Discussion
The lungs are commonly affected by ROS, and almost all cell types of this organ can produce these highly reactive molecules. Under basal conditions, ROS are synthesized by NADPH oxidase and act as second messengers to activate numerous intracellular proteins and enzymes involved in physiological and pathological stages [
19,
32]. In the lungs, pulmonary fibroblasts produce ROS, especially after stimulation by inflammatory cytokines [
33] with a diverse range of stimuli. However, uncontrolled ROS production could directly cause DNA damage, morphologic transformations in cells, and lung injury if the antioxidant system cannot control the generation of ROS [
34,
35]. Besides, in the environment, many toxicants can induce these prooxidant forces [
36,
37]. Nevertheless, there is no previous information about ROS production in human lung fibroblasts induced by HMs, a by-product of disinfection procedures to produce drinking water. The current study found that the toxicants understudy could induce ROS
in vitro, particularly O
2•. Similar findings were also observed in peripheral blood mononuclear cells (PBMCs) of
Cyprinus carpio carpio i.p dosed with CH
2Cl
2, CHCl
3 and BrCHCl
3 [
19]. Despite this lack of information, we can propose that the generation of ROS in human fibroblasts treated with halomethanes is due to a series of events, not necessarily consecutive, as well as its feedback with prooxidant forces and oxidative damage summarized as follows:
1) Biotransformation processes are fundamental in the generation of ROS.
2) ROS generation is not directly regulated by mitochondrial activity.
3) The higher the number of halogens involved, the greater the probability of ROS generation.
4) Some physicochemical properties, such as the electronegativity of the metabolites generated during biotransformation, influence the ROS generation rate.
5) ROS induction depends mainly on specific and non-specific antioxidant defenses. The arguments presented are discussed below. ROS could be generated more by an independent pathway than mitochondrial complex I activity [
19], such as the case of decoupling the flow of electrons during redox processes of cytochrome P450 isoenzymes (i.g. cytochrome P450 isoform 2E1 “CYP 2E1”). Other enzymes are involved in the HM’s bioactivation process, as in the isoform
tetha of the glutathione S-transferase (GSTT) [
8,
19,
38,
39,
40]. Regarding the topic, the obtained results showed that the degree of chlorination of HMs is a critical factor in the induction of ROS in MRC-5 cells. In the first step of the biotransformation of the HMs, the three compounds under study can induce similar amounts of O
2• mediated by the activity of CYP 2E1 [
19]. Next to this step, the HMs possessing three halogens suffer reductions, dehalogenations, and reductive dehalogenation mediated by some enzymes such as GSTT and other enzymes [
8,
41], allowing ROS generation [
19]. However, the higher levels of O
2• elicited by CHCl
3 compared with BrCHCl
2 could be explained by HCl release, which possesses an electronegativity value (δ=1.06) higher than HBr (δ=0.86) released from BrCHCl
2 after dehalogenation processes [
19]. Contrary to these halomethanes, the amounts of O
2• were the lowest in the cells treated with CH
2Cl
2. This response could be explained by a lesser number of reductive dehalogenations involved in the biotransformation processes that this compound undergoes [
19]. Also, on the final step of biotransformation of CHCl
3 mediated by GSTT [
42], the release of H
+ after reductive dehalogenation probably enhances the respiratory rate and proton motive force [
19], allowing ROS generation by mitochondrial complex I activity as well as by mitochondrial electron transport chain. In contrast, the significant diminution of O
2•levels in MRC-5 fibroblasts exposed to high concentrations of CH
2Cl
2 is feasible to explain by enzymatic dismutation of H
2O
2 through SOD activity as demonstrated in the current study. The activity of SOD plays a crucial role in the antioxidant system in human lungs [
6,
43,
44]. This enzyme plays a vital role in the dismutation of O
2• into H
2O
2 and prevents chain reactions that induce •OH and peroxynitrite radicals (ONOO-). However, the O
2• may further convert into OH• in the presence of metal ions via Fenton and Haber-Weiss reactions [
45]. The excessive production of ROS in the lungs can disrupt the oxidant/antioxidant balance, vital in developing pathologies in the airways [
46,
47,
48,
49]. Besides, excessive ROS levels can alter critical cellular components such as proteins, unsaturated lipids, and DNA by exposure to free radicals, thus compromising cell homeostasis [
45,
50].
On the other hand, the H
2O
2 generation by dismutation of O
2• could occur spontaneously (non-catalyzed) and too mediated by the activity of SOD [
45]. The highest levels of H
2O
2 observed in MRC-5 cells treated with CH
2Cl
2 at 10
-20 mol were not accompanied by the highest concentration of its precursor (O
2•), which reinforces the hypothesis about a non-catalyzed dismutation, which probably occurred early in the present study, nor the enzymatic activity of SOD had a role, as has been previously reported [
45]. In the current study, the H
2O
2 was related to O
2• levels in MRC-5 fibroblasts exposed to CHCl
3 but not with SOD activity. This finding suggests a non-catalyzed dismutation of O
2• to H
2O
2 in the presence of H
+ generated during biotransformation at the late stage [
19] as observed - for MRC-5 cells treated with CH
2Cl
2. If H
2O
2 levels remain stable in the intracellular medium and escape to the enzymatic action of either CAT or GPx or some non-enzymatic element of the antioxidant system. This compound can activate signaling pathways that stimulate diverse cellular processes such as cell proliferation, differentiation, migration, or apoptosis [
19,
51,
52,
53,
54,
55]. In contrast, levels of H
2O
2 in MRC-5 fibroblasts treated with CH
2Cl
2 and with BrCHCl
2 were statistically related to O
2• concentration and showed a similar response in the low treatment. However, in the fibroblasts exposed to BrCHCl
2, the diminution of H
2O
2 was linked with the activity of CAT and GPx. Despite the crucial role of CAT in the dismutation of H
2O
2, the role of GPx in maintaining the redox status of lung cells exposed to BrCHCl
2 stands out. This enzyme also removes organic hydroperoxides and participates in glutathione’s redox cycle, which is abundant in the alveolar epithelial lining fluid [
56]. Regardless of the enzymes involved in the antioxidant defense activity, ROS induction elicited lipid peroxidation
in vitro in human lung fibroblasts.
In addition to the cytotoxicity elicited by HMs previously reported in MRC-5 cells [
8], the current study shows that CHCl
3-induced lipid peroxidation evaluated as TBARS on this cell line exposed to higher concentrations of this toxicant. The induction of lipid peroxidation was dependent on halomethane concentrations in MRC-5 cells. Indeed, the damage is induced by the remaining amount of ROS generated and not detoxified through the antioxidant defense, as previously discussed. However, this imbalance between the prooxidant forces (ROS) and antioxidants (SOD, CAT, and GPx) produces a series of radical reactions, including,
1) the initiation caused by the predominance of pro-oxidant forces,
2) the propagation of free radicals to reach deep levels of damage and destruction of the lipid bilayer and
3) formation of polymers between free radicals might occur in the best of cases. The background on the subject is set out below. Lipid peroxidation is an oxidative process that involves the deterioration of the lipid bilayer mediated by hydroxyl radical (
•OH) to form carbon-carbon double bonds at the initial stage. - The
•OH is considered the most critical oxidizing chemical species in the initiation of this process [
57]. However, through chain reactions, other chemical species such as alkyl hydroperoxides (LOOH), alkyl peroxyl radicals (LOO•), and alkoxyl radicals (LO) may also be formed [
45]. The level of lipid peroxidation induced by CHCl
3 was related to H
2O
2 concentration. This oxidation probably implies damages in lung fibroblast membranes (both in the cell and organelles) that can reach several levels of gravity, such as altering the fluidity of the lipid bilayer on its functionality even on its breaking. The consequences can be diverse, but it is possible to emphasize modifications in hormone receptors and proteins involved in signal transduction pathways [
45]. A previous report found that a high concentration of CHCl
3 elicited a decrease in cell growth of MRC-5 cells in all concentrations assessed; however, higher concentrations of this toxicant induced cytotoxicity [
8]. It is possible to speculate that lipid peroxidation elicited by CHCl
3 was directly related to various damages, including diminution of lung fibroblast growth and cytotoxicity; however, the lungs could counteract these damages. In this regard, Kornbrust and Mavis [
58] reported that lipid peroxidation is about 25 to 50 times higher in the lungs and heart, which are highly oxygenated tissues, than in the liver, kidneys, testes, and brain. Indeed, in lung cells, lipid peroxidation is a process that occurs regularly, but pulmonary fibroblasts are highly effective in remodeling the cells and repairing the damaged tissues. In this regard, the lung can regenerate new cells efficiently after injury. Lung regeneration comprises progenitor cell activation and cell replacement through the proliferation of intact cells [
59,
60,
61]. Interestingly, the higher concentration of CHCl
3 induced a diminution in TBARS levels in MRC-5 cells about the treatment of 10
-6 mol, which could be due to GPx activity. However, a significant inverse relationship between TBARS and CAT activity also stands out. This response suggests that oxidative damage was provoked by H
2O
2, whose observed decrease to 10
-4 for 10
-6 was consistent with increased CAT activity and a decrease in lipid peroxidation. In contrast, in human lung fibroblasts exposed to BrCHCl
2, only the median concentrations elicited an increase in lipid peroxidation, which was related to levels of ROS. However, higher concentrations of this toxicant cause a significant reduction in this oxidative damage, which could be related to increase GPx activity even though a significant relationship was not found. Considering the lung’s capacity to repair damaged cells or regenerate new cells after an oxidative injury elicited by ROS, observing the relationships among other biomarkers involved in regulating the cell cycle case of NF-κB is probably helpful.
Remarkably, only the HMs possessing three halogens (BrCl2CH and Cl3CH) were able to increase the levels of phospho-NF-κB/p65 protein at Ser536; however, the magnitude of this induction and relationships with the biomarkers of oxidative stress response was different. Although the quantitative structure-activity relationships (QSARs) would offer some explanations about the differentiated response of NF-κB translocation to the nucleus, the molecular docking tools provide detailed information about the disruption of the IκBα- NF-κB/p65 complex by the effects of HMs treatment.
The
in-silico results using the software molecular operating environment (MOE) showed that the estimated free energy of binding with the p65 subunit of NF-κB was -7.0 kcal/mol for BrCHCl
2 and Cl
3CH binds with -6.5 kcal/mol. The interaction energy values implied that these ligands could bind with the nuclear factor with sufficient bond strength, which coincided with the higher phosphorylation levels of NF-κB/p65 in a concentration-dependent manner, particularly for BrCHCl
2. The bonds of chloride and bromide from HM have occurred in the C-terminal domain of p65 of NF-κB/p65; this is a relevant finding because the C-terminal transcription activation domain (TAD) regulates the gene expression, and is highly active [
62]. Notably, we found that Cl
1, Cl
2, and Br are in contact with four contiguous amino acid residues (Thr
191, Ala
192, Glu
193, Leu
194) into region RHD (Rel homology domain), which is involved in DNA-binding. In this sense, it has been reported that the IPT domain is the dimerization segment of the NF-κB/p65 and related transcription factors [
63,
64], covering 101 a.a. from Thr
191 to Asp
291. Also, the IPT domain acts as a dominant regulator of diverse stressful conditions, including physical stress, oxidative stress, and exposure to certain chemicals [
65,
66,
67]. Besides, there are critical conserved residues in this domain: Arg198, Glu211, Leu215, and Cys216 in the dimer interface, which is dominated by hydrophobic interactions close to residues of the bond of Cl
1, Cl
2 and bromide anions [
68].
On the other hand, post-translational modifications such as phosphorylation, acetylation, and methylation of transcription factors may affect NF-κB transcriptional activity [
69]. Remarkably, the phosphorylation of IκB proteins followed by ubiquitination and degradation by proteasomes releases the NF-κB/p65 homodimer, allowing the bind of p65 with DNA as a homodimer or as heterodimers [
70,
71,
72]. In this context, there are several phosphorylation sites in the IPT domain close to the binding residues of Cl
1, Cl
2 and bromide anions of the HMs, i.e., residues 254, 276, 281, particularly stands out the contact with Ser
281. A previous study on a site-directed mutagenesis screen for potential phosphorylation sites within the p65 RHD identified S205, S276, and S281 essential for p65 transcriptional activity [
17,
73]. These studies and current results suggest that phosphorylation of the NF-κB subunits profoundly affects the function of NF-κB. Certain phosphorylation events contribute to the selective regulation of NF-κB transcriptional activity in a gene-specific manner [
74]. In agreement with our result, the binding of the backbone carbon of CHCl
3 with Ser281 may affect the complex IκBα/p65 conformation, allowing the hiper-proliferation of MRC-5 as documented in a previous study [
8]. In this regard, it has been reported that the phosphorylation of p65 induces a conformational change, which impacts p65 ubiquitination and stability and protein-protein interactions. However, NF-κB/p65 is a highly dynamic and flexible protein not fixed in a globular and well-ordered shape but can change dynamically after NF-κB activation [
75]. We consider that the bond of Cl
1, Cl
2 and bromide of BrCHCl
2 and CHCl
3 would affect the structure of the p65 homodimer as well as its interaction with IκBα, and the dissociation of the Iκβα-NF-κB complex involved in the activation of target genes transcription. However, to evaluate the effects of these HMs in human fibroblasts, the influence of some of the more reactive metabolites derived from biotransformation processes must be considered, such as in the case of ROS.
Despite the preceding results showing relationships among oxidative stress response, and NF-κB/p65 activation [
76,
77,
78,
79,
80,
81,
82,
83,
84,
85,
86] there is no previous information about ROS production in human lung fibroblasts induced by HMs. However, in a previous study, ROS was inducted in peripheral blood mononuclear cells (PBMCs) of
Cyprinus carpio carpio i.p treated with CH
2Cl
2, CHCl
3 and BrCHCl
2 was documented [
19]. Results of the current study showed that the toxicants under study induced ROS in vitro with an estimated free energy of binding with NF-κB/p65 of -7.0 kcal/mol for H
20
2 and -7.6 kcal/mol for O
2•. The interactions of these ROS occurred in the IPT domain of NF-κB/p65, suggesting the activation of target genes is involved in the antioxidant response, among others, as shown by Pearson correlation analysis. Certain NF-κB-regulated genes play a major role in regulating the amount of ROS in the cell because ROS have various inhibitory or stimulatory roles in NF-κB signaling. The interactions of ROS with cysteine residues are essential to disturb the NF-кB pathways. For example, hydrogen peroxide inhibits IKK activation by acting on its cysteine residues, eliciting, in this way, inhibitory effects in the catalytic domains of tyrosine phosphatases [
87,
88].
Biochemical analysis in MRC-5 cells treated with HMs and computational study using MOE2019 showed that BrCHCl
2 and CHCl
3 induce the phosphorylation of NF-κB/p65 by activating the IPT domain. The consequence of these interactions is probably linked with the hyper-proliferation of these cells, as previously documented [
8]. However, the ROS generated during the metabolism of these toxicants disturb the activation of the IPT domain involved in the antioxidant response mediated by NF-κB/p65 as observed by negative relationships among ROS with SOD, CAT, and GPx found in cells treated with CHCl
3.