagents and its implications. Chem. Rev. 2010, 110, 6961–7001. [Google Scholar] [CrossRef] [PubMed]
  63. Bartberger, M.D.; Liu, W.; Ford, E.; Miranda, K.M.; Switzer, C.; Fukuto, J.M.; Farmer, P.J.; Wink, D.A.; Houk, K.N. The reduction potential of nitric oxide (NO) and its importance to NO biochemistry. Proc. Natl. Acad. Sci. USA 2002, 99, 10958–10963. [Google Scholar] [CrossRef]
  64. Koppenol, W.H.; Stanbury, D.M.; Bounds, P.L. Electrode potentials of partially reduced oxygen species, from dioxygen to water. Free Radic. Biol. Med. 2010, 49, 317–322. [Google Scholar] [CrossRef]
  65. Finosh, G.Y.; Joyabalan, M. Reactive oxygen species—Control and management using amphiphilic biosynthetic hydrogels for cardiac applications. Adv. Biosci. Biotech. 2013, 4, 1134–1146. [Google Scholar] [CrossRef]
  66. Karogodina, T.Y.; Sergeeva, S.V.; Stass, D.V. Stability and reactivity of free radicals: A physicochemical perspective with biological implications. Hemoglobin 2011, 35, 262–275. [Google Scholar] [CrossRef]
  67. Phaniendra, A.; Jestadi, D.B.; Periyasamy, L. Free radicals: Properties, sources, targets, and their implication in various diseases. Indian J. Clin. Biochem. 2015, 30, 11–26. [Google Scholar] [CrossRef]
  68. Pryor, W.A. Oxy-radicals and related species: Their formation, lifetimes, and reactions. Annu. Rev. Physiol. 1986, 48, 657–667. [Google Scholar] [CrossRef] [PubMed]
  69. Sies, H. Strategies of antioxidant defense. Eur. J. Biochem. 1993, 215, 213–219. [Google Scholar] [CrossRef] [PubMed]
  70. Beckman, J.S. Peroxynitrite versus hydroxyl radical: The role of nitric oxide in superoxide-dependent cerebral injury. Ann. N. Y. Acad. Sci. 1994, 738, 69–75. [Google Scholar] [CrossRef]
  71. Bonini, M.G.; Augusto, O. Carbon dioxide stimulates the production of thiyl, sulfinyl, and disulfide radical anion from thiol oxidation by peroxynitrite. J. Biol. Chem. 2001, 276, 9749–9754. [Google Scholar] [CrossRef]
  72. Augusto, O.; Miyamoto, S. Oxygen radicals and related species. In Principles of Free Radical Biomedicine; Pantopoulos, K., Schipper, H.M., Eds.; Nova Science Publishers: Hauppauge, NY, USA, 2011; Volume 1, pp. 1–23. [Google Scholar]
  73. Winterbourn, C.C. Reconciling the chemistry and biology of reactive oxygen species. Nat. Chem. Biol. 2008, 4, 278–286. [Google Scholar] [CrossRef]
  74. Sueishi, M.; Hori, M.; Ishikawa, M.; Matsu-Ura, K.; Kamogawa, E.; Honda, Y.; Kita, M.; Ohara, K. Scavenging rate constants of hydrophilic antioxidants against multiple reactive oxygen species. J. Clin. Biochem. Nutr. 2014, 54, 67–74. [Google Scholar] [CrossRef] [PubMed]
  75. Keszler, A.; Zhang, Y.; Hogg, N. Reaction between nitric oxide, glutathione, and oxygen in the presence and absence of protein: How are S-nitrosothiols formed? Free Radic. Biol. Med. 2010, 48, 55–64. [Google Scholar] [CrossRef]
  76. Adams, G.E.; Boag, J.W.; Currant, J.; Michael, B.D. Absolute rate constants for the reaction of the hydroxyl radical with organic compounds. In Pulse Radiolysis; Ebert, M., Keene, J.P., Swallow, A.J., Baxendale, J.H., Eds.; Academic Press: New York, NY, USA, 1965; pp. 131–143. [Google Scholar]
  77. Liphard, M.; Bothe, E.; Schulte-Frohlinde, D. The influence of glutathione on single-strand breakage in single-stranded DNA irradiated in aqueous solution in the absence and presence of oxygen. Int. J. Radiat. Biol. 1990, 58, 589–602. [Google Scholar] [CrossRef]
  78. Gupta, V.; Carroll, K.S. Sulfenic acid chemistry, detection and cellular lifetime. Biochim. Biophys. Acta 2014, 1840, 847–875. [Google Scholar] [CrossRef]
  79. Ross, F.; Alberta, B.; Ross, A.B. Selected Specific Rates of Reactions of Transients from Water in Aqueous Solution. III. Hydroxyl Radical and Perhydroxyl Radical and Their Radical Ions; Natl. Stand. Ref. Data Syst. No. 59; National Bureau of Standards: Washington, DC, USA, 1977.
  80. Eriksen, T.E.; Fransson, G. Formation of reducing radicals on radiolysis of glutathione and some related compounds in aqueous solution. J. Chem. Soc. Perkin Trans. 1988, 2, 1117–1122. [Google Scholar] [CrossRef]
  81. Carballal, S.; Bartesaghi, S.; Radi, R. Kinetic and mechanistic considerations to assess the biological fate of peroxynitrite. Biochim. Biophys. Acta 2014, 1840, 768–780. [Google Scholar] [CrossRef] [PubMed]
  82. Engstrom, P.C.; Easterling, L.; Baker, R.R.; Matalon, S. Mechanisms of extracellular hydrogen peroxide clearance by alveolar type II pneumocytes. J. Appl. Physiol. 1985, 69, 2078–2084. [Google Scholar] [CrossRef] [PubMed]
  83. Kramarenko, G.G.; Hummel, S.G.; Martin, S.M.; Buettner, G.R. Ascorbate reacts with singlet oxygen to produce hydrogen peroxide. Photochem. Photobiol. 2006, 82, 1634–1637. [Google Scholar] [CrossRef] [PubMed]
  84. Shimizu, R.; Yagi, M.; Kikuchi, A. Suppression of riboflavin-sensitized singlet oxygen generation by l-ascorbic acid, 3-O-ethyl-l-ascorbic acid and Trolox. J. Photochem. Photobiol. B 2019, 191, 116–122. [Google Scholar] [CrossRef]
  85. Sies, H.; Jones, D.P. Reactive oxygen species (ROS) as pleiotropic physiological signalling agents. Nat. Rev. Mol. Cell Biol. 2020, 21, 363–383. [Google Scholar] [CrossRef] [PubMed]
  86. Ferrer-Sueta, G.; Campolo, N.; Trujillo, M.; Bartesaghi, S.; Carballal, S.; Romero, N.; Alvarez, B.; Radi, R. Biochemistry of peroxynitrite and protein tyrosine nitration. Chem. Rev. 2018, 118, 1338–1408. [Google Scholar] [CrossRef]
  87. Sheng, Y.; Abreu, I.A.; Cabelli, D.E.; Maroney, M.J.; Miller, A.F.; Teixeira, M.; Valentine, J.S. Superoxide dismutases and superoxide reductases. Chem. Rev. 2014, 114, 854–918. [Google Scholar] [CrossRef]
  88. Quijano, C.; Castro, L.; Peluffo, G.; Valez, V.; Radi, R. Enhanced mitochondrial superoxide in hyperglycemic endothelial cells: Direct measurements and formation of hydrogen peroxide and peroxynitrite. Am. J. Physiol. Heart Circ. Physiol. 2007, 293, H3404–H3414. [Google Scholar] [CrossRef]
  89. Asada, K.; Kanematsu, S. Reactivity of Thiols with Superoxide Radicals. Agric. Biol. Chem. 1976, 40, 1891–1892. [Google Scholar] [CrossRef]
  90. Kwon, B.G.; Kim, J.-O.; Kwon, J.-K. An Advanced Kinetic Method for HO2∙/O2-∙ Determination by Using Terephthalate in the Aqueous Solution. Environm. Engin. Res. 2012, 17, 205–210. [Google Scholar] [CrossRef]
  91. Bartlett, D.; Church, D.F.; Bounds, P.L.; Koppenol, W.H. The kinetics of the oxidation of L-ascorbic acid by peroxynitrite. Free Radic. Biol. Med. 1995, 18, 85–92. [Google Scholar] [CrossRef] [PubMed]
  92. Hakim, T.S.; Sugimori, K.; Camporesi, E.M.; Anderson, G. Half-life of nitric oxide in aqueous solutions with and without haemoglobin. Physiol. Meas. 1996, 17, 267–277. [Google Scholar] [CrossRef]
  93. Zhong, Z.J.; Yao, Z.P.; Shi, Z.Q.; Liu, Y.D.; Liu, L.F.; Xin, G.Z. Measurement of intracellular Nnitric oxide with a quantitative mass spectrometry probe approach. Anal. Chem. 2021, 93, 8536–8543. [Google Scholar] [CrossRef]
  94. Hetrick, E.M.; Schoenfisch, M.H. Analytical chemistry of nitric oxide. Annu. Rev. Anal. Chem. 2009, 2, 409–433. [Google Scholar] [CrossRef]
  95. Thomas, D.D.; Ridnour, L.A.; Isenberg, J.S.; Flores-Santana, W.; Switzer, C.H.; Donzelli, S.; Hussain, P.; Vecoli, C.; Paolocci, N.; Ambs, S.; et al. The chemical biology of nitric oxide: Implications in cellular signaling. Free Radic. Biol. Med. 2008, 45, 18–31. [Google Scholar] [CrossRef]
  96. Ford, E.; Hughes, M.N.; Wardman, P. Kinetics of the reactions of nitrogen dioxide with glutathione, cysteine, and uric acid at physiological pH. Free Radic. Biol. Med. 2002, 32, 1314–1323. [Google Scholar] [CrossRef]
  97. Augusto, O.; Bonini, M.G.; Amanso, A.M.; Linares, E.; Santos, C.C.; De Menezes, S.L. Nitrogen dioxide and carbonate radical anion: Two emerging radicals in biology. Free Radic. Biol. Med. 2002, 32, 841–859. [Google Scholar] [CrossRef]
  98. Koppenol, W.H.; Moreno, J.J.; Pryor, W.A.; Ischiropoulos, H.; Beckman, J.S. Peroxynitrite, a cloaked oxidant formed by nitric oxide and superoxide. Chem. Res. Toxicol. 1992, 5, 834–842. [Google Scholar] [CrossRef]
  99. Quijano, C.; Alvarez, B.; Gatti, R.M.; Augusto, O.; Radi, R. Pathways of peroxynitrite oxidation of thiol groups. Biochem. J. 1997, 322 Pt 1, 167–173. [Google Scholar] [CrossRef]
  100. Kissner, R.; Nauser, T.; Bugnon, P.; Lye, P.G.; Koppenol, W.H. Formation and properties of peroxynitrite as studied by laser flash photolysis, high-pressure stopped-flow technique, and pulse radiolysis. Chem. Res. Toxicol. 1997, 10, 1285–1292. [Google Scholar] [CrossRef] [PubMed]
  101. Radi, R. Peroxynitrite, a stealthy biological oxidant. J. Biol. Chem. 2013, 288, 26464–26472. [Google Scholar] [CrossRef] [PubMed]
  102. Radi, R. Oxygen radicals, nitric oxide, and peroxynitrite: Redox pathways in molecular medicine. Proc. Natl. Acad. Sci. USA 2018, 115, 5839–5848. [Google Scholar] [CrossRef] [PubMed]
  103. Bartesaghi, S.; Radi, R. Fundamentals on the biochemistry of peroxynitrite and protein tyrosine nitration. Redox Biol. 2018, 14, 618–625. [Google Scholar] [CrossRef]
  104. Halliwell, B. Free radicals and antioxidants: A personal view. Nutr. Rev. 1994, 52, 253–265. [Google Scholar] [CrossRef]
  105. Bianconi, E.; Piovesan, A.; Facchin, F.; Beraudi, A.; Casadei, R.; Frabetti, F.; Vitale, L.; Pelleri, M.C.; Tassani, S.; Piva, F.; et al. An estimation of the number of cells in the human body. Ann. Hum. Biol. 2013, 40, 463–471. [Google Scholar] [CrossRef] [PubMed]
  106. Majima, H.J.; Oberley, T.D.; Furukawa, K.; Mattson, M.P.; Yen, H.-C.; Szweda, L.I.; St Clair, D.K. Prevention of mitochondrial injury by manganese superoxide dismutase reveals a primary mechanism for alkaline-induced cell death. J. Biol. Chem. 1998, 273, 8217–8224. [Google Scholar] [CrossRef]
  107. Itoh, K.; Chiba, T.; Takahashi, S.; Ishii, T.; Igarashi, K.; Katoh, Y.; Oyake, T.; Hayashi, N.; Satoh, K.; Hatayama, I.; et al. An Nrf2/small Maf heterodimer mediates the induction of phase II detoxifying enzyme genes through antioxidant response elements. Biochem. Biophys. Res. Commun. 1997, 236, 313–322. [Google Scholar] [CrossRef]
  108. Itoh, K.; Wakabayashi, N.; Katoh, Y.; Ishii, T.; Igarashi, K.; Engel, J.D.; Yamamoto, M. Keap1 represses nuclear activation of antioxidant responsive elements by Nrf2 through binding to the amino-terminal Neh2 domain. Genes Dev. 1999, 13, 76–86. [Google Scholar] [CrossRef]
  109. Gross, E.; Bedlack, R.S., Jr.; Loew, L.M. Dual-wavelength ratiometric fluorescence measurement of the membrane dipole potential. Biophys. J. 1994, 67, 208–216. [Google Scholar] [CrossRef]
  110. Sarkar, S.; Chattopadhyay, A. Membrane dipole potential: An emerging approach to explore membrane organization and function. J. Phys. Chem. B 2022, 126, 4415–4430. [Google Scholar] [CrossRef] [PubMed]
  111. Salemi, G.; Gueli, M.; D’Amelio, M.; Saia, V.; Mangiapane, P.; Aridon, P.; Ragonese, P.; Lupo, I. Blood levels of homocysteine, cysteine, glutathione, folic acid, and vitamin B12 in the acute phase of atherothrombotic stroke. Neurol. Sci. 2009, 30, 361–364. [Google Scholar] [CrossRef] [PubMed]
  112. Richie, J.P., Jr.; Skowronski, L.; Abraham, P.; Leutzinger, Y. Blood glutathione concentrations in a large-scale human study. Clin. Chem. 1996, 42, 64–70. [Google Scholar] [CrossRef] [PubMed]
  113. Thomas, D.D.; Flores-Santana, W.; Switer, C.H.; Wink, D.A.; Ridnour, L.A. Impact of cell signaling processes. In Nitric Oxide Biology and Pathology, 2nd ed.; Section I. Chemical Biology, Chapter 1, Determinants of nitric oxide chemistry; Ignapro, L.J., Ed.; Academic Press: Amsterdam, The Netherlands, 2010; pp. 3–26. [Google Scholar]
  114. Lynch, R.E.; Fridovich, I. Permeation of the erythrocyte stroma by superoxide radical. J. Biol. Chem. 1978, 253, 4697–4699. [Google Scholar] [PubMed]
  115. Porcelli, A.; Ghelli, A.; Zanna, C.; Pinton, P.; Rizzuto, R.; Rugolo, M. pH difference across the outer mitochondrial membrane measured with a green fluorescent protein mutant. Biochem. Biophys. Res. Commun. 2005, 326, 799–804. [Google Scholar] [CrossRef]
  116. Mondal, P.; Ishigami, I.; Gérard, E.F.; Lim, C.; Yeh, S.R.; de Visser, S.P.; Wijeratne, G.B. Proton-coupled electron transfer reactivities of electronically divergent heme superoxide intermediates: A kinetic, thermodynamic, and theoretical study. Chem. Sci. 2021, 12, 8872–8883. [Google Scholar] [CrossRef]
  117. De Grey, A.D. HO2•: The forgotten radical. DNA Cell Biol. 2002, 21, 251–257. [Google Scholar] [CrossRef]
  118. Cordeiro, R.M. Reactive oxygen species at phospholipid bilayers: Distribution, mobility and permeation. Biochim. Biophys. Acta 2014, 1838 Pt B, 438–444. [Google Scholar] [CrossRef]
  119. Tikunov, A.; Johnson, C.B.; Pediaditakis, P.; Markevich, N.; Macdonald, J.M.; Lemasters, J.J.; Holmuhamedov, E. Closure of VDAC causes oxidative stress and accelerates the Ca2+-induced mitochondrial permeability transition in rat liver mitochondria. Arch. Biochem. Biophys. 2010, 495, 174–181. [Google Scholar] [CrossRef]
Figure 1. In the mitochondria, 2~3% of electrons leak from the electron transport chain (ETC), and then oxygen traps the electrons, turning them into superoxide anions (O2•−), and subsequently various ROS are produced: •OH, 1O2, H2O2, O2•−, HO2•, •NO, •NO2, ONOO, and ONOOH. In the intermembrane space, ten times higher amounts of H+ (protons) exist compared to those in the matrix. Among the ROS, ONOO and HOO• (HO2•) can couple with H+, and ONOOH and HOO• are produced and penetrate through the membrane, entering the cytosol to initiate intracellular signals, such as NF-κB and Nrf2.
Figure 1. In the mitochondria, 2~3% of electrons leak from the electron transport chain (ETC), and then oxygen traps the electrons, turning them into superoxide anions (O2•−), and subsequently various ROS are produced: •OH, 1O2, H2O2, O2•−, HO2•, •NO, •NO2, ONOO, and ONOOH. In the intermembrane space, ten times higher amounts of H+ (protons) exist compared to those in the matrix. Among the ROS, ONOO and HOO• (HO2•) can couple with H+, and ONOOH and HOO• are produced and penetrate through the membrane, entering the cytosol to initiate intracellular signals, such as NF-κB and Nrf2.
Biomolecules 14 00128 g001
Table 1. Calculated dipole moment and experimental permeability coefficient of ROS and RNS.
Table 1. Calculated dipole moment and experimental permeability coefficient of ROS and RNS.
ROS or RNSCalculated Dipole Moment/DPermeability Coefficient/cm s−1
H2OWater1.892.3 × 10−3 [48]
H2O2Hydrogen peroxide0.00 (permeable)6.1 × 10−3, 6.6 × 10−4 [49]
•OHHydroxyl radical1.67
O2•−Superoxide01 × 10−6 (pH 7.3, 25 °C) [50]
(7.6 + 0.3) × 10−8 [51]
HO2Hydroperoxyl radical2.234.9 × 10−4 [51]
•NONitric oxide0.14 (permeable)93 (20 °C) [52]
•NO2Nitrogen dioxide0.35[53] and discussion in the text
ONOOPeroxynitrite2.14Through anion exchanger [54]
8.0 × 10−4 [55]
ONOOHPeroxynitrous acid1.774–13 × 10−4 [56,57,58]
Table 2. Predictive performance of mitochondria-originating reactive oxygen species.
Table 2. Predictive performance of mitochondria-originating reactive oxygen species.
ROS or RNSHalf-Life TimeAmount/CellDiffusion Distance (µm)Permeability Coefficients (Pm) (cm s−1)Eo′; One-Electron Reduction Potential (V) at pH 7pKak (AscH)/M−1 s−1k (GSH)/M−1 s−1
H2OWater––––––3.3 × 10−3 (EYPC) [59]−2.87 [60]
−2.87 [61]
15.7 [62]––––
O2Oxygen––––––12 (DMPC) [59]
125 (DMPC) [59]
114 (DOPC) [59]
157 (POPC) [59]
50 (EYPC: 30% Chol) [59]
38 (RBC human) [59]
21 (CHO cells) [59]
42 (CHO cells) [59]
−0.18 (pH 7) [60]
−0.33 [61]
−0.16 [63]
−0.18 (pH 7, 25 °C) [64]
––––––
•OHPeroxynitrous acid10−9 s [65]
10−9~10−6 s (diffusion-controlled reactivity) [66]
10−10 s [67]
10−9 s (1 M, 37 °C) [68]
10−9 s [69]
––3 Å [70]
A large flux of hydroxyl radicals would be required to inactivate a substantial fraction of any biological target [70]
0.02 (GSH+) [71]
––+2.32 (pH 7) [60]
+2.31 [61]
+2.31 (pH 7, 25 °C) [64]
+2.31 [72]
+2.31 (pH 7) [73]
11.9 [62]
11.6 [74]
1.1 × 1010 (pH 7.4) [61]1.0 × 1010 [72]
1.64 ± 0.01 × 1010 [74]
1 × 109 [75]
8.8 × 109 (pH 1.0) [76]
9.0 × 109 (pH 7.6) [77]
1 × 1010 [78]
1.1 × 1010 (oxidized GSH) [79]
1.4 × 1010 (reduced GSH) [79]
1.4 ± 0.1 × 1010 (pH 7.8) [80]
4.4 ± 0.5 × 1010 (pH 10.6) [80]
2.3 × 1010 [81]
4.4 ± 0.5 × 1010 (pH 10.6) [82]
1O2Singlet oxygen10−6 s [65]
10−6 s [67]
10−6 s (solvent, 37 °C) [68]
10−5 s [69]
10−9~10−6 s [83]
––––––+0.81 (pH 7, 25 °C) [64]––3.2 × 108 [83]
1.8 × 108 [84]
9.39 ± 0.07 × 108 [74]
H2O2Hydrogen peroxideStable [65]
Stable)
Stable, decomposed by catalase and GSH peroxidase and by EDTA and ADP [68]
Enzymatic [69]
18.1 ± 2.7 min [82]
Physiological condition (proliferation/differentiation/migration/angiogenesis): 0.001~0.1 µM)
Stress responses/adaptation (e.g.,
NRF2): 0.05~5.0 µM [85]
Inflammation/fibrogenesis/tumor growth/metastasis: 0.01~10.0 µM [85]
Growth arrest/cell death: 1.0~10.0 µM [85]
1600 (GSH+) [78]6 × 10−4 (RBC horse) [59]
3 × 10−3 (peroxisome rat liver) [59]
1.2 × 10−2 (RBC rat) [59]
2 × 10−4 (Jurkat T cells) [59]
3.6 × 10−4 (Chara coralina) [59]
1.6 × 10−3 (Escherichia coli) [59]
4 × 10−4 (PC12 cells) [59]
1.6 × 10−3 (HUVEC cells) [59]
1.1 × 10−3 (IMR-90 cells) [59]
4.4 × 10−4 (HeLa cells) [59]
+0.39 (pH 7) [60]
+0.32 [61]
+1.77 [72]
+1.8 [78]
+0.39 (pH 7, 25 °C) [86]
11.6 [62]
11.75 (pH 7.2) [78]
––9 × 10−1 [72]
9 × 10−1 [73]
9 × 10−1 (pH 7.4, 37 °C) [78]
8.7 × 10−1 [81]
O2•−Superoxide10−6 s [65]
1 s (pH 10) [66]
10−6 s (diffusion-controlled reactivity) [66]
10−6 s [67]
The lifetime of superoxide in a cellular environment in water would be expected to be very short, too short to permit diffusion for great distances [68]
Enzymatic [69]
3000 ms (10−6 M) [87]
175 ms (10−6 M + SOD 10−9 M) [87]
hours (10−9 M) [87]
175 ms (10−9 M + SOD 10−9 M) [87]
0.175 ms (10−9 M + SOD 10−6 M) [87]
28.4 pM (normal condition)/mitochondria [88]
Formation rate (to 6 µM/s) [88]
MnSOD-catalyzed dismutation (k = 2 × 109 M−1 s−1) [88]
9.15 × 10−8 pmol production/s/mitochondria *
690 nM production/s/mitochondria *
5.5 × 104 superoxide molecules /s/mitochondria *
––2.1 × 10−6 (SBPC) [59]
7.6 × 10−8 (EYPC) [59]
+0.94 [72]
+0.94 [73]
––1 × 105 (pH 7.4) [61]
2.7 × 105 (pH 7.4) [61]
~10 to 103 [72]
2 × 102 [81]
1.1 ± 0.04 × 103 [74]
6.7 × 105 (reduced GSH) (pH 7.8) [89]
HO2Hydroperoxyl radical51~422 s (pH 2~10) [90]
HO2• radicals in organic or lipophilic media could have a longer half-life. The half-life of superoxide cannot be calculated unless the concentrations of SOD and all reactive substrates are known [67]
9.15 × 10−8 pmol production/s/mitochondria *
690 nM production/s/mitochondria *
5.5 × 104 superoxide molecules/s/mitochondria *
––4.9 × 10−4 (EYPC) [59]+1.05 (pH 7) [60]
+1.06 [72]
+1.05 (pH 7, 25 °C) [86]
4 [62]
4.8 [89]
4.8 [90]
4.8 [91]
1 × 105 (pH 7.4) [61]
2.7 × 105 (pH 7.4) [61]
––
•NONitric oxidems to s depending on the available concentration of O2, otherwise stable [66]
Second [67]
1~10 s [69]
445 s [92]
•NO:1200 nM in saline:
binding with Hb: 2 × 105 M−1 s−1 [92]
Seconds [93]
pM~μM [93]
pM~µM in physiological milieu [94]
cGMP-mediated
processes; <1~30 nM [95]
Akt phosphorylation; = 30~100 nM
stabilization of HIF-1α; = 100~300 nM [95] phosphorylation of p53; > 400 nM [95]
nitrosative stress; 1 μM [95]
––73 (EYPC) [59]
66 (EYPC: 30% Chol) [59]
18 (RBC human) [59]
–0.52 (pH 7) [60]
–0.35 [63]
–0.80 [72]
–0.80 [73]
––––Nondetectable [72]
1.0 × 101 [75]
•NO2Nitrogen dioxide Second [67]
<10 µs [96]
Typically 0.2~0.3 µM [96]0.4 (GSH+) [78]
0.2 in the cytoplasm [96]
<0.8 in blood plasma [96]
~5 (EYPC) [59]+1.04 (pH 7) [60]
+1.04 [63]
+1.04 [72]
+1.04 [73]
––1.8 × 107 [96]
3.5 × 108 [96]
3.54 × 106 (pH 5.4~6.5, 55 °C) [97]
3.0 × 107 [72] 2.2 × 107 [75]
3 × 107 [78]
2 × 107 [81]
~2 × 107 [96]
ONOOPeroxynitrite0.8 s (pH 7.4) [64]
10−3 s [67]
0.05~1 s [69]
0.8 s (pH 7.4) [98]
0.9 s [98]
Stable [98]
Relatively stable [99]
Less than 1 s (pH 7.4, 37 °C) [99]
0.8 s (pH 7.4) [100]
A total peroxynitrite and peroxynitrous acid concentration that
exceeds 0.1 mM [101]
60 (GSH+) [78]
0.42 [101]
–––––––7 × 102 [78]
2.35 ± 0.04 × 102, 25 °C [91]
6.6 × 102 (pH 7.4, 25 °C) [71]
7.0 × 102 [73]
6.6 × 102 [75]
1.36 × 103 (pH 7.4, 37 °C) [78]
2.81 × 102 (pH 5.75, 37 °C) [100]
ONOOHPeroxynitrous acidFairly stable [67]
0.90 s, 25 °C [98]
Less than 1 s at physiological pH and 37 °C [99]
0.6 s; 1.13 s−1 in phosphate buffer (pH 7.4, 37 °C) [102]
A total peroxynitrite and peroxynitrous acid concentration that exceeds 0.1 mM [101]––8 × 10−4 (DMPC) [59]
1.3 × 10−3 (EYPC) [59]
6.3 × 10−4 (DMPC) [59]
4 × 10−4 (DPPC) [59]
+1.40 [72]6.8 [86]
6.8 [90]
6.8 [98]
6.8 [103]
––6.6 × 102 [72]
1.35 × 103 [81]
RO•Alkoxyl radicals10−6 s [67]
10−6 s (100 mM) [68]
10−6 s [69]
––––+1.60 [61]
+1.60 [72]
~+1.60 [73]
––1.6 × 109(pH 7.4) [61]2.76 ± 0.15 × 106 [74]
ROO•Peroxyl radicalsSeconds to hours depending on conditions [66]
17 s [67]
7 s (100 mM, 37 °C) [68]
7 s [69]
––––+1.00 [61]
+0.77~1.44 [73]
+1.00 [72]
––1-2 × 106 (pH 7.4) [61]––
Abbreviations: Chol, cholesterol; DLPC, dilauroylphosphatidylcholine; DMPC, dimyristoylphosphatidylcholine; DOPC, dioleoylphosphatidylcholine; DPPC, dipalmitoylphosphatidylcholine; EYPC, egg yolk phosphatidylcholine; POPC, palmitoyloleoylphosphatidylcholine; RBC, red blood cell. * Assuming a 70 kg man, O2 consumption/day is estimated as 14.7 mol/day [14,104]. Assuming that 2% of electrons leak from the mitochondrial electron transport chain (ETC) and that these are trapped by oxygen and made into superoxide, the superoxide production from the ETC is thus estimated as 3402.8 nmol/s. The number of cells/body is estimated as 3.72 × 1013 [105]. Thus, superoxide production is calculated as 5.51 × 107 mol/s/mitochondria. Assuming that the volume of mitochondria is 1.32 × 107 m3, then considering mitochondrial volume, 1.32 × 10−16 m3, superoxide production is estimated as 6.90 × 102 µmol/s/m3. It is noted that this number is the amount of superoxide produced and that superoxide is modified by other molecules and enzymes, and thus the amount of superoxide existing in cells is much less.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Masuda, D.; Nakanishi, I.; Ohkubo, K.; Ito, H.; Matsumoto, K.-i.; Ichikawa, H.; Chatatikun, M.; Klangbud, W.K.; Kotepui, M.; Imai, M.; et al. Mitochondria Play Essential Roles in Intracellular Protection against Oxidative Stress—Which Molecules among the ROS Generated in the Mitochondria Can Escape the Mitochondria and Contribute to Signal Activation in Cytosol? Biomolecules 2024, 14, 128. https://doi.org/10.3390/biom14010128

AMA Style

Masuda D, Nakanishi I, Ohkubo K, Ito H, Matsumoto K-i, Ichikawa H, Chatatikun M, Klangbud WK, Kotepui M, Imai M, et al. Mitochondria Play Essential Roles in Intracellular Protection against Oxidative Stress—Which Molecules among the ROS Generated in the Mitochondria Can Escape the Mitochondria and Contribute to Signal Activation in Cytosol? Biomolecules. 2024; 14(1):128. https://doi.org/10.3390/biom14010128

Chicago/Turabian Style

Masuda, Daisuke, Ikuo Nakanishi, Kei Ohkubo, Hiromu Ito, Ken-ichiro Matsumoto, Hiroshi Ichikawa, Moragot Chatatikun, Wiyada Kwanhian Klangbud, Manas Kotepui, Motoki Imai, and et al. 2024. "Mitochondria Play Essential Roles in Intracellular Protection against Oxidative Stress—Which Molecules among the ROS Generated in the Mitochondria Can Escape the Mitochondria and Contribute to Signal Activation in Cytosol?" Biomolecules 14, no. 1: 128. https://doi.org/10.3390/biom14010128

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop