Skip to main content
Erschienen in: Cardiovascular Toxicology 5/2020

Open Access 25.07.2020

The Molecular Mechanism of Aluminum Phosphide poisoning in Cardiovascular Disease: Pathophysiology and Diagnostic Approach

verfasst von: Seyed Farzad Hosseini, Mehdi Forouzesh, Mohsen Maleknia, Samira Valiyari, Mahmood Maniati, Azin Samimi

Erschienen in: Cardiovascular Toxicology | Ausgabe 5/2020

Abstract

Nowadays, poisoning with metal phosphides, especially aluminum phosphide (ALP), is one of the main health threats in human societies. Patients suffer from significant complications due to this type of poisoning, and the heart is one of the main organs targeted by ALP. Therefore, in this study, we discussed the effect of phosphine on cardiac function. This study is based on data obtained from PubMed, between 2002 and 2020. The key keywords included “Aluminum phosphide,” “Oxidative Stress,” “Mitochondria,” “Cardiovascular disease,” and “Treatment.” The results showed that ALP produced reactive oxygen species (ROS) due to mitochondrial dysfunction. ROS production leads to red blood cell hemolysis, decreased ATP production, and induction of apoptosis in cardiomyocytes, which eventually results in cardiovascular disease. Since ALP has the most significant effect on cardiomyocytes, the use of appropriate treatment strategies to restore cell function can increase patients’ survival.
Hinweise
Communicated by Mitzi C. Glover.
The original version of this article was revised: This article published originally with open access. With the author’s decision to step back from Open Choice, the copyright of the article has been changed.
A correction to this article is available online at https://​doi.​org/​10.​1007/​s12012-020-09604-3.

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Introduction

The use of pesticides has led to an increase in the quantity and quality of agricultural products in some developing countries. However, improper use of these chemicals can lead to severe acute and chronic poisoning [1]. Poisoning with pesticides can be intentional, accidental, occupational, or for murder purposes, and it leads to the deaths of about 300,000 people worldwide each year, making it a global health problem [2]. Metal phosphides make a large part of pesticides that are currently used in many countries [3]. These metal phosphides cause a wide range of effects, including electrolyte imbalances, metabolic disorders, cell toxicity, oxidative stress, inhibition of cholinesterase, circulatory failure, and damage to various organs [4]. Aluminum phosphide (ALP) is one of the most widely used metal phosphides and a highly toxic substance with a high mortality rate (40–80%) [5]. The most important causes of death are cardiac shock, severe metabolic acidosis, and hypotension [6]. ALP produces phosphine gas (PH3) after exposure to moisture and acidic stomach conditions, which leads to severe toxicity in poisoned individuals [7]. PH3 is a highly toxic metabolite known as a mitochondrial toxin [8]. However, organs that are mainly dependent on oxidative phosphorylation are quickly affected [9]. So far, the pathophysiological mechanism of ALP poisoning has not been established in humans. However, inhibition of cytochrome C oxidase activity, oxidative stress and Ca2+ and Mg2+ compounds are among the proposed ALP poisoning mechanisms. In addition to inhibiting cytochrome C oxidase, ALP affects other enzymes and consequently it causes severe energy failure. Also, by increasing the production of reactive oxygen species (ROS), it can cause tissue damage synergistically [911]. Cardiomyocytes are one of the main targets for PH3, and cardiac toxicity has been reported as the main cause of death in ALP poisoning cases [12, 13]. Despite supportive care and medical treatment, most patients die after ALP poisoning in the hospital. Therefore, it is necessary to investigate the mechanisms involved in causing tissue damage, especially the heart, as the main organ involved in this poisoning. The diagnosis of ALP poisoning is mainly based on statements from the patients or witnesses, as well as the inspection of remnants or residues. The toxicological analysis of gastric contents and viscera, along with the typical pungent garlicky odor perceptible in many cases upon opening the stomach, may serve as an indication of a potential ALP casualty. Diagnosis of phosphide poisoning can be doubtlessly confirmed by detecting the phosphine gas in samples. Quantitative methods for the detection of phosphine include a simple colorimetric method and gas chromatograph [14]. In this study, we collected scientific data to identify the risk factors for cardiac damage and treatment strategies with the aim of early diagnosis, improvement of therapeutic conditions, and enhanced patient survival.

Mitochondrial Complex as the Main Target of ALP

Mitochondria provide more than 90% of total ATP for eukaryotic cells [15]. Phosphine, by changing in electron transfer chain, reacts with the mitochondrial respiratory chain as the main source of free radical production and interferes with oxidative phosphorylation, leading to high production of ROS and decreased ATP levels. This causes a cell energy crisis. Therefore mitochondria are known to be the main target of phosphine [11, 16].
By producing ROS including superoxide (O20−) and H2O2, cellular oxidative stress acts the same as reactive nitrogen species (RNS) which mainly include NO and peroxynitrite as by-products of a set of enzymes that participate in electron transfer. ROS/RNS potentially damage biological macromolecules, leading to cell death [16]. Although many studies have so far been performed to identify the mechanism of phosphine toxicity, its exact mechanism is still unclear. However, most studies have estimated an increase in oxidative stress and a decrease in antioxidant capacity as the primary mechanism of toxicity [17, 18]. Phosphine acts at the mitochondrial level, and after systemic absorption, impairs the synthesis of enzymes and proteins [1, 19]. Also, its toxicity mechanism involves formation of highly reactive hydroxyl radicals. Phosphine is a highly reactive radical that penetrates the intracellular space and disrupts mitochondrial function by reacting with the mitochondrial respiratory chain as the primary source of ROS production and creating extensive oxidative stress [7]. Evidence suggests that Complex IV and cytochrome C oxidase are the main sites of interaction between phosphine and the electron transport chain, and phosphine inhibits the potential of mitochondrial membrane (Δψm) by inhibiting this enzyme at the site of Complex IV [20]. Besides, phosphine reduces the activity of complexes I and II, which reduce the activity of mitochondrial complexes and inhibit its aerobic respiration, leading to mass production of ROS, impaired ATP synthesis, and energy failure [17, 18]. ROS production resulted by phosphine toxicity is a fatal cause of energy deficiency, and a decrease in ROS production along with a reduction in energy metabolism can increase tolerance to phosphine [21].
Inhibition of cytochrome oxidase by phosphine, along with decreased catalase and peroxidase activity, leads to hydrogen peroxide (H2O2) accumulation and formation of hydroxyl radicals (⋅OH) [11]. In the meantime, ROS can damage or alter mitochondrial DNA, which can result in inefficient and effective respiration and overcoming genes that encode resistance to phosphine toxicity. Suppression of mitochondrial respiration chain genes leads to increased phosphine resistance, and its persistence may be the result of activation of genes that encode the components of the respiratory chain, including Complexes I (NADH/ubiquinone) and III (cytochrome c reductase) [22, 23]. Resistance is stronger when gene complexes III are suppressed, indicating that this complex plays a more critical role in phosphine resistance [24, 25]. In addition to increasing H2O2 production, phosphine performance is associated with increased lipid peroxidation (LPO) following glutathione (GSH) reduction. Reduced concentration of GSH in various tissues in ALP poisoning can also explain cellular damage because GSH catalyzes H2O2 to O2 and H2O, which is way to protect against oxidation [26, 27]. In general, toxicity of PH3 is closely related to the availability and metabolic absorption of oxygen and is increased in the presence of oxygen while decreased under anoxic conditions [4]. Due to its high oxygen consumption, the heart is the organ most vulnerable to ALP poisoning. Cardiac toxicity has been attributed to the increase in free radicals produced by inhibiting respiratory chain complexes. These free radicals cause LPO, DNA damage, and ultimately oxidative stress, targeting the apoptotic process in cardiomyocytes [12, 20]. Finally, in case of PH3 poisoning, the rate of damage to the heart can be assessed by identifying markers related to cardiac damage such as troponin. The efficiency of cardiomyocytes in patients can be increased using appropriate treatment strategies.

Relevance of Intravascular Hemolysis and ALP

Following the uptake of phosphine through gastric mucosa, vascular wall degeneration, hemolysis, and methemoglobinemia (Met-Hb) occur, leading to organ damage [28]. ALP can directly damage blood vessels and the RBC membrane, or by inducing free radicals, it can cause hemoglobinemia and intravascular hemolysis, in which oxidative stress plays a significant role in the formation of these lesions [29, 30].
Exposure to chemicals that oxidize ferrous hemoglobin to ferric form can lead to the production of Met-Hb [31]. Decreased Met-Hb capacity to deliver enough oxygen to tissues could be another reason for multiple organ failure following ALP poisoning [29]. Clinical manifestations of Met-Hb are due to a decrease in oxygen transport capacity, and consequently tissue hypoxia, which helps to worsen the patient's condition [32]. These manifestations depend on the blood concentration, and according to recent studies there is a clear and significant relationship between blood levels of Met-Hb and mortality in poisoned individuals. Intravascular hemolysis can have a more significant impact on the transmission of oxygen to the target tissues [29]. ALP oxidizing properties can oxidize hemoglobin protein and precipitate it as Heinz bodies, leading to intravascular hemolysis [33]. Intravascular hemolysis usually occurs in patients with G6PD deficiency, and when G6PD levels are low, the ability of erythrocytes to produce NADPH is impaired, and cells are prone to hemolysis [34, 35]. Hemolysis is also possibly due to metabolic acidosis, which is a common feature of ALP poisoning. However, hemolysis is rarely reported despite the common G6PD deficiency following this poisoning, which is due to cardiogenic shock and death before hemolysis [36]. Decreased G6PD also impairs NO production and increases vascular oxidative stress, leading to cardiovascular disease (CVD) progression [35, 37]. However, although hemolysis is a rare symptom of ALP poisoning, it is essential to diagnose it because of its serious consequences.

Effect of ALP on Cardiomyocyte

The heart is the main organ affected by ALP poisoning. Cardiovascular disorders due to ALP poisoning, which include refractory hypotension, dysrhythmia, and congestive heart failure, occur within 12 to 24 h of exposure [38]. In general, cardiac toxicity, cardiac dysfunction, and circulatory collapse that lead to cardiomyocyte death have been identified as the leading causes of death in ALP poisoning [39]. Cardiomyocytes make 75% of cardiac tissue and play a vital role in cardiac circulation. Mitochondria, as critical organelles, are abundant in cardiomyocytes, and by producing ATP through the process of oxidative phosphorylation, contribute to the contractile function of cardiomyocytes and provide 90% of the energy of these cells [15].
One of the most important and prominent features of ALP poisoning is impaired hemostasis of cardiac energy [19], with ALP directly affecting cardiac myocytes. ALP also disrupts the electron transport chain, which in turn disrupts cell energy demand, inhibits cytochrome c oxidase activity as one of the enzymes in the electron transport chain (ETC), reduces ATP levels, and ultimately reduces myocardial energy. In addition to reducing energy, the production of free radicals, especially ROS, and oxidative stress, which lead to LPO, contributes to ALP-induced cardiac toxicity [15, 40]. In general, the heart is very sensitive to oxidative damage due to its high O2 intake, limited antioxidant system, and high metabolic activity [41]. In addition, inhibiting antioxidant enzymes and producing superoxide radicals, reduce NO bioavailability, which increases the adhesion of neutrophils to coronary arteries and leads to vasoconstriction [42]. Excess radical superoxide reacts with NO, leading to increased LPO, which in turn causes cell damage and activation of the apoptotic process in cardiomyocytes (Fig. 1) [43]. ALP-induced myocardial damage is also associated with changes in biochemical biomarkers such as creatine phosphokinase (CPK), creatine kinase myocardial band (CK-MB), and Troponin-T in some cases [44]. However, some reports suggest a change in these biomarkers following ALP poisoning, but according to studies performed by Soltaninejad and her colleagues, these markers are unreliable, and despite ECG changes in acute poisoning, there are conflicting reports of normal and abnormal levels of CPK-MB [45]. In this regard, Karami and Mohajeri state that the average level of these enzymes could not rule out cardiac toxicity, while its high level can confirm myocardial damage [46]. Therefore, it can be said that in the future, with further studies determining the role of CPK and CK-MB markers as diagnostic markers, it is possible to predict ALP damage to cardiomyocytes and to prevent disease progression using appropriate treatment strategies.

Therapeutic Approaches and Their Challenges

Due to the increasing prevalence of ALP poisoning in recent years, many studies have been conducted on this topic, especially to find effective treatment (Table 1). Given that ALP intoxication has no particular counteractant, the foundation of treatment is supportive care. To manage ALP poisoning, gastric lavage with potassium permanganate or a mix with coconut oil and sodium bicarbonate, and injection of charcoal, along with alleviative treatment are used. Besides, acidosis can be treated with early intravenous injection of sodium bicarbonate, cardiogenic shock with liquid vasopressor, and refractory cardiogenic shock with intra-aortic balloon pump or digoxin [47, 48]. However, some disagree with the utilization of sodium bicarbonate for metabolic acidosis treatment in ALP intoxication. They justify their claim by experiments in which no beneficial hemodynamic effects of sodium bicarbonate was found in patients with lactic acidosis. Alternatives for therapy modalities mentioned are intravenous methylene blue for methemoglobinemia, N-acetylcysteine, digoxin, hyperbaric oxygen, trimetazidine, and boric acid [13]. The administration of exogenous drugs has not been a compelling factor in ALP poisoning. The main reason for this is the low efficiency of these drugs, their symptomatic interventions, and the absence of any change in the pathogenic mechanisms following the use of these drugs [12]. Of course, this is not true with drugs that have antioxidant properties. However, the lack of specific antidote and effective supportive care results in high mortality of this type of poisoning.
Table 1
Summary of some studies for the treatment of ALP poison patients
Conducted by
Type of study
Number of patients
Type of sample
Clinical finding
Refs.
Taghaddosinejad et al
Case–control
63
Human
N-Acetyl-cysteine (NAC) cause inhibition cardiotoxicity induced by ALP poison
[78]
Mehrpour et al
Case report
1
Human
Use of intra-aortic balloon pump cause increased survival of patient
[39]
Mohan et al
Case study
7
Human
Use of Extracorporeal membrane oxygenation cause restore cardiomyocyte function
[79]
Marashi et al
Review study
Human
Use of hydroxyethyl starch can restore cardiac function by reducing hypotension
[80]
Pannu et al
Pilot study
60
Human
Improve survival and hemodynamics in patients
[81]
Maleki et al
Prospective study
Animal
Use of selegiline cause improve cardiac and gastrointestinal injury
[82]
Melatonin is a powerful antioxidant that can easily cross all cellular parts with the highest concentration in mitochondria and protect cells against oxidative stress [49]. Studies show that melatonin affects mitochondrial energy homeostasis and can neutralize the reduced activity of complexes I and IV in toxicity and increase ATP production by increasing activity and expression of ETC complexes [50, 51]. Melatonin and its metabolites (C3OHM and AMK) have also been known as a strong lipid peroxyl radical (LOO) scavenger [52, 53]. Melatonin stimulates the expression of genes and the activity of antioxidant enzymes such as glutathione peroxidase (GPx), superoxide dismutase (SOD), glutathione reductase (GR) and catalase (CAT), reducing LPO and ROS levels [38]. Besides, melatonin inhibits selective inhibition of iNOS/i-mt NO, which prevents high NO production in cardiac mitochondria [54]. On the other hand, activation of melatonin receptors reduces cAMP levels and induces PI4,5 biphosphate, which leads to vasocontraction. In general, it regulates blood pressure through an adjustment mechanism [55]. Melatonin also has antiapoptotic effects and prevents apoptosis by preventing the release of cytochrome C and direct inhibition of mitochondrial transition pore opening (MPTP) opening [56].
Another useful substance that has antioxidant properties is Triiodothyramin (T3) [57]. The use of thyroid hormones in the control and treatment of ALP poisoning is mainly due to their antioxidant effect, which plays a significant role in their cytoprotective effects. This antioxidant effect is due to the increased activity of uncoupling proteins in mitochondria, which is accompanied by decreased ROS production with no reduction in ATP synthesis and an increase in activity and expression of mitochondrial ATP-sensitive potassium (mitoK ATP). Studies also show the effect of T3 on improving and increasing serum levels of Total Thiol Molecule (TTM), which plays an essential role against ROS [40, 57, 58].
Minocycline also has a protective effect on the heart due to its radical scavenger activity as well as its crucial role in apoptosis and inflammation pathways [59, 60]. Minocycline improves mitochondrial function by acting on MPTP and inhibiting mitochondrial diffusion in the apoptosis and inflammation processes. These events lead to an increase in ST height, which indicates myocardial and pericardial damage that is significantly associated with mortality following poisoning. The prolonged QTC interval is observed following ALP poisoning that indicates left ventricular dysfunction, arrhythmia, and cardiac cell death. Also, the proliferation of QRS complex occurs as a change in the potential of right and left ventricular depolarization [61, 62]. Haghi-Aminjan and colleagues reported that minocycline had a significant effect on improving these cardiac parameters. This effect is due to its accumulation in cardiomyocytes several times more than the level of plasma and its ability to be chelated into bivalent cations such as Ca2+ cardiomyocytes [61].
In addition to oxidative changes, acidosis and hypoxia also occur after ALP poisoning, which reduces pH and intracellular energy, affects contraction, and causes arrhythmias. Also, following hypoxia, there will be a decrease in cellular Mg2+ but an increase in mitochondrial Ca2+ consumption, leading to inhibition of ATP synthesis [63, 64]. Recent studies have shown that 25Mg PMC16 increases ATP in the heart by activating the transfer of Mg2+ into cardiac cells. It is interesting to know note that sodium bicarbonate (NaHCO3) as standard treatment was not able to improve ALP poisoning in terms of QRS, HR, and BP changes. However, concomitant treatment with 25mg PMC16 was very successful [65]. In general, the effects of Mg2+ are attributed to three main mechanisms, namely rapid distribution to the myocardium, strong antioxidant potential, and the ability to increase intracellular energy [66, 67].
l-carnitine is an essential cofactor in the metabolism of fatty acids, which is the primary source of energy for the cardiac muscle. Decreased oxidation of fatty acids after myocardial ischemia leads to the accumulation of long-chain Acyl-CoA esters, which leads to a decrease in the energy of cardiomyocytes and affects muscle contraction and electrical conduction of the heart [68]. ALP poisoning leads to myocardial necrosis similar to the one that occurs during ischemia and is associated with the accumulation of long-chain Acyl-CoA metabolites. According to previous studies, reducing the effects of these metabolites on cardiac metabolism leads to good results [69].
However, the primary therapeutic challenge for ALP poisoning is the effective management of resistant blood pressure and shock due to reduced cardiac contraction which has a high priority. Vasopressors and phosphodiesterase inhibitors can support this hemodynamic condition, but they are not harmless [70, 71]. Jafari et al. found that vasopressin had beneficial effects on severe hypotension and shock. Milrinone, as a phosphodiesterase III inhibitor, was able to improve acute heart failure by increasing intracellular cAMP in myocytes and strong inotropic effects [72]. This inotropic compound increased the contractile force of the myocardium by reducing the O2 cardiac demand of catecholamines [73]. Also, evidence indicates that it causes corrective effects on ST, QRS complex, and QTC changes as electrical conductivity indicators that occur after ALP poisoning [72]. Adding digoxin and glucagon can be helpful for poisoned patients who still have refractory shock despite the use of conventional vasopressors and inotropic drugs. Digoxin is a Na–K ATPase inhibitor that increases myocardial contraction and inotropy without significantly affecting peripheral arteries [74, 75]. Early detection and intervention at the onset of cardiac damage appear to be effective in delaying their progression. In recent years, many studies have shown that changes in red cell distribution width (RDW) levels as a predictor marker are associated with prognosis and assessment of the severity and progression of CVD [76]. Increased causative stress in RBCs worsens their mechanical properties and leads to tissue perfusion disorders, which are reflected in increased RDW levels [77]. Therefore, RDW may be considered a predictor of cardiovascular complications from ALP poisoning.

Conclusion and Future Perspective

Nowadays, poisoning with metal phosphides, especially aluminum phosphide (ALP), is one of the main health threats in human societies. So far, many studies have been performed on ALP pathogenesis in patients, and the results of most studies emphasize that ALP causes mitochondrial dysfunction. Mitochondrial dysfunction leads to decreased ATP production and increased intravascular hemolysis of RBCs. These disorders result in decreased myocytes function and CVD. Since cardiac disorders, especially cardiac shock, are the leading causes of death in ALP patients, it can be argued that identifying markers associated with cardiac diseases in the early stages of poisoning can lead to the best choices of treatment strategies.

Acknowledgements

We wish to thank all our colleagues in legal medicine organization, Ahvaz, Iran.

Compliance with Ethical Standards

Conflict of interest

Authors declare that they have no conflict of interest.

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Unsere Produktempfehlungen

e.Med Interdisziplinär

Kombi-Abonnement

Jetzt e.Med zum Sonderpreis bestellen!

Für Ihren Erfolg in Klinik und Praxis - Die beste Hilfe in Ihrem Arbeitsalltag

Mit e.Med Interdisziplinär erhalten Sie Zugang zu allen CME-Fortbildungen und Fachzeitschriften auf SpringerMedizin.de.

Jetzt bestellen und 100 € sparen!

e.Med Innere Medizin

Kombi-Abonnement

Mit e.Med Innere Medizin erhalten Sie Zugang zu CME-Fortbildungen des Fachgebietes Innere Medizin, den Premium-Inhalten der internistischen Fachzeitschriften, inklusive einer gedruckten internistischen Zeitschrift Ihrer Wahl.

Jetzt bestellen und 100 € sparen!

Literatur
1.
Zurück zum Zitat Hashemi-Domeneh, B., Zamani, N., Hassanian-Moghaddam, H., Rahimi, M., Shadnia, S., Erfantalab, P., et al. (2016). A review of aluminium phosphide poisoning and a flowchart to treat it. Archives of Industrial Hygiene and Toxicology, 67(3), 183–193.PubMed Hashemi-Domeneh, B., Zamani, N., Hassanian-Moghaddam, H., Rahimi, M., Shadnia, S., Erfantalab, P., et al. (2016). A review of aluminium phosphide poisoning and a flowchart to treat it. Archives of Industrial Hygiene and Toxicology, 67(3), 183–193.PubMed
2.
Zurück zum Zitat Gunnell, D., & Eddleston, M. (2003). Suicide by intentional ingestion of pesticides: A continuing tragedy in developing countries. International Journal of Epidemiology., 32, 903–909. Gunnell, D., & Eddleston, M. (2003). Suicide by intentional ingestion of pesticides: A continuing tragedy in developing countries. International Journal of Epidemiology., 32, 903–909.
3.
Zurück zum Zitat Hassanian-Moghaddam, H. (2014). Phosphides and phosphine: Mechanisms for toxicity and range of the problem. Clinical Toxicology, 52(4), 397–398. Hassanian-Moghaddam, H. (2014). Phosphides and phosphine: Mechanisms for toxicity and range of the problem. Clinical Toxicology, 52(4), 397–398.
4.
Zurück zum Zitat Sciuto, A. M., Wong, B. J., Martens, M. E., Hoard-Fruchey, H., & Perkins, M. W. (2016). Phosphine toxicity: A story of disrupted mitochondrial metabolism. Annals of the New York Academy of Sciences, 1374(1), 41.PubMedPubMedCentral Sciuto, A. M., Wong, B. J., Martens, M. E., Hoard-Fruchey, H., & Perkins, M. W. (2016). Phosphine toxicity: A story of disrupted mitochondrial metabolism. Annals of the New York Academy of Sciences, 1374(1), 41.PubMedPubMedCentral
5.
Zurück zum Zitat Mathai, A., & Bhanu, M. S. (2010). Acute aluminium phosphide poisoning: Can we predict mortality? Indian Journal of Anaesthesia, 54(4), 302.PubMedPubMedCentral Mathai, A., & Bhanu, M. S. (2010). Acute aluminium phosphide poisoning: Can we predict mortality? Indian Journal of Anaesthesia, 54(4), 302.PubMedPubMedCentral
6.
Zurück zum Zitat Mehrpour, O., Natalie, N., & Patrick, N. G. (2019). "Is cytochrome oxidase inhibition the primary mechanism in aluminum phosphide poisoning? Environmental Science and Pollution Research International., 129, 613–614. Mehrpour, O., Natalie, N., & Patrick, N. G. (2019). "Is cytochrome oxidase inhibition the primary mechanism in aluminum phosphide poisoning? Environmental Science and Pollution Research International., 129, 613–614.
7.
Zurück zum Zitat Kariman, H., Heydari, K., Fakhri, M., Shahrami, A., Dolatabadi, A. A., Mohammadi, H. A., et al. (2018). Aluminium phosphide poisoning and oxidative stress. Journal of Medical Toxicology, 8(3), 281–284. Kariman, H., Heydari, K., Fakhri, M., Shahrami, A., Dolatabadi, A. A., Mohammadi, H. A., et al. (2018). Aluminium phosphide poisoning and oxidative stress. Journal of Medical Toxicology, 8(3), 281–284.
8.
Zurück zum Zitat Hena, Z., McCabe, M. E., Perez, M. M., Sharma, M., Sutton, N. J., Peek, G. J., et al. (2018). Aluminum phosphide poisoning: Successful recovery of multiorgan failure in a pediatric patient. International Journal of Pediatrics and Adolescent Medicine, 5(4), 155–158.PubMed Hena, Z., McCabe, M. E., Perez, M. M., Sharma, M., Sutton, N. J., Peek, G. J., et al. (2018). Aluminum phosphide poisoning: Successful recovery of multiorgan failure in a pediatric patient. International Journal of Pediatrics and Adolescent Medicine, 5(4), 155–158.PubMed
9.
Zurück zum Zitat Anand, R., Sharma, D., Verma, D., Bhalla, A., Gill, K., & Singh, S. (2013). Mitochondrial electron transport chain complexes, catalase and markers of oxidative stress in platelets of patients with severe aluminum phosphide poisoning. Human & Experimental Toxicology, 32(8), 807–816. Anand, R., Sharma, D., Verma, D., Bhalla, A., Gill, K., & Singh, S. (2013). Mitochondrial electron transport chain complexes, catalase and markers of oxidative stress in platelets of patients with severe aluminum phosphide poisoning. Human & Experimental Toxicology, 32(8), 807–816.
10.
Zurück zum Zitat Singh, S., Bhalla, A., Verma, S. K., Kaur, A., & Gill, K. (2006). Cytochrome-c oxidase inhibition in 26 aluminum phosphide poisoned patients. Clinical Toxicology, 44(2), 155–158.PubMed Singh, S., Bhalla, A., Verma, S. K., Kaur, A., & Gill, K. (2006). Cytochrome-c oxidase inhibition in 26 aluminum phosphide poisoned patients. Clinical Toxicology, 44(2), 155–158.PubMed
11.
Zurück zum Zitat Proudfoot, A. T. (2009). Aluminium and zinc phosphide poisoning. Clinical toxicology, 47(2), 89–100.PubMed Proudfoot, A. T. (2009). Aluminium and zinc phosphide poisoning. Clinical toxicology, 47(2), 89–100.PubMed
12.
Zurück zum Zitat Asghari, M. H., Abdollahi, M., de Oliveira, M. R., & Nabavi, S. M. (2017). A review of the protective role of melatonin during phosphine-induced cardiotoxicity: Focus on mitochondrial dysfunction, oxidative stress and apoptosis. Journal of Pharmacy and Pharmacology, 69(3), 236–243.PubMed Asghari, M. H., Abdollahi, M., de Oliveira, M. R., & Nabavi, S. M. (2017). A review of the protective role of melatonin during phosphine-induced cardiotoxicity: Focus on mitochondrial dysfunction, oxidative stress and apoptosis. Journal of Pharmacy and Pharmacology, 69(3), 236–243.PubMed
13.
Zurück zum Zitat Baruah, U., Sahni, A., & Sachdeva, H. C. (2015). Successful management of aluminium phosphide poisoning using intravenous lipid emulsion: Report of two cases. Indian Journal of Critical Care Medicine, 19(12), 735.PubMedPubMedCentral Baruah, U., Sahni, A., & Sachdeva, H. C. (2015). Successful management of aluminium phosphide poisoning using intravenous lipid emulsion: Report of two cases. Indian Journal of Critical Care Medicine, 19(12), 735.PubMedPubMedCentral
14.
Zurück zum Zitat Yan, H., Xiang, P., Zhang, S., Shen, B., & Shen, M. (2017). Diagnosis of aluminum phosphide poisoning using a new analytical approach: Forensic application to a lethal intoxication. International Journal of Legal Medicine, 131(4), 1001–1007.PubMed Yan, H., Xiang, P., Zhang, S., Shen, B., & Shen, M. (2017). Diagnosis of aluminum phosphide poisoning using a new analytical approach: Forensic application to a lethal intoxication. International Journal of Legal Medicine, 131(4), 1001–1007.PubMed
15.
Zurück zum Zitat Baghaei, A., Solgi, R., Jafari, A., Abdolghaffari, A. H., Golaghaei, A., & Asghari, M. H. (2016). Molecular and biochemical evidence on the protection of cardiomyocytes from phosphine-induced oxidative stress, mitochondrial dysfunction and apoptosis by acetyl-l-carnitine. Environmental Toxicology and Pharmacology, 42, 30–37.PubMed Baghaei, A., Solgi, R., Jafari, A., Abdolghaffari, A. H., Golaghaei, A., & Asghari, M. H. (2016). Molecular and biochemical evidence on the protection of cardiomyocytes from phosphine-induced oxidative stress, mitochondrial dysfunction and apoptosis by acetyl-l-carnitine. Environmental Toxicology and Pharmacology, 42, 30–37.PubMed
16.
Zurück zum Zitat Ježek, P., Plecitá-Hlavatá, L., Smolková, K., & Rossignol, R. (2010). Distinctions and similarities of cell bioenergetics and the role of mitochondria in hypoxia, cancer, and embryonic development. The International Journal of Biochemistry & Cell Biology, 42(5), 604–622. Ježek, P., Plecitá-Hlavatá, L., Smolková, K., & Rossignol, R. (2010). Distinctions and similarities of cell bioenergetics and the role of mitochondria in hypoxia, cancer, and embryonic development. The International Journal of Biochemistry & Cell Biology, 42(5), 604–622.
17.
Zurück zum Zitat Anand, R., Kumari, P., Kaushal, A., Bal, A., Wani, W. Y., Sunkaria, A., et al. (2012). Effect of acute aluminum phosphide exposure on rats: A biochemical and histological correlation. Toxicology Letters, 215(1), 62–69.PubMed Anand, R., Kumari, P., Kaushal, A., Bal, A., Wani, W. Y., Sunkaria, A., et al. (2012). Effect of acute aluminum phosphide exposure on rats: A biochemical and histological correlation. Toxicology Letters, 215(1), 62–69.PubMed
18.
Zurück zum Zitat Valmas, N., Zuryn, S., & Ebert, P. R. (2008). Mitochondrial uncouplers act synergistically with the fumigant phosphine to disrupt mitochondrial membrane potential and cause cell death. Toxicology, 252(1–3), 33–39.PubMed Valmas, N., Zuryn, S., & Ebert, P. R. (2008). Mitochondrial uncouplers act synergistically with the fumigant phosphine to disrupt mitochondrial membrane potential and cause cell death. Toxicology, 252(1–3), 33–39.PubMed
19.
Zurück zum Zitat Dua, R., & Gill, K. D. (2004). Effect of aluminium phosphide exposure on kinetic properties of cytochrome oxidase and mitochondrial energy metabolism in rat brain. Biochimica et Biophysica Acta (BBA)., 1674(1), 4–11.PubMed Dua, R., & Gill, K. D. (2004). Effect of aluminium phosphide exposure on kinetic properties of cytochrome oxidase and mitochondrial energy metabolism in rat brain. Biochimica et Biophysica Acta (BBA)., 1674(1), 4–11.PubMed
20.
Zurück zum Zitat Nath, N. S., Bhattacharya, I., Tuck, A. G., Schlipalius, D. I., & Ebert, P. R. (2011). Mechanisms of phosphine toxicity. Journal of Toxicology., 179, 1–8. Nath, N. S., Bhattacharya, I., Tuck, A. G., Schlipalius, D. I., & Ebert, P. R. (2011). Mechanisms of phosphine toxicity. Journal of Toxicology., 179, 1–8.
21.
Zurück zum Zitat Tang, P.-A., Duan, J.-Y., Wu, H.-J., Ju, X.-R., & Yuan, M.-L. (2017). Reference gene selection to determine differences in mitochondrial gene expressions in phosphine-susceptible and phosphine-resistant strains of Cryptolestes ferrugineus, using qRT-PCR. Scientific Reports, 7(1), 1–11. Tang, P.-A., Duan, J.-Y., Wu, H.-J., Ju, X.-R., & Yuan, M.-L. (2017). Reference gene selection to determine differences in mitochondrial gene expressions in phosphine-susceptible and phosphine-resistant strains of Cryptolestes ferrugineus, using qRT-PCR. Scientific Reports, 7(1), 1–11.
22.
Zurück zum Zitat Lee, S. S., Lee, R. Y., Fraser, A. G., Kamath, R. S., Ahringer, J., & Ruvkun, G. (2003). A systematic RNAi screen identifies a critical role for mitochondria in C. elegans longevity. Nature Genetics, 33(1), 40–48.PubMed Lee, S. S., Lee, R. Y., Fraser, A. G., Kamath, R. S., Ahringer, J., & Ruvkun, G. (2003). A systematic RNAi screen identifies a critical role for mitochondria in C. elegans longevity. Nature Genetics, 33(1), 40–48.PubMed
23.
Zurück zum Zitat Zuryn, S., Kuang, J., & Ebert, P. (2008). Mitochondrial modulation of phosphine toxicity and resistance in Caenorhabditis elegans. Toxicological Sciences, 102(1), 179–186.PubMed Zuryn, S., Kuang, J., & Ebert, P. (2008). Mitochondrial modulation of phosphine toxicity and resistance in Caenorhabditis elegans. Toxicological Sciences, 102(1), 179–186.PubMed
24.
Zurück zum Zitat Jagadeesan, R., Collins, P. J., Daglish, G. J., Ebert, P. R., & Schlipalius, D. I. (2012). Phosphine resistance in the rust red flour beetle, Tribolium castaneum (Coleoptera: Tenebrionidae): inheritance, gene interactions and fitness costs. PLoS ONE, 7(2), e31582.PubMedPubMedCentral Jagadeesan, R., Collins, P. J., Daglish, G. J., Ebert, P. R., & Schlipalius, D. I. (2012). Phosphine resistance in the rust red flour beetle, Tribolium castaneum (Coleoptera: Tenebrionidae): inheritance, gene interactions and fitness costs. PLoS ONE, 7(2), e31582.PubMedPubMedCentral
25.
Zurück zum Zitat Daglish, G. J., Nayak, M. K., Pavic, H., & Smith, L. W. (2015). Prevalence and potential fitness cost of weak phosphine resistance in Tribolium castaneum (Herbst) in eastern Australia. Journal of stored products research, 61, 54–58. Daglish, G. J., Nayak, M. K., Pavic, H., & Smith, L. W. (2015). Prevalence and potential fitness cost of weak phosphine resistance in Tribolium castaneum (Herbst) in eastern Australia. Journal of stored products research, 61, 54–58.
26.
Zurück zum Zitat Aminjan, H. H., Abtahi, S. R., Hazrati, E., Chamanara, M., Jalili, M., & Paknejad, B. (2019). Targeting of oxidative stress and inflammation through ROS/NF-kappaB pathway in phosphine-induced hepatotoxicity mitigation. Life Sciences, 232, 116607. Aminjan, H. H., Abtahi, S. R., Hazrati, E., Chamanara, M., Jalili, M., & Paknejad, B. (2019). Targeting of oxidative stress and inflammation through ROS/NF-kappaB pathway in phosphine-induced hepatotoxicity mitigation. Life Sciences, 232, 116607.
27.
Zurück zum Zitat Circu, M. L., & Aw, T. Y. (2012). Glutathione and modulation of cell apoptosis. Biochimica et Biophysica Acta (BBA)., 1823(10), 1767–1777.PubMed Circu, M. L., & Aw, T. Y. (2012). Glutathione and modulation of cell apoptosis. Biochimica et Biophysica Acta (BBA)., 1823(10), 1767–1777.PubMed
28.
Zurück zum Zitat Soltaninejad, K., Nelson, L. S., Khodakarim, N., Dadvar, Z., & Shadnia, S. (2011). Unusual complication of aluminum phosphide poisoning: Development of hemolysis and methemoglobinemia and its successful treatment. Indian Journal of Critical Care Medicine, 15(2), 117.PubMedPubMedCentral Soltaninejad, K., Nelson, L. S., Khodakarim, N., Dadvar, Z., & Shadnia, S. (2011). Unusual complication of aluminum phosphide poisoning: Development of hemolysis and methemoglobinemia and its successful treatment. Indian Journal of Critical Care Medicine, 15(2), 117.PubMedPubMedCentral
29.
Zurück zum Zitat Mostafazadeh, B., Pajoumand, A., Farzaneh, E., Aghabiklooei, A., & Rasouli, M. R. (2011). Blood levels of methemoglobin in patients with aluminum phosphide poisoning and its correlation with patient’s outcome. Journal of Medical Toxicology, 7(1), 40–43.PubMed Mostafazadeh, B., Pajoumand, A., Farzaneh, E., Aghabiklooei, A., & Rasouli, M. R. (2011). Blood levels of methemoglobin in patients with aluminum phosphide poisoning and its correlation with patient’s outcome. Journal of Medical Toxicology, 7(1), 40–43.PubMed
30.
Zurück zum Zitat Vosooghi, A. A., & Salmasi, M. (2018). G6PD deficiency and aluminum phosphide poisoning. Journal of Research in Medical Sciences, 23(1), 83.PubMedPubMedCentral Vosooghi, A. A., & Salmasi, M. (2018). G6PD deficiency and aluminum phosphide poisoning. Journal of Research in Medical Sciences, 23(1), 83.PubMedPubMedCentral
31.
Zurück zum Zitat Lakshmi, B. (2002). Methemoglobinemia with aluminum phosphide poisoning. The American journal of Emergency Medicine, 20(2), 130–132.PubMed Lakshmi, B. (2002). Methemoglobinemia with aluminum phosphide poisoning. The American journal of Emergency Medicine, 20(2), 130–132.PubMed
32.
Zurück zum Zitat Chui, J., Poon, W., Chan, K., Chan, A., & Buckley, T. (2005). Nitrite-induced methaemoglobinaemia–aetiology, diagnosis and treatment. Anaesthesia, 60(5), 496–500.PubMed Chui, J., Poon, W., Chan, K., Chan, A., & Buckley, T. (2005). Nitrite-induced methaemoglobinaemia–aetiology, diagnosis and treatment. Anaesthesia, 60(5), 496–500.PubMed
33.
Zurück zum Zitat Afolabi, O. K., Oyewo, E. B., Adeleke, G. E., Badmus, J. A., & Wusu, A. D. (2019). Mitigation of aluminium phosphide-induced hematotoxicity and ovarian oxidative damage in Wistar Rats by Hesperidin. American Journal of Biochemistry, 9(1), 7–16. Afolabi, O. K., Oyewo, E. B., Adeleke, G. E., Badmus, J. A., & Wusu, A. D. (2019). Mitigation of aluminium phosphide-induced hematotoxicity and ovarian oxidative damage in Wistar Rats by Hesperidin. American Journal of Biochemistry, 9(1), 7–16.
34.
Zurück zum Zitat Malakar, S., Negi, B. D., Dutt, K., & Raina, S. (2019). Intravascular hemolysis in aluminum phosphide poisoning. Indian Journal of Critical Care Medicine, 23(2), 106.PubMedPubMedCentral Malakar, S., Negi, B. D., Dutt, K., & Raina, S. (2019). Intravascular hemolysis in aluminum phosphide poisoning. Indian Journal of Critical Care Medicine, 23(2), 106.PubMedPubMedCentral
35.
Zurück zum Zitat Cappellini, M. D., & Fiorelli, G. (2008). Glucose-6-phosphate dehydrogenase deficiency. The Lancet, 371(9606), 64–74. Cappellini, M. D., & Fiorelli, G. (2008). Glucose-6-phosphate dehydrogenase deficiency. The Lancet, 371(9606), 64–74.
36.
Zurück zum Zitat Srinivas, R., Agarwal, R., Jairam, A., & Sakhuja, V. (2007). Intravascular haemolysis due to glucose-6-phosphate dehydrogenase deficiency in a patient with aluminium phosphide poisoning. Emergency Medicine Journal, 24(1), 67–68.PubMed Srinivas, R., Agarwal, R., Jairam, A., & Sakhuja, V. (2007). Intravascular haemolysis due to glucose-6-phosphate dehydrogenase deficiency in a patient with aluminium phosphide poisoning. Emergency Medicine Journal, 24(1), 67–68.PubMed
37.
Zurück zum Zitat Kuhn, V., Diederich, L., Keller, T. S., Kramer, C. M., Lückstädt, W., Panknin, C., et al. (2017). Red blood cell function and dysfunction: redox regulation, nitric oxide metabolism, anemia. Antioxidants & Redox Signaling, 26(13), 718–742. Kuhn, V., Diederich, L., Keller, T. S., Kramer, C. M., Lückstädt, W., Panknin, C., et al. (2017). Red blood cell function and dysfunction: redox regulation, nitric oxide metabolism, anemia. Antioxidants & Redox Signaling, 26(13), 718–742.
38.
Zurück zum Zitat Asghari, M. H., Moloudizargari, M., Baeeri, M., Baghaei, A., Rahimifard, M., Solgi, R., et al. (2017). On the mechanisms of melatonin in protection of aluminum phosphide cardiotoxicity. Archives of Toxicology, 91(9), 3109–3120.PubMed Asghari, M. H., Moloudizargari, M., Baeeri, M., Baghaei, A., Rahimifard, M., Solgi, R., et al. (2017). On the mechanisms of melatonin in protection of aluminum phosphide cardiotoxicity. Archives of Toxicology, 91(9), 3109–3120.PubMed
39.
Zurück zum Zitat Mehrpour, O., Amouzeshi, A., Dadpour, B., Oghabian, Z., Zamani, N., Amini, S., et al. (2014). Successful treatment of cardiogenic shock with an intraaortic balloon pump following aluminium phosphide poisoning. Archives of Industrial Hygiene and Toxicology, 65(1), 121–127.PubMed Mehrpour, O., Amouzeshi, A., Dadpour, B., Oghabian, Z., Zamani, N., Amini, S., et al. (2014). Successful treatment of cardiogenic shock with an intraaortic balloon pump following aluminium phosphide poisoning. Archives of Industrial Hygiene and Toxicology, 65(1), 121–127.PubMed
40.
Zurück zum Zitat Goharbari, M., Taghaddosinejad, F., Arefi, M., Sharifzadeh, M., Mojtahedzadeh, M., Nikfar, S., et al. (2018). Therapeutic effects of oral liothyronine on aluminum phosphide poisoning as an adjuvant therapy: A clinical trial. Human & Experimental Toxicology, 37(2), 107–117. Goharbari, M., Taghaddosinejad, F., Arefi, M., Sharifzadeh, M., Mojtahedzadeh, M., Nikfar, S., et al. (2018). Therapeutic effects of oral liothyronine on aluminum phosphide poisoning as an adjuvant therapy: A clinical trial. Human & Experimental Toxicology, 37(2), 107–117.
41.
Zurück zum Zitat Kaiserová, H., Šimůnek, T., Štěrba, M., den Hartog, G. J., Schröterová, L., Popelová, O., et al. (2007). New iron chelators in anthracycline-induced cardiotoxicity. Cardiovascular Toxicology, 7(2), 145–150.PubMed Kaiserová, H., Šimůnek, T., Štěrba, M., den Hartog, G. J., Schröterová, L., Popelová, O., et al. (2007). New iron chelators in anthracycline-induced cardiotoxicity. Cardiovascular Toxicology, 7(2), 145–150.PubMed
42.
Zurück zum Zitat Gouda, A. S., El-Nabarawy, N. A., & Ibrahim, S. F. (2018). Moringa oleifera extract (Lam) attenuates Aluminium phosphide-induced acute cardiac toxicity in rats. Toxicology reports, 5, 209–212.PubMedPubMedCentral Gouda, A. S., El-Nabarawy, N. A., & Ibrahim, S. F. (2018). Moringa oleifera extract (Lam) attenuates Aluminium phosphide-induced acute cardiac toxicity in rats. Toxicology reports, 5, 209–212.PubMedPubMedCentral
43.
Zurück zum Zitat Dua, R., Sunkaria, A., Kumar, V., & Gill, K. D. (2010). Impaired mitochondrial energy metabolism and kinetic properties of cytochrome oxidase following acute aluminium phosphide exposure in rat liver. Food and Chemical Toxicology, 48(1), 53–60.PubMed Dua, R., Sunkaria, A., Kumar, V., & Gill, K. D. (2010). Impaired mitochondrial energy metabolism and kinetic properties of cytochrome oxidase following acute aluminium phosphide exposure in rat liver. Food and Chemical Toxicology, 48(1), 53–60.PubMed
44.
Zurück zum Zitat Soltaninejad, K., Beyranvand, M.-R., Momenzadeh, S.-A., & Shadnia, S. (2012). Electrocardiographic findings and cardiac manifestations in acute aluminum phosphide poisoning. Journal of Forensic and Legal Medicine, 19(5), 291–293.PubMed Soltaninejad, K., Beyranvand, M.-R., Momenzadeh, S.-A., & Shadnia, S. (2012). Electrocardiographic findings and cardiac manifestations in acute aluminum phosphide poisoning. Journal of Forensic and Legal Medicine, 19(5), 291–293.PubMed
45.
Zurück zum Zitat Soltaninejad, K., Shadnia, S., Ziyapour, B., & Brent, J. (2009). Aluminum phosphide intoxication mimicking ischemic heart disease led to unjustified treatment with streptokinase Letter to the editor. Clinical Toxicology, 47(9), 908–909.PubMed Soltaninejad, K., Shadnia, S., Ziyapour, B., & Brent, J. (2009). Aluminum phosphide intoxication mimicking ischemic heart disease led to unjustified treatment with streptokinase Letter to the editor. Clinical Toxicology, 47(9), 908–909.PubMed
46.
Zurück zum Zitat Karami-Mohajeri, S., Jafari, A., & Abdollahi, M. (2013). Comprehensive review of the mechanistic approach and related therapies to cardiovascular effects of aluminum phosphide. International Journal of Pharmacology, 9(8), 493–500. Karami-Mohajeri, S., Jafari, A., & Abdollahi, M. (2013). Comprehensive review of the mechanistic approach and related therapies to cardiovascular effects of aluminum phosphide. International Journal of Pharmacology, 9(8), 493–500.
47.
Zurück zum Zitat Ghasemi Tousi, A. (2017). Successful treatment of aluminium phosphide, is it possible? Asia Pacific Journal of Medical Toxicology, 6(3), 90–96. Ghasemi Tousi, A. (2017). Successful treatment of aluminium phosphide, is it possible? Asia Pacific Journal of Medical Toxicology, 6(3), 90–96.
48.
Zurück zum Zitat Karimani, A., Mohammadpour, A. H., Zirak, M. R., Rezaee, R., Megarbane, B., & Tsatsakis, A. (2018). Antidotes for aluminum phosphide poisoning: An update. Toxicology Reports, 5, 1053–1059.PubMedPubMedCentral Karimani, A., Mohammadpour, A. H., Zirak, M. R., Rezaee, R., Megarbane, B., & Tsatsakis, A. (2018). Antidotes for aluminum phosphide poisoning: An update. Toxicology Reports, 5, 1053–1059.PubMedPubMedCentral
49.
Zurück zum Zitat Korkmaz, A., Reiter, R. J., Topal, T., Manchester, L. C., Oter, S., & Tan, D.-X. (2009). Melatonin: An established antioxidant worthy of use in clinical trials. Molecular Medicine, 15(1), 43–50.PubMed Korkmaz, A., Reiter, R. J., Topal, T., Manchester, L. C., Oter, S., & Tan, D.-X. (2009). Melatonin: An established antioxidant worthy of use in clinical trials. Molecular Medicine, 15(1), 43–50.PubMed
50.
Zurück zum Zitat Martin, M., Macias, M., Escames, G., Reiter, R. J., Agapito, M., Ortiz, G., et al. (2000). Melatonin-induced increased activity of the respiratory chain complexes I and IV can prevent mitochondrial damage induced by ruthenium red in vivo. Journal of Pineal Research, 28(4), 242–248.PubMed Martin, M., Macias, M., Escames, G., Reiter, R. J., Agapito, M., Ortiz, G., et al. (2000). Melatonin-induced increased activity of the respiratory chain complexes I and IV can prevent mitochondrial damage induced by ruthenium red in vivo. Journal of Pineal Research, 28(4), 242–248.PubMed
52.
Zurück zum Zitat Marchetti, C., Sidahmed-Adrar, N., Collin, F., Jore, D., Gardès-Albert, M., & Bonnefont-Rousselot, D. (2011). Melatonin protects PLPC liposomes and LDL towards radical-induced oxidation. Journal of Pineal Research, 51(3), 286–296.PubMed Marchetti, C., Sidahmed-Adrar, N., Collin, F., Jore, D., Gardès-Albert, M., & Bonnefont-Rousselot, D. (2011). Melatonin protects PLPC liposomes and LDL towards radical-induced oxidation. Journal of Pineal Research, 51(3), 286–296.PubMed
53.
Zurück zum Zitat Mekhloufi, J., Vitrac, H., Yous, S., Duriez, P., Jore, D., Gardès-Albert, M., et al. (2007). Quantification of the water/lipid affinity of melatonin and a pinoline derivative in lipid models. Journal of Pineal Research, 42(4), 330–337.PubMed Mekhloufi, J., Vitrac, H., Yous, S., Duriez, P., Jore, D., Gardès-Albert, M., et al. (2007). Quantification of the water/lipid affinity of melatonin and a pinoline derivative in lipid models. Journal of Pineal Research, 42(4), 330–337.PubMed
54.
Zurück zum Zitat Ortiz, F., García, J. A., Acuña-Castroviejo, D., Doerrier, C., López, A., Venegas, C., et al. (2014). The beneficial effects of melatonin against heart mitochondrial impairment during sepsis: Inhibition of i NOS and preservation of n NOS. Journal of Pineal Research, 56(1), 71–81.PubMed Ortiz, F., García, J. A., Acuña-Castroviejo, D., Doerrier, C., López, A., Venegas, C., et al. (2014). The beneficial effects of melatonin against heart mitochondrial impairment during sepsis: Inhibition of i NOS and preservation of n NOS. Journal of Pineal Research, 56(1), 71–81.PubMed
55.
Zurück zum Zitat Paulis, L., & Šimko, F. (2007). Blood pressure modulation and cardiovascular protection by melatonin: potential mechanisms behind. Physiological Research, 56(6), 671–684.PubMed Paulis, L., & Šimko, F. (2007). Blood pressure modulation and cardiovascular protection by melatonin: potential mechanisms behind. Physiological Research, 56(6), 671–684.PubMed
56.
Zurück zum Zitat Radogna, F., Albertini, M. C., De Nicola, M., Diederich, M., Bejarano, I., & Ghibelli, L. (2015). Melatonin promotes Bax sequestration to mitochondria reducing cell susceptibility to apoptosis via the lipoxygenase metabolite 5-hydroxyeicosatetraenoic acid. Mitochondrion, 21, 113–121.PubMed Radogna, F., Albertini, M. C., De Nicola, M., Diederich, M., Bejarano, I., & Ghibelli, L. (2015). Melatonin promotes Bax sequestration to mitochondria reducing cell susceptibility to apoptosis via the lipoxygenase metabolite 5-hydroxyeicosatetraenoic acid. Mitochondrion, 21, 113–121.PubMed
57.
Zurück zum Zitat Abdolghaffari, A. H., Baghaei, A., Solgi, R., Gooshe, M., Baeeri, M., Navaei-Nigjeh, M., et al. (2015). Molecular and biochemical evidences on the protective effects of triiodothyronine against phosphine-induced cardiac and mitochondrial toxicity. Life Sciences, 139, 30–39.PubMed Abdolghaffari, A. H., Baghaei, A., Solgi, R., Gooshe, M., Baeeri, M., Navaei-Nigjeh, M., et al. (2015). Molecular and biochemical evidences on the protective effects of triiodothyronine against phosphine-induced cardiac and mitochondrial toxicity. Life Sciences, 139, 30–39.PubMed
58.
Zurück zum Zitat Cioffi, F., Senese, R., Lanni, A., & Goglia, F. (2013). Thyroid hormones and mitochondria: With a brief look at derivatives and analogues. Molecular and Cellular Endocrinology, 379(1–2), 51–61.PubMed Cioffi, F., Senese, R., Lanni, A., & Goglia, F. (2013). Thyroid hormones and mitochondria: With a brief look at derivatives and analogues. Molecular and Cellular Endocrinology, 379(1–2), 51–61.PubMed
59.
Zurück zum Zitat Abbaszadeh, A., Darabi, S., Hasanvand, A., Amini-Khoei, H., Abbasnezhad, A., Choghakhori, R., et al. (2018). Minocycline through attenuation of oxidative stress and inflammatory response reduces the neuropathic pain in a rat model of chronic constriction injury. Iranian Journal of Basic Medical Sciences, 21(2), 138.PubMedPubMedCentral Abbaszadeh, A., Darabi, S., Hasanvand, A., Amini-Khoei, H., Abbasnezhad, A., Choghakhori, R., et al. (2018). Minocycline through attenuation of oxidative stress and inflammatory response reduces the neuropathic pain in a rat model of chronic constriction injury. Iranian Journal of Basic Medical Sciences, 21(2), 138.PubMedPubMedCentral
60.
Zurück zum Zitat Haghi-Aminjan, H., Asghari, M. H., Goharbari, M. H., & Abdollahi, M. (2017). A systematic review on potential mechanisms of minocycline in kidney diseases. Pharmacological Reports, 69(4), 602–609.PubMed Haghi-Aminjan, H., Asghari, M. H., Goharbari, M. H., & Abdollahi, M. (2017). A systematic review on potential mechanisms of minocycline in kidney diseases. Pharmacological Reports, 69(4), 602–609.PubMed
61.
Zurück zum Zitat Haghi-Aminjan, H., Baeeri, M., Rahimifard, M., Alizadeh, A., Hodjat, M., Hassani, S., et al. (2018). The role of minocycline in alleviating aluminum phosphide-induced cardiac hemodynamic and renal toxicity. Environmental Toxicology and Pharmacology, 64, 26–40.PubMed Haghi-Aminjan, H., Baeeri, M., Rahimifard, M., Alizadeh, A., Hodjat, M., Hassani, S., et al. (2018). The role of minocycline in alleviating aluminum phosphide-induced cardiac hemodynamic and renal toxicity. Environmental Toxicology and Pharmacology, 64, 26–40.PubMed
62.
Zurück zum Zitat Zhu, S., Stavrovskaya, I. G., Drozda, M., Kim, B. Y., Ona, V., Li, M., et al. (2001). Minocycline inhibits cytochrome c release and delays progression of amyotrophic lateral sclerosis in mice. Nature, 417, 74–78. Zhu, S., Stavrovskaya, I. G., Drozda, M., Kim, B. Y., Ona, V., Li, M., et al. (2001). Minocycline inhibits cytochrome c release and delays progression of amyotrophic lateral sclerosis in mice. Nature, 417, 74–78.
63.
Zurück zum Zitat Shadnia, S., Rahimi, M., Pajoumand, A., Rasouli, M.-H., & Abdollahi, M. (2005). Successful treatment of acute aluminium phosphide poisoning: Possible benefit of coconut oil. Human & Experimental Toxicology, 24(4), 215–218. Shadnia, S., Rahimi, M., Pajoumand, A., Rasouli, M.-H., & Abdollahi, M. (2005). Successful treatment of acute aluminium phosphide poisoning: Possible benefit of coconut oil. Human & Experimental Toxicology, 24(4), 215–218.
64.
Zurück zum Zitat Shadnia, S., Sasanian, G., Allami, P., Hosseini, A., Ranjbar, A., Amini-Shirazi, N., et al. (2009). A retrospective 7-years study of aluminum phosphide poisoning in Tehran: Opportunities for prevention. Human & Experimental Toxicology, 28(4), 209–213. Shadnia, S., Sasanian, G., Allami, P., Hosseini, A., Ranjbar, A., Amini-Shirazi, N., et al. (2009). A retrospective 7-years study of aluminum phosphide poisoning in Tehran: Opportunities for prevention. Human & Experimental Toxicology, 28(4), 209–213.
65.
Zurück zum Zitat Baeeri, M., Shariatpanahi, M., Baghaei, A., Ghasemi-Niri, S. F., Mohammadi, H., Mohammadirad, A., et al. (2013). On the benefit of magnetic magnesium nanocarrier in cardiovascular toxicity of aluminum phosphide. Toxicology and Industrial Health, 29(2), 126–135.PubMed Baeeri, M., Shariatpanahi, M., Baghaei, A., Ghasemi-Niri, S. F., Mohammadi, H., Mohammadirad, A., et al. (2013). On the benefit of magnetic magnesium nanocarrier in cardiovascular toxicity of aluminum phosphide. Toxicology and Industrial Health, 29(2), 126–135.PubMed
66.
Zurück zum Zitat Mohammadi, H., Karimi, G., Shafiee, H., Nikfar, S., Baeeri, M., Sabzevari, O., et al. (2011). Benefit of nanocarrier of magnetic magnesium in rat malathion-induced toxicity and cardiac failure using non-invasive monitoring of electrocardiogram and blood pressure. Toxicology and Industrial Health, 27(5), 417–429.PubMed Mohammadi, H., Karimi, G., Shafiee, H., Nikfar, S., Baeeri, M., Sabzevari, O., et al. (2011). Benefit of nanocarrier of magnetic magnesium in rat malathion-induced toxicity and cardiac failure using non-invasive monitoring of electrocardiogram and blood pressure. Toxicology and Industrial Health, 27(5), 417–429.PubMed
67.
Zurück zum Zitat Shafiee, H., Mohammadi, H., Rezayat, S. M., Hosseini, A., Baeeri, M., Hassani, S., et al. (2010). Prevention of malathion-induced depletion of cardiac cells mitochondrial energy and free radical damage by a magnetic magnesium-carrying nanoparticle. Toxicology Mechanisms and Methods, 20(9), 538–543.PubMed Shafiee, H., Mohammadi, H., Rezayat, S. M., Hosseini, A., Baeeri, M., Hassani, S., et al. (2010). Prevention of malathion-induced depletion of cardiac cells mitochondrial energy and free radical damage by a magnetic magnesium-carrying nanoparticle. Toxicology Mechanisms and Methods, 20(9), 538–543.PubMed
68.
Zurück zum Zitat Wang, Z.-Y., Liu, Y.-Y., Liu, G.-H., Lu, H.-B., & Mao, C.-Y. (2018). l-Carnitine and heart disease. Life Sciences, 194, 88–97.PubMed Wang, Z.-Y., Liu, Y.-Y., Liu, G.-H., Lu, H.-B., & Mao, C.-Y. (2018). l-Carnitine and heart disease. Life Sciences, 194, 88–97.PubMed
69.
Zurück zum Zitat Elabbassi, W., Chowdhury, M. A., & Fachtartz, A. A. N. (2014). Severe reversible myocardial injury associated with aluminium phosphide toxicity: A case report and review of literature. Journal of the saudi Heart Association, 26(4), 216–221.PubMed Elabbassi, W., Chowdhury, M. A., & Fachtartz, A. A. N. (2014). Severe reversible myocardial injury associated with aluminium phosphide toxicity: A case report and review of literature. Journal of the saudi Heart Association, 26(4), 216–221.PubMed
70.
Zurück zum Zitat Barry, J. D., Durkovich, D. W., & Williams, S. R. (2006). Vasopressin treatment for cyclic antidepressant overdose. The Journal of Emergency Medicine, 31(1), 65–68.PubMed Barry, J. D., Durkovich, D. W., & Williams, S. R. (2006). Vasopressin treatment for cyclic antidepressant overdose. The Journal of Emergency Medicine, 31(1), 65–68.PubMed
71.
Zurück zum Zitat Lescan, M., Scheule, A., Neumann, B., Haller, C., Westendorff, J., Wendel, H. P., et al. (2013). Beneficial effects on cardiac performance and cardioprotective properties of milrinone after cold ischemia. Journal of Cardiovascular Diseases & Diagnosis, 1(124), 2. Lescan, M., Scheule, A., Neumann, B., Haller, C., Westendorff, J., Wendel, H. P., et al. (2013). Beneficial effects on cardiac performance and cardioprotective properties of milrinone after cold ischemia. Journal of Cardiovascular Diseases & Diagnosis, 1(124), 2.
72.
Zurück zum Zitat Jafari, A., Baghaei, A., Solgi, R., Baeeri, M., Chamanara, M., Hassani, S., et al. (2015). An electrocardiographic, molecular and biochemical approach to explore the cardioprotective effect of vasopressin and milrinone against phosphide toxicity in rats. Food and Chemical Toxicology, 80, 182–192.PubMed Jafari, A., Baghaei, A., Solgi, R., Baeeri, M., Chamanara, M., Hassani, S., et al. (2015). An electrocardiographic, molecular and biochemical approach to explore the cardioprotective effect of vasopressin and milrinone against phosphide toxicity in rats. Food and Chemical Toxicology, 80, 182–192.PubMed
73.
Zurück zum Zitat Tosaka, S., Makita, T., Tosaka, R., Maekawa, T., Cho, S., Hara, T., et al. (2007). Cardioprotection induced by olprinone, a phosphodiesterase III inhibitor, involves phosphatidylinositol-3-OH kinase-Akt and a mitochondrial permeability transition pore during early reperfusion. Journal of Anesthesia, 21(2), 176–180.PubMed Tosaka, S., Makita, T., Tosaka, R., Maekawa, T., Cho, S., Hara, T., et al. (2007). Cardioprotection induced by olprinone, a phosphodiesterase III inhibitor, involves phosphatidylinositol-3-OH kinase-Akt and a mitochondrial permeability transition pore during early reperfusion. Journal of Anesthesia, 21(2), 176–180.PubMed
74.
Zurück zum Zitat Oghabian, Z., & Mehrpour, O. (2016). Treatment of aluminium phosphide poisoning with a combination of intravenous glucagon, digoxin and antioxidant agents. Sultan Qaboos University Medical Journal, 16(3), e352.PubMedPubMedCentral Oghabian, Z., & Mehrpour, O. (2016). Treatment of aluminium phosphide poisoning with a combination of intravenous glucagon, digoxin and antioxidant agents. Sultan Qaboos University Medical Journal, 16(3), e352.PubMedPubMedCentral
75.
Zurück zum Zitat Mehrpour, O., Farzaneh, E., & Abdollahi, M. (2011). Successful treatment of aluminum phosphide poisoning with digoxin: A case report and review of literature. International Journal of Pharmacology, 7(7), 761–764. Mehrpour, O., Farzaneh, E., & Abdollahi, M. (2011). Successful treatment of aluminum phosphide poisoning with digoxin: A case report and review of literature. International Journal of Pharmacology, 7(7), 761–764.
77.
Zurück zum Zitat Li, N., Heng, Z., & Qizhu, T. (2017). Red blood cell distribution width: a novel predictive indicator for cardiovascular and cerebrovascular diseases. Disease Markers, 2017, 7089493.PubMedPubMedCentral Li, N., Heng, Z., & Qizhu, T. (2017). Red blood cell distribution width: a novel predictive indicator for cardiovascular and cerebrovascular diseases. Disease Markers, 2017, 7089493.PubMedPubMedCentral
78.
Zurück zum Zitat Taghaddosinejad, F., Farzaneh, E., Ghazanfari-Nasrabad, M., Eizadi-Mood, N., Hajihosseini, M., & Mehrpour, O. (2016). The effect of N-acetyl cysteine (NAC) on aluminum phosphide poisoning inducing cardiovascular toxicity: A case–control study. Springerplus, 5(1), 1948.PubMedPubMedCentral Taghaddosinejad, F., Farzaneh, E., Ghazanfari-Nasrabad, M., Eizadi-Mood, N., Hajihosseini, M., & Mehrpour, O. (2016). The effect of N-acetyl cysteine (NAC) on aluminum phosphide poisoning inducing cardiovascular toxicity: A case–control study. Springerplus, 5(1), 1948.PubMedPubMedCentral
79.
Zurück zum Zitat Mohan, B., Gupta, V., Ralhan, S., Gupta, D., Puri, S., Wander, G. S., et al. (2015). Role of extracorporeal membrane oxygenation in aluminum phosphide poisoning–induced reversible myocardial dysfunction: A novel therapeutic modality. The Journal of Emergency Medicine, 49(5), 651–656.PubMed Mohan, B., Gupta, V., Ralhan, S., Gupta, D., Puri, S., Wander, G. S., et al. (2015). Role of extracorporeal membrane oxygenation in aluminum phosphide poisoning–induced reversible myocardial dysfunction: A novel therapeutic modality. The Journal of Emergency Medicine, 49(5), 651–656.PubMed
80.
Zurück zum Zitat Marashi, S. M., Arefi, M., Behnoush, B., Nasrabad, M. G., & Nasrabadi, Z. N. (2011). Could hydroxyethyl starch be a therapeutic option in management of acute aluminum phosphide toxicity? Medical Hypotheses, 76(4), 596–598.PubMed Marashi, S. M., Arefi, M., Behnoush, B., Nasrabad, M. G., & Nasrabadi, Z. N. (2011). Could hydroxyethyl starch be a therapeutic option in management of acute aluminum phosphide toxicity? Medical Hypotheses, 76(4), 596–598.PubMed
82.
Zurück zum Zitat Maleki, A., Hosseini, M. J., Rahimi, N., Abdollahi, A., Akbarfakhrabadi, A., Javadian, N., et al. (2019). Adjuvant potential of selegiline in treating acute toxicity of aluminium phosphide in rats. Basic & Clinical Pharmacology & Toxicology, 125(1), 62–74. Maleki, A., Hosseini, M. J., Rahimi, N., Abdollahi, A., Akbarfakhrabadi, A., Javadian, N., et al. (2019). Adjuvant potential of selegiline in treating acute toxicity of aluminium phosphide in rats. Basic & Clinical Pharmacology & Toxicology, 125(1), 62–74.
Metadaten
Titel
The Molecular Mechanism of Aluminum Phosphide poisoning in Cardiovascular Disease: Pathophysiology and Diagnostic Approach
verfasst von
Seyed Farzad Hosseini
Mehdi Forouzesh
Mohsen Maleknia
Samira Valiyari
Mahmood Maniati
Azin Samimi
Publikationsdatum
25.07.2020
Verlag
Springer US
Erschienen in
Cardiovascular Toxicology / Ausgabe 5/2020
Print ISSN: 1530-7905
Elektronische ISSN: 1559-0259
DOI
https://doi.org/10.1007/s12012-020-09592-4

Weitere Artikel der Ausgabe 5/2020

Cardiovascular Toxicology 5/2020 Zur Ausgabe