Skip to content
Publicly Available Published by De Gruyter May 6, 2020

Novel selective anticancer agents based on Sn and Au complexes. Mini-review

  • Elena R. Milaeva EMAIL logo , Dmitry B. Shpakovsky , Yulia A. Gracheva , Taisiya A. Antonenko , Tatyana D. Ksenofontova , Evgeny A. Nikitin and Daria A. Berseneva

Abstract

Cancer is one of the most common causes of death in modern medicine. Molecular design of novel substances with pharmacological activity is one of the goals of medicinal inorganic chemistry. Platinum complexes are widely used in the treatment of cancer, despite high efficacy their use is limited by side effects, as well as primary or acquired resistance. In this regard, the search for novel metal-containing antitumor compounds is underway. Organotins and gold compounds are promising pharmacological agents with anti-cancer properties. The introduction of protective antioxidant fragments into inorganic compounds molecules is a way to reduce the side effects of anti-cancer drugs on healthy cells. 2,6-dialkylphenols belonging to vitamin E (α-tocopherol) mimetics are widely used as antioxidants and stabilizers. The properties of Ph3SnCl (Sn-I), Ph3PAuCl (Au-I) and complexes Ph3SnSR (Sn-II) and Ph3PAuSR (Au-II) based on 2,6-di-tert-butyl-4-mercaptophenol (RSH) as radical scavengers and reducing agents were studied in model reactions. For Sn-II and Au-II the comparative study of cytotoxic action was made and the IC50 values on different cancer cell lines were found to be depended on the nature of metal. In general, Sn(IV) complexes possessed higher cytotoxicity than Au(I) complexes. In order to clarify the mechanism of cytotoxic mode of action the effect of compounds on Fe3+-induced lipid peroxidation, mitochondrial potential and mitochondrial permeability, cell cycle and induction of apoptosis was studied. Organotin compounds can bind tubulin SH-groups and inhibit its polymerization by a dose-dependent mechanism, whereas gold compounds inhibit Thioredoxin reductase (TrxR). In vivo experiments on acute toxicity of Sn-II and Au-II proved their moderate toxic action that opens prospects for the further study as antitumor agents.

Introduction

The discovery of the anticancer properties of cisplatin and its analogues (oxaliplatin and carboplatin) – Pt (II) compounds, their wide clinical success in modern cancer treatment methods have caused great interest in the study of anticancer drugs based on metal compounds [1].

Compounds of the general formula RnSnX4-n, (X=Cl, Br) contain a Sn(IV) atom and are classified as mono-, di-, tri-, and tetraorganotins depending on the number of alkyl or aryl groups. The rapid growth of industrial, agricultural, and biological uses of organotin compounds over the past few decades has led to an increase in their concentration in the environment and in biological systems [2]. Since organotins molecules contain lipophilic organic groups, they are able to easily penetrate through the cell membrane, which leads to a rapid increase in their concentration in lipids. Migrating along the trophic chain, these toxicants pose a threat to human health [3]. In addition, Sn is a biogenic element necessary for calcification processes; therefore, the correct development of the skeleton is impossible without it. The biological activity of Sn(IV) compounds is sharply different from the action of elemental tin and Sn(II) inorganic salts. An intensive study of organotin compounds began after the discovery of their antitumor activity by M. Gielen in the 1980s. The researcher’s attention was attracted mainly by organotin compounds which possess high antitumor activity that opens up the possibility of their study as promising pharmacological substances [4], [5], [6], [7], [8]. Organotin(IV) carboxylates with o- or p-hydroxybenzoic acids showed high cytotoxicity against sarcoma cancer cells [9]. The Bu3Sn and Ph3Sn-complexes with 2-thiobarbituric acid demonstrate higher cytotoxic activity than that of cisplatin against human breast adenocarcinoma cells (MCF-7), their IC50 values were 272 and 179-fold lower than that of cisplatin, respectively [10], [11].

It is well known that the Sn atom interacts with free SH groups in proteins, which leads to distortion of the structure and loss of protein functionality. Organotin chlorides can promote lipid peroxidation in cell membranes and cause oxidative stress [12]. It has been shown that RnSnX4-n exhibiting antitumor activity bind and unwind DNA, affect cell death as a result of necrosis, increase the influx of extracellular Ca2+ and generate reactive oxygen species (ROS) that leads to apoptosis [13]. Compounds Sn(IV) act on cells by various mechanisms: changes in membrane permeability, suppression of the oxidative phosphorylation and ATP synthesis in mitochondria, accumulation in the Golgi apparatus and endoplasmic reticulum, influence on DNA, RNA and protein synthesis, induction of apoptosis [14]. Studies of the binding of DNA to organotin compounds are of great importance for the production of new antitumor agents [15]. Moreover, the correlation between anti-tumor action of organotin(IV) complexes with thioamide ligands and inhibition of enzyme lipoxygenase by compounds in catalytic oxidation of linoleic acid to hydroperoxylinoleic acid was discovered [16], [17], [18]. The tin drug Purlytin (Rostaporfin) is used in clinical practice (Scheme 1) [19]. Purlytin – tin etiopurpurin – belongs to the class of metal complexes with tetrapyrrole macrocycles as ligands, and is used as a photosensitizer to treat age-related macular degeneration. The drug was approved by the FDA in 2004 for therapy in ophthalmology. However, Miravant Medical Technologies is currently conducting phase II/III clinical trials for PDT for metastatic breast cancer.

Scheme 1: 
Examples of Sn(IV) and Au(I) drugs purlytin, auranofin, aurothioglucose and aurothiomalate.
Scheme 1:

Examples of Sn(IV) and Au(I) drugs purlytin, auranofin, aurothioglucose and aurothiomalate.

Among several classes of metal complexes (Ru, Pd, Ti, Cu, Au, etc.) which were noted as promising anticancer agents, many studies focus on various gold compounds in oxidation states +1 and +3. This is because the compounds Au(I) are considered as possible antiproliferative agents in connection with their long-term use in the clinic for the treatment of rheumatoid arthritis. Au(III) compounds form flat square complexes, isoelectronic and isostructural Pt (II) complexes, due to which they can have a similar biological effect.

Gold and its complexes have historically been used to treat a wide range of diseases, but the rational use of Au-based compounds in medicine began in the early 1920s, when the bacteriostatic effect of the metal in vitro was proved. The characteristic oxidation states for gold are Au(I) and Au(III), and Au(0) is found in colloidal solutions. The first known gold salt acting as a drug was gold cyanide K[Au(CN)2], used by bacteriologist Robert Koch to kill mycobacteria, the causative agent of tuberculosis [20]. In 1929, J. Forestier proposed the use of gold salts for the treatment of rheumatoid arthritis [21]. Gold compounds such as auranofin, aurothioglucose and aurothiomalate in which gold is bound to a sulfur atom are widely used as antirheumatic drugs (Scheme 1).

Water-soluble compounds are used in the clinic and contain hydrophilic groups (e.g. glucose) as well as an aurothiogroup (–S–Au–). The high affinity for sulfur and the inhibitory effect of gold salts on various enzymes suggested that the therapeutic effect of Au compounds may be due to inhibition of binding with sulfhydryl residues [22]. Recent studies have shown that auranofin and other gold compounds bearing biologically relevant molecules demonstrate antitumor properties and they are able to overcome cisplatin resistance in cancer cells of the human ovary, confirming the suggestion that the mechanism of action of this compound is different from the damage to DNA caused by cisplatin [23]. In particular, it was found that auranofin acts as a powerful inhibitor of thioredoxin reductase (TrxR) and causes a change in the cell redox status, which leads to an increase in the production of superoxide radical anion, H2O2 and oxidation of the components of the thioredoxin (Trx) system [24], [25], [26].

The lipoxygenase inhibition and binding affinity towards the calf thymus DNA by triphenylphosphine gold(I) complexes with the thioamides was studied [27], [28], [29].

Main problems in creating drugs are the rational “activity-toxicity” ratio and the reduction of side effects. The development of organometallic therapeutic agents is hampered by their toxicity. Since Sn(IV) and Au(I) in auranofin analogs can induce oxidative stress leading to cytotoxicity, many attempts have been made to minimize the pro-oxidant effect and generation of reactive oxygen species (ROS) [30]. Including of additional 2,6-di-tert-butylphenol group in organotin complexes decreased the cytotoxicity of compounds [31]. These results indicated that attenuation of pro-oxidative properties of tin and gold containing drugs by introduction of radical scavenging moiety is a key prerequisite for the rational design of anticancer therapeutics with balanced redox potential [32]. Hence, numerous studies are being conducted on the modification of approved preparations based on metal compounds, including gold and tin. To reduce the side effects of anticancer drugs on healthy cells the introduction of substituents with protective (antioxidant) activity in the molecule seems reasonable.

Antioxidants play an important role in the regulation of free-radical processes in the body. One of the most promising directions in the chemistry of antioxidants is the synthesis of hybrid compounds combining antioxidant activity with the ability of structural interaction with a biosystem. Such compounds include 2,4- and 2,6-dialkylsubstituted phenols, in particular, 2,6-di-tert-butylphenols. These compounds exhibit a wide range of biological activity: anti-inflammatory, antimicrobial, antiviral, etc. The antioxidant defense system in a living body regulates fluctuations in the prooxidant-antioxidant balance and protects cells from damage caused by high levels of oxidative stress. Analogues of the natural antioxidant α-tocopherol are widely used as inhibitors of the formation of free radicals in the oxidative destruction of natural and synthetic substrates [33]. Phenolic compounds such as 2,6-di-tert-butyl-4-methoxyphenol (BGT, E320) and 2,6-di-tert-butyl-4-methylphenol (BHT, E321) are examples of some synthetic antioxidants. Their activity is related with mobility of the hydrogen atom of the phenolic OH group, which is easily cleaved by the action of peroxyl radicals, forming a relatively unreactive phenoxyl radical, breaking off the radical chain process [34].

Sterically hindered phenols inhibit effectively the oxidation of various organic substrates [35]. In the case of sterically hindered phenols, the formation of hydrogen bonds is impossible. The effectiveness of 2,6-di-tert-butylphenols as inhibitors of the oxidative destruction of hydrocarbons is determined by the nature of the group in the para position of the aromatic ring, which affects the stability of the corresponding phenoxyl radicals formed during the oxidation process [36]. It is worth to note that in cancer organs and tissues where the pH value is close to five the principal OH group of phenolic antioxidants is protonated that can bock the activity of antioxidant in cancer media. On the other hand, in normal tissues and cells at physiological pH the antioxidant is active.

Our group have reported the antiproliferative activity of organotin(IV) complexes bearing 2,6-di-tert-butylphenol fragment with heterocyclic thioamides. The IC50 value for bis-(3,5-di-tert-butyl-4-hydroxyphenyl)tin complex R2Sn(MPMT)2 (where MPMT=2-mercapto-4-methylpyrimidine, R=3,5-di-tert-butyl-4-hydroxyphenyl) was 32-fold more cytotoxic than cisplatin at MCF-7 cancer cells [37]. In order to lower the toxic effect of organotins the use of antioxidants is suggested. A series of organotin complexes based on 2,6-di-tert-butyl-4-mercaptophenol were synthesized and tested against cancer and normal cells, also the ability of complexes to decrease the content of SH groups in tubulin allows one to consider these compounds as potential antimitotic agents. Complexes exhibited significantly lower cytotoxic activity against normal MRC-5 cell line compared to the tumor cell lines MCF-7 and HeLa [31].

Living organism possesses sophisticated protective antioxidant system, which prevents oxidative stress. It consists of low-molecular natural antioxidants (ascorbic acid, tocopherols, lipoic acid, and carotenoids, etc.) and enzymatic antioxidants such as superoxide dismutase (SOD), catalase, glutathione and glutathione dependent enzymes, regulates the toxic impact of ROS and prevents cell damage. There are several mechanisms of antioxidants action depending on the interaction of antioxidative agents with various ROS: a dismutation of superoxide radical-anion by SOD, scavenging of hydroxyl radical HO˙˙, interaction either with peroxyl radicals of lipids LOO˙ or lipid hydroperoxides LOOH breaking down the chain radical process of lipid peroxidation. Recent results in testing metal complexes with 2,6-di-tert-butylphenol moiety showed that they could potentially facilitate the scavenging of excess ROS, and thus normalize redox balance in the damaged cells and organs [38].

In this regard, the study of the biological activity of tin and gold compounds with 2,6-di-tert-butylphenol fragment of urgent importance.

Biological activity of Sn(IV) and Au(I) thiolates

The activity of organotins is determined by: (1) the nature of the substituents; (2) the coordination availability of the Sn atom to form a bond with the target; (3) a relatively stable L–Sn bond (for example, S–Sn); (4) slow hydrolytic destruction [39]. Synthesis and complex study of the biological activity of Sn and Au complexes containing antioxidant 2,6-di-tert-butylphenol fragment was carried out recently. The structures of compounds (Scheme 2) were confirmed by 1H, 13C NMR, X-ray diffraction and elemental analysis. The mechanisms of their biological action and “structure – activity” correlations has been intensively studied. A comparative study of precursors Ph3SnCl (Sn-I), AuPPh3Cl (Au-I) and 2,6-di-tert-butylphenol (RSH) and their complexes Ph3SnSR (Sn-II) and Ph3PAuSR (Au-II) in the model processes of interaction with ROS and electron transport was carried out. Cytotoxicity in vitro on human cancer cell lines, cell cycle and influence of various structural fragments on the activity of compounds were studied [31], [32], [40].

Scheme 2: 
Structures of tin and gold complexes based on 2,6-di-tert-butyl-4-mercaptophenol. Molecular structure of compounds Sn-II and Au-II (hydrogen atoms are omitted for clarity).
Scheme 2:

Structures of tin and gold complexes based on 2,6-di-tert-butyl-4-mercaptophenol. Molecular structure of compounds Sn-II and Au-II (hydrogen atoms are omitted for clarity).

Antioxidant properties in model systems

The study of redox activity using the CUPRAC test

The redox activity was studied in CUPRAC (Cupric Ion Reducing Antioxidant Capacity) test [41]. This method is based on the reduction of Cu(II) ions by the studied compound forming Cu(I) complex with neocuproin (2,9-dimethyl-1,10-phenanthroline), which has a maximum absorption at 450 nm. Trolox (6-hydroxy-2,5,7,8-tetramethylchromane-2-carboxylic acid) was used as standard, the results for complexes and starting compounds are presented in Trolox equivalents antioxidant capacity (TEAC) (Table 1) [32], [42].

Table 1:

TEAC values for the studied compounds and Trolox.

Compound Trolox RSH Au-I Au-II Sn-I Sn-II
TEAC 1.00±0.03 1.24±0.03 0.12±0.02 1.05±0.07 0.23± 0.02 1.69±0.13

The initial chlorides Au-I and Sn-I are inactive in the Cu(II) reduction. The most active reducing agent was Sn-II for which the Trolox equivalent was 1.7 times higher than Trolox and 1.4 times higher than organic precursor RSH. Introduction of a phenolic fragment in gold compound Au-II resulted in less antioxidant activity than the Sn-II. It can be assumed that SnPh3 fragment increases the stability of the phenoxyl radical is responsible for the antioxidant activity.

DPPH-test

The radical scavenging activity of compounds has been studied in the process of hydrogen atom transfer from the phenol moiety to the stable free radical 2,2-diphenyl-1-picrylhydrazyl (DPPH) to give diphenylpicrylhydrazine and phenoxyl radicals (Scheme 3) which can undergo further reactions such as coupling, fragmentation and addition [43].

Scheme 3: 
Mechanism of radical scavenging activity of 2,6-di-tert-butylphenols with DPPH.
Scheme 3:

Mechanism of radical scavenging activity of 2,6-di-tert-butylphenols with DPPH.

Antiradical activity was evaluated as the amount of antioxidant necessary to decrease the initial concentration of DPPH by 50% (Efficient Concentration=EC50). The EC50 values for Sn-II and Au-II were 15±4 and 31±2 μM, respectively, whereas the activity of RSH was several times lower (55±10 μM). Compounds Au-I and Sn-I were not active in this test since they do not possess active OH group. The antioxidant activity depends on metal atom nature in the phenol containing complexes. Thus, the antioxidant activity of gold and tin compounds bearing hindered phenol group is mainly determined by the ability to hydrogen atom abstraction from the –OH group and such compounds possess radical scavenging activity.

Evaluation of the ability of compounds to react with NADPH

NADPH affects the pool of adenine nucleotides which plays an important role in antioxidant defense system. The ability of compounds to oxidize NADPH was monitored spectrophotometrically in the presence of NADPH at λmax 340 nm for 30 min. Since the absorbance and structure of the spectra did not change during incubation, the absence of a reaction between the substances and NADPH can be postulated.

Reaction with ROS

Interaction with superoxide radical anion O2˙

The activity of the complexes with respect to the superoxide radical anion O2˙ generated in the xanthine/xanthine oxidase enzymatic system was determined by the change in the reduction rate of tetrazolium blue to formazan (Fig. 1) [44]. All compounds were practically inactive, the initial Au-I gold complex showed little activity (Fig. 1b).

Fig. 1: 
Kinetic curves of formazane accumulation in the xanthine/xanthine oxidase system in the presence of different compounds (a) and different concentrations of the Au-I (b).
Fig. 1:

Kinetic curves of formazane accumulation in the xanthine/xanthine oxidase system in the presence of different compounds (a) and different concentrations of the Au-I (b).

Interaction with hydroxyl radical ˙OH

The oxidation of Fe2+ ions by hydrogen peroxide (Fenton reaction) was used to generate the hydroxyl radical which reacts with D-deoxyribose to form malonic dialdehyde (MDA), whose concentration was determined spectrophotometrically by reaction with thiobarbituric acid leading to trimethine complex with maximum absorption λmax 532 nm [45]. The results are presented in % of control (Table 2). The antioxidant ionol (BHT) was used for comparison.

Table 2:

The activity of the studied compounds and ionol in the reaction with ˙OH.

Compound ionol RSH Au-I Au-II Sn-I Sn-II
Activity, % 77 74 83 112 103 88

Since RSH demonstrates high antioxidant activity, the introduction of the RS fragment into the organotin complex increases its activity, the opposite situation is observed for the Au-II. Apparently, for the initial Au-I, a ligand exchange reaction involving ˙OH can take place.

Assessment of the ability of compounds to react with H2O2

Spectrophotometric monitoring of the reactivity of the studied compounds with respect to H2O2 in the 200–350 nm range [46] showed no changes in the structure of the spectra during incubation (10 min) indicating the absence of interaction between substances.

Fe3+-induced lipid peroxidation

Mitochondria are the main source of intracellular ROS generation. Abnormal level of ROS causes the oxidative damage of intracellular structures as proteins, nucleic acids, lipids, membranes and other biomolecules with consequent injury to cells. ROS are proposed to be involved in all the steps of carcinogenesis [38]. Lipid peroxidation of mitochondrial membranes leads to disturbances of oxidative phosphorylation, a major source of energy. The monitoring of lipid peroxidation in mitochondrial suspension can be followed by the accumulation of substances that reacted with thiobarbituric acid (TBARs) [47]. Oxidative damage (lipid peroxidation) was induced by Fe3+. RSH was found to be effective antioxidant with (IC50 0.2 μM), whereas Sn-II demonstrated moderate antioxidant activity (IC50 3.1 μM) unlike Sn-I with prooxidant properties. This effect is observed most likely due to the presence of a metal atom (Sn) in their structure [48].

The influence of gold complex Au-II and their precursors RSH and Au-I on lipid peroxidation of isolated rat brain mitochondria was studied. In contrast to Au-I, RSH and the Au-II similarly decreased the TBARs concentration upon Fe3+-induced lipid peroxidation of mitochondria (Fig. 2). The tendency to decrease the TBARs level in the presence of phenolcontaining complexes and RSH leads one to suggest the main role of the phenol moiety in the antioxidant mechanism.

Fig. 2: 
Influence of gold complex Au-II and precursors (RSH and Au-I) on Fe3+-induced lipid peroxidation of isolated rat brain mitochondria.
Fig. 2:

Influence of gold complex Au-II and precursors (RSH and Au-I) on Fe3+-induced lipid peroxidation of isolated rat brain mitochondria.

Toxicity of Sn and Au compounds

MTT test

Cytotoxicity of compounds RSH, RSAuPPh3 (Au-II), RSSnPh3 (Sn-II), their precursors ClAuPPh3 (Au-I), ClSnPh3 (Sn-I), as well as mixtures of RSH with Au-I and Sn-I in 1:1 ratio were studied by MTT test on the cancer cell lines MCF-7, HCT-116 and A549 (lung cancer) and conditionally normal cells MCF10A (breast epithelial cells) and VA13 (embryonic lung fibroblasts). Cell survival curves were obtained and IC50 values were determined (Table 3). The introduction of a phenolic fragment in the case of Au-II reduced the toxicity compared to precursors Au-I and Sn-I. RSH and Au-II demonstrated low toxicity up to >50 μM. Moreover, the study of mixtures of Sn-I and RSH showed no significant change. Sn-II was 6 times less toxic in comparison with Sn-I at the HCT-116 cells. Consequently its toxicity was studied on other cancer cell lines (Table 4) [42].

Table 3:

IC50 values for the studied compounds on HCT-116 and MCF-7 cell lines.

Compound IC50, μM
HCT-116 MCF-7
RSH >50 >50
Au-I 7.7±1.3 12.1±1.5
Au-II >50 >50
Sn-I 0.01±0.001 0.07±0.01
Sn-II 0.06±0.01 [42] 0.09±0.01 [42]
Cisplatin 6.5±0.5 18.5±0.2 [31]
Table 4:

The IC50 values for the compounds of Sn-II and cisplatin on different cancer cell lines [42].

Compound IC50, μM
CaCo-2 SNB-19 A-498 M-14 NCl-H332M
Sn-II 0.09±0.02 0.06±0.03 0.04±0.01 0.15±0.03 0.10±0.01
Cisplatin 62.3±1.2 0.81±0.03 17.5±1.2 10.8±4.5 2.7±0.3
  1. Colon (CaCo-2), glioblastoma (SNB-19), renal (A498), melanoma (M14), non-small cell lung (NCI-H322M) cell lines.

Study of apoptosis and cell cycle

The effect of compounds on cell death of cultured cells during incubation for 48 h with compounds at 2·IC50 concentration was studied. After incubation, the cells were removed from the substrate, washed and the percentage of apoptotic and necrotic cells was detected using a cytofluorimetry. This analysis is base on the detection of phosphatidylserine (PS) on the surface of apoptotic cells using fluorescently labeled Annexin V in combination with the dead cell marker 7-AAD [49]. Annexin V is a Ca2+-dependent phospholipid binding protein that has a high affinity for PS, a membrane component usually localized to the inner surface of the cell membrane [50]. At the beginning of apoptosis, PS molecules move to the outer surface of the cell membrane, where Annexin V can easily bind them. At the late stage apoptotic cells show a loss of membrane integrity. Membrane-impervious dye 7-AAD is used to distinguish dead cells from early apoptotic cells. Analysis of apoptotic profile can distinguish four cell populations:

(1) Viable cells that do not undergo apoptosis: Annexin V (−) and 7-AAD (−); (2) Early apoptotic cells: Annexin V (+) and 7-AAD (−); (3) Late apoptotic cells: Annexin V (+) and 7-AAD (+); (4) Cells that have died by non-apoptotic pathway: Annexin V (−) and 7-AAD (+). The obtained results of the distribution of 7-AAD-positive and negative and AnV-positive and negative cells in % are presented in Fig. 3 and Table 5 for MCF-7 and HCT-116 cells.

Fig. 3: 
Apoptotic profile of cancer cells MCF-7 after treatment with different compounds at μM after 48 h. Concentration of compounds 2·IC50.
Fig. 3:

Apoptotic profile of cancer cells MCF-7 after treatment with different compounds at μM after 48 h. Concentration of compounds 2·IC50.

Table 5:

Distribution (%) of 7-AAD-positive and negative and Annexin V (AnV)-positive and negative MCF-7 and HCT-116 cells.

Compound AnV(−)/7-AAD(−) AnV(+)/7-AAD(−) AnV(−)/7-AAD(+) AnV(+)/7-AAD(+)
MCF-7
 Control 87.4% 9.5% 0.2% 2.9%
 RSH 90.5% 9.4% 0.1% 0.1%
 Au-I 24.4% 40.0% 0.8% 34.8%
 Au-II 32.1% 31.7% 1.0% 35.2%
 Sn-I 46.3% 36.7% 0.7% 16.3%
 Sn-II 39.3% 54.5% 0.3% 5.9%
HCT-116
 Control 71.4% 13.0% 2.4% 13.3%
 RSH 72.4% 16.7% 7.3% 3.6%
 Au-I 11.5% 3.2% 0.2% 85.2%
 Au-II 26.9% 10.1% 2.1% 61.0%
 Sn-I 3.9% 17.3% 0.5% 78.3%
 Sn-II 10.9% 26.0% 0.2% 63.0%

For the MCF-7 cells Au-I and Au-II complexes after 48 h of incubation the percentage of late apoptotic cells is maximal. However, there are no significant differences in the percentage of cells in the pool of early and late apoptotic cells for the Au-I compound and Au-II with phenol phragment. The RSH does not induce tumor cell apoptosis. It was shown previously that the Sn-II induces not only depolarization of mitochondria but also mitochondrial swelling and opens the mitochondrial permeability transition pores (MPTp) leading to the cell death [51].

Incubation with compound Sn-II increases in the pool of cells of early apoptosis by 1.5 times compared to the original compound Sn-I, but the percentage of late apoptotic cells was 3 times less. After 48 h of incubation HCT-116 cells were more sensitive to the action of the studied compounds. The percentage of late apoptotic cells for tin and gold complexes was higher than 50%. For Sn-II and Au-II complexes containing a phenolic group the number of late apoptotic cells was lower than early apoptotic cells compared to the precursors Au-I and Sn-I. Thus, for HCT-116 cells the effect of the antioxidant fragment R in Au-II and Sn-II retards the development of apoptosis. For both cell lines, the percentage of late apoptotic cells in the starting substances was higher than in the complexes, furthermore, the number of cells in the early apoptosis stage was higher for compounds with an antioxidant fragment (Fig. 4).

Fig. 4: 
Percentage of MCF-7 and HCT-116 cells in early and late apoptosis after 48 h of incubation.
Fig. 4:

Percentage of MCF-7 and HCT-116 cells in early and late apoptosis after 48 h of incubation.

The effect of compounds in 2·IC50 concentration on the cell cycle was studied by mixture of reagents including the nuclear DNA intercalating dye propidium iodide (PI) and RNases in a proprietary formulation after incubation for 24 and 72 h [50].

The technique is based on the determination of DNA content by PI staining to recognize and measure the percentage of cells in each phase of the cell cycle (G0/G1, S and G2/M). PI distinguishes cells at different stages of the cell cycle based on different DNA content in the presence of RNase which increases the specificity of DNA staining. Cells in the G0/G1 phase contain two copies of each chromosome. As cells progress through the cell cycle, they synthesize chromosomal DNA (S-phase). The fluorescence intensity from PI increases until all chromosomal DNA doubles (G2/M phase). At this stage, G2/M cells fluoresce at twice the intensity of the G0/G1 population. G2/M cells divide into two daughter cells at the end of the cycle. The effect of the studied compounds on the cell cycle for MCF-7 and HCT-116 cells is presented in Table 6. For the MCF-7 cell line the effect of the compounds is insignificant. It can be noted that Au-II complex presumably blocks G0/G1 phase. For Au-II and Sn-II complexes decreased the part of cells in G0/G1 phase that seems to be associated with the appearance of dead cells whose DNA is destroyed. The influence of the studied compounds on the cell cycle for HCT-116 cells is presented in Figs. 5 and 6.

Table 6:

Effect of compounds on the cell cycle of MCF-7 and HCT-116.

Compound Incubation time, h MCF-7 HCT-116
G0/G1 S G2/M G0/G1 S G2/M
Control 24 58.9 15.8 20.1 36.0 25.3 22.7
72 74.3 12.1 11.1 47.8 22.8 21.2
RSH 24 58.6 16.0 19.9 44.6 17.1 26.0
72 73.6 8.6 12.9 45.1 18.8 23.2
Au-I 24 54.6 19.6 19.9 39.2 27.2 19.7
72 72.0 10.6 14.3 50.5 24.3 19.5
Au-II 24 65.7 15.8 14.9 43.6 22.7 22.7
72 49.6 18.3 22.2 52.1 20.5 22.5
Sn-I 24 45.5 15.7 31.3 39.3 24.0 26.7
72 69.0 13.3 14.5 50.5 20.6 23.2
Sn-II 24 66.1 15.9 12.1 31.7 20.3 31.2
72 52.4 11.6 14.7 36.8 18.3 32.5
Fig. 5: 
Cell cycle analysis by flow cytometry in HCT-116 cancer cells were treated by RSH, Sn-I and Sn-II for 24 and 72h, concentration of compounds 2·IC50.
Fig. 5:

Cell cycle analysis by flow cytometry in HCT-116 cancer cells were treated by RSH, Sn-I and Sn-II for 24 and 72h, concentration of compounds 2·IC50.

Fig. 6: 
Cell cycle analysis by flow cytometry in HCT-116 cancer cells were treated by Au-I and Au-II for 24 and 72h, concentration of compounds 2·IC50.
Fig. 6:

Cell cycle analysis by flow cytometry in HCT-116 cancer cells were treated by Au-I and Au-II for 24 and 72h, concentration of compounds 2·IC50.

Au-I and Sn-I do not change essentially the cell cycle of HCT-116 cells. The Au-II causes a block at the G0/G1 stage. RSH results in a minor G1 phase block after 24 h of incubation. Sn-II causes a marked block of the cell cycle in the G2/M phase after 24 and 72 h and initiate cell death. The results demonstrate that the effect on cell cycle depends on compounds and cells type. The pronounced influence was found for HCT-116 cells. High cytotoxicity of Sn-II towards MCF7 and HCT-116 cancer cells indicates the perspectives of this compound for preclinical studies.

Mechanisms of Sn and Au cytotoxicity

The possible mechanisms of Sn-II cytotoxicity including tubulin binding and influence on tubulin assembly in microtubules was investigated previously [51]. Sn-II revealed the significant inhibition of tubulin polymerization without disturbances of microtubules structure (Fig. 7).

Fig. 7: 
The influence of Sn-II on tubulin polymerization (a) kinetic curve of tubulin polymerization as the change in fluorescence intensity; (b) the concentration dependence of tubulin polymerization rate (Vp) in the presence of compound Sn-II.
Fig. 7:

The influence of Sn-II on tubulin polymerization (a) kinetic curve of tubulin polymerization as the change in fluorescence intensity; (b) the concentration dependence of tubulin polymerization rate (Vp) in the presence of compound Sn-II.

Since tubulin is a sulphhydryl-rich protein with 20 cysteine residues distributed across both subunits, it can react with sulphhydryl-directed reagents, the interaction of Sn-II with tubulin did not provide distinct evidences for the putative binding modes and modes of action of the compounds. For insight of mode of interaction between Sn-II and tubulin, docking simulation for this compound towards well-studied colchicine, paclitaxel and vinblastine binding pockets in the tubulin molecule were carried out. The main contribution into the calculated binding energy score of compoundы was provided by van-der-Waals interactions. A large size of the molecule hampers binding with colchicine site. Predicted binding modes of the compound in paclitaxel and vinblastine binding sites are much more reasonable (Fig. 8).

Fig. 8: 
The best scored orientations of compound Sn-II (a, b) in the ligand binding sites of 4EB6 (a) and 1JFF (b). Cognate ligands [vinblastine (a) and paclitaxel (b)] are shown as blue translucent surfaces, organotin is shown as balls-and sticks and red coloured surface [51].
Fig. 8:

The best scored orientations of compound Sn-II (a, b) in the ligand binding sites of 4EB6 (a) and 1JFF (b). Cognate ligands [vinblastine (a) and paclitaxel (b)] are shown as blue translucent surfaces, organotin is shown as balls-and sticks and red coloured surface [51].

Since organotins form significantly fewer hydrogen bonds than natural compounds, their binding is ruled mostly by shape complementarity. Sn-II can overlap with paclitaxel and vinblastine sites, but their totally different shape requires specific binding modes and this compound cannot penetrate into the deepest parts of the pockets.

Thioredoxin reductase (TrxR) – redox enzyme which maintains the balance between pro- and anti-oxidative states in living cells. TrxR as a major target of auranofin been inhibited by auranofin leads to an oxygen stress, cytotoxicity and induced damage of tissues in which auranofin accumulates that complicate the practical use of auranofin. Many attempts have been made to minimize the pro-oxidant effect of auranofin. Modulation of mammalian TrxR was investigated as a tentative drug target. The Au-II inhibited TrxR with an IC50 value of 0.57 ±0.15 mM. Auranofin showed a higher activity (IC50 0.016±0.003 mM) whereas RSH had no effect at up to 10 mM [32]. These results justified our attempts of chemical modification of Au-I leading to derivatives with an attenuated prooxidant potency.

Toxicity of Sn and Au compounds in vivo

The experiments were performed on white Balb/c mice weighing 18–20 g, age 8–10 weeks. Compounds Sn-II and Au-II were injected into the abdominal cavity in physiological buffer solution. The volume of injected solutions is 0.5 ml, the volume of solvent (DMSO) does not exceed 125 μl. In the initial experiments, the upper limit of tolerability of administered doses was estimated. Two doses were chosen: 62.5 mg/kg and 125 mg/kg (1.25 mg/mouse and 2.5 mg/mouse, respectively). Each group had 4 animals. After 7 days all mice in both groups (62.5 mg/kg Sn-II or Au-II) were alive. In Sn-II group (125 mg/kg) two mice died on third day. In Au-II (125 mg/kg) group all mice died after first day of experiment. The results of the initial evaluation showed that for compounds Sn-II and Au-II, a dose of 125 mg/kg was lethal. In the following experiments, animals received doses of compounds Sn-II and Au-II from 70 mg/kg to 110 mg/kg once intraperitoneal injection in each group, 6 mice. Observations were continued for 21 days. Table 7 shows the results of the study of acute general resorptive toxicity of compounds Sn-II and Au-II. The introduction was accompanied by toxic manifestations (behavioral-stooping, numbness, sometimes inactivity). When administered doses of 70–80 mg/kg, these manifestations lasted from several hours to 1 day, after which the condition of the animals was restored and, in the future, did not differ from that of intact mice. For compounds Sn-II and Au-II, administration of doses >80 mg/kg was fatal for some animals. Thus, 80 mg/kg is the maximum tolerated dose.

Table 7:

Survivability of Balb/c mice with intraperitoneal injection of compounds Sn-II and Au-II.

Compound Dose, mg/kg
70 80 90 100 110
Sn-II 0/6 0/6 3/6 5/6 5/6
Au-II 0/6 0/6 3/6 5/6 6/6

Conclusions

The synthesis of Sn(IV) and Au(I) complexes bearing 2,6-di-tert-butylphenol moiety has been accomplished by our research group. The results of studies of Sn(IV) and Au(I) compounds indicate a difference in the modes of compounds action depending on the nature of the metal type and presence of antioxidant fragment. While Ph3SnCl and Ph3PAuCl compounds act as prooxidants, complexes Ph3SnSR and Ph3PAuSR based on 2,6-di-tert-butyl-4-mercaptophenol (RSH) can act as radical scavengers, reducing agents and effective inhibitors of induced mitochondrial lipid peroxidation and provides the transport to biological targets in normal cells. Complexes of Sn exhibit significantly higher cytotoxicity than Au compounds. Tin compounds exhibit an extraordinary antiproliferative properties in different cancer cells and 2,6-di-tert-butyl-4-mercaptophenol ligand decreases cytotoxic activity. The introduction of hindered phenol groups decreases the cytotoxicity of complexes. Organotin complex Ph3SnSR depolarizes the mitochondria, induces the mitochondrial permeability transition and causes blocking of the stage preceding mitosis that may be the main reasons of its cytotoxicity. Ph3PAuSR attenuated the potency of the reference drug Auranofin to inhibit TrxR and can acts as a broad specificity organ protector against oxygen burst. Experiments in vivo demonstrated that the 5 mg·kg−1 daily for 14 days was safe and efficient dose.

The polytopic Sn(IV) and Au(I) compounds combining 2,6-di-tert-butylphenol pendant and metal center in one molecule are membrane active compounds and may be studied with the aim to find novel metal-based drugs with low undesirable toxicity.


Article note

A collection of invited papers based on presentations at 21st Mendeleev Congress on General and Applied Chemistry (Mendeleev-21), held in Saint Petersburg, Russian Federation, 9–13 September 2019.


Acronyms

A498

Renal cell line

A549

Lung cancer cell line

7-AAD

7-Aminoactinomycin D

BHT

2,6-di-tert-butyl-4-methylphenol

BGT

2,6-di-tert-butyl-4-methoxyphenol

CaCo-2

Colon cell line

CUPRAC

Cupric Ion Reducing Antioxidant Capacity

DCFH-DA

2′,7′-Dichlorofluorescin diacetate

DNA

Deoxyribonucleic acid

DPPH

2,2-diphenyl-1-picrylhydrazyl

DMSO

Dimethyl sulfoxide

EC50

Efficient Concentration

FDA

Food and Drug Administration

GFP

Green fluorescent proteins

HCT-116

Human colon cancer cell line

MDA

Malonic dialdehyde

MCF-7

Breast cancer cell line

MCF10A

Conditionally normal cells (breast epithelial cells)

M14

Melanoma

MitoSOX

Mitochondrial Superoxide Indicator

MPTp

Mitochondrial permeability transition pores

NADPH

Nicotinamide adenine dinucleotide phosphate

NCI-H322M

Non-small-cell lung carcinoma cell line

PDT

Photodynamic therapy

PI

Propidium iodide

PS

Phosphatidylserine

ROS

Reactive oxygen species

RNA

Ribonucleic acid

SNB-19

Glioblastoma cell line

TBARs

Thiobarbituric acid reactive substances

TEAC

Trolox equivalents antioxidant capacity

TrxR

Thioredoxin reductase

Trx

Thioredoxin

VA13

Embryonic lung fibroblasts

  1. Funding: The authors are grateful to Kharitonashvili E.V. for fruitful discussion. Experiments in vivo were supported by the Russian Science Foundation grant № 19-13-00084 and the Russian Foundation for Basic Research grants № 18-03-00203, № 19-33-90236 (synthesis of complexes), № 19-53-12042 (biological experiments).

References

[1] E. R. Milaeva, V. Y. Tyurin. Pure Appl. Chem. 89, 1065 (2017).10.1515/pac-2016-1130Search in Google Scholar

[2] A. Pagliarani, F. Trombetti, V. Ventrella (Eds.), Biochemical and Biological Effects of Organotins, Bentham eBooks, Sharjah (2012).10.2174/97816080526531120101Search in Google Scholar

[3] E. R. Milaeva, V. S. Petrosyan, N. T. Berberova, Y. T. Pimenov, L. Pellerito. Bioinorg. Chem. Appl. 2, 69 (2004).10.1155/S1565363304000068Search in Google Scholar

[4] C. N. Banti, S. K. Hadjikakou, T. Sismanoglu, N. Hadjiliadis. J. Inorg. Biochem. 194, 114 (2019).10.1016/j.jinorgbio.2019.02.003Search in Google Scholar

[5] J. Devi, J. Yadav. Anti-Cancer Agents Med. Chem. 18, 335 (2018).10.2174/1871520617666171106125114Search in Google Scholar

[6] S. K. Hadjikakou, N. Hadjiliadis. Coord. Chem. Rev. 253, 235 (2009).10.1016/j.ccr.2007.12.026Search in Google Scholar

[7] M. N. Xanthopoulou, S. K. Hadjikakou, N. Hadjiliadis, M. Schürmann, K. Jurkschat, J. Binolis, S. Karkabounas, K. Charalabopoulos. Bioinorg. Chem. Appl. 1, 227 (2003).10.1155/S1565363303000189Search in Google Scholar

[8] M. N. Xanthopoulou, S. K. Hadjikakou, N. Hadjiliadis, M. Schürmann, K. Jurkschat, A. Michaelides, S. Skoulika, T. Bakas, J. Binolis, S. Karkabounas, K. Charalabopoulos. J. Inorg. Biochem. 96, 425 (2003).10.1016/S0162-0134(03)00178-8Search in Google Scholar

[9] M. A. Abdellah, S. K. Hadjikakou, N. Hadjiliadis, M. Kubicki, T. Bakas, N. Kourkoumelis, Y. V. Simos, S. Karkabounas, M. M. Barsan, I. S. Butler. Bioinorg. Chem. Appl. 2009, 542979 (2009).10.1155/2009/542979Search in Google Scholar

[10] V. I. Balas, I. I. Verginadis, G. D. Geromichalos, N. Kourkoumelis, L. Male, M. B. Hursthouse, K. H. Repana, E. Yiannaki, K. Charalabopoulos, T. Bakas, S. K. Hadjikakou. Eur. J. Med. Chem. 46, 2835 (2011).10.1016/j.ejmech.2011.04.005Search in Google Scholar PubMed

[11] V. I. Balas, S. K. Hadjikakou, N. Hadjiliadis, N. Kourkoumelis, M. E. Light, M. Hursthouse, A. K. Metsios, S. Karkabounas. Bioinorg. Chem. Appl. MS. ID. 654137 (2008).10.1155/2008/654137Search in Google Scholar PubMed PubMed Central

[12] M. N. Kolyada, V. P. Osipova, N. T. Berberova, Yu. T. Pimenov, E. R. Milaeva. Macroheterocycles 10, 57 (2017).10.6060/mhc160542oSearch in Google Scholar

[13] A. Gennari, B. Viviani, C. L. Galli, M. Marinovich, R. Pieters, E. Corsini. Toxicol. Appl. Pharmacol. 169, 185 (2000).10.1006/taap.2000.9076Search in Google Scholar PubMed

[14] L. Niu, Y. Li, Q. Li. Inorg. Chim. Acta 423, 2 (2014).10.1016/j.ica.2014.05.007Search in Google Scholar

[15] G. K. Latsis, C. N. Banti, N. Kourkoumelis, C. Papatriantafyllopoulou, N. Panagiotou, A. Tasiopoulos, A. Douvalis, A. G. Kalampounias, T. Bakas, S. K. Hadjikakou. Int. J. Mol. Sci. 19, 2055 (2018).10.3390/ijms19072055Search in Google Scholar PubMed PubMed Central

[16] M. N. Xanthopoulou, S. K. Hadjikakou, N. Hadjiliadis, E. R. Milaeva, J. A. Gracheva, V. Y. Tyurin, N. Kourkoumelis, K. C. Christoforidis, A. K. Metsios, S. Karkabounas, K. Charalabopoulos. Eur. J. Med. Chem. 43, 327 (2008).10.1016/j.ejmech.2007.03.028Search in Google Scholar PubMed

[17] M. N. Xanthopoulou, S. K. Hadjikakou, H. N. Kourkoumelis, E. R. Milaeva, Y. A. Gracheva, V. Y. Tyurin, I. Verginadis, S. Karkabounas, M. Baril, I. S. Butler. Russ. Chem. Bull. 56, 767 (2007).10.1007/s11172-007-0115-zSearch in Google Scholar

[18] M. N. Xanthopoulou, N. Kourkoumelis, S. K. Hadjikakou, N. Hadjiliadis, M. Kubicki, S. Karkabounas, T. Bakas. Polyhedron 27, 3318 (2008).10.1016/j.poly.2008.07.030Search in Google Scholar

[19] Drugs R&D 5, 58 (2004). https://doi.org/10.2165/00126839-200405010-00013.10.2165/00126839-200405010-00013Search in Google Scholar PubMed

[20] R. Kour Sodhi, S. Paul. Cancer Ther. Oncol. Int. J. 14, MS.ID.555883 (2019).Search in Google Scholar

[21] T. Zou, C. T. Lum, C.-N. Lok, J.-J. Zhanga, C.-M. Che. Chem. Soc. Rev. 44, 8786 (2015).10.1039/C5CS00132CSearch in Google Scholar PubMed

[22] S. Tian, F.-M. Siu, S. C. F. Kui, C.-N. Lok, C.-M. Che. Chem. Comm. 47, 9318 (2011).10.1039/c1cc11820jSearch in Google Scholar PubMed

[23] V. Fernández-Moreira, R. P. Herrera, M. C. Gimeno. Pure Appl. Chem. 91, 247 (2019).10.1515/pac-2018-0901Search in Google Scholar

[24] C. Roder, M. J. Thomson. Drugs R&D 15, 13 (2015).10.1007/s40268-015-0083-ySearch in Google Scholar PubMed PubMed Central

[25] A. Casini, L. Messori. Curr. Top. Med. Chem. 11, 2647 (2011).10.2174/156802611798040732Search in Google Scholar PubMed

[26] C. Marzano, V. Gandin, A. Folda, G. Scutari, A. Bindoli, M. P. Rigobello. Free Rad. Biol. Med. 42, 872 (2007).10.1016/j.freeradbiomed.2006.12.021Search in Google Scholar PubMed

[27] K. N. Kouroulis, S. K. Hadjikakou, N. Kourkoumelis, M. Kubicki, L. Male, M. Hursthouse, S. Skoulika, A. K. Metsios, V. Y. Tyurin, A. V. Dolganov, E. R. Milaeva, N. Hadjiliadis. Dalton Trans. 47, 10446 (2009).10.1039/b909587jSearch in Google Scholar PubMed

[28] E. Georgiou, A. Metsios, N. Kourkoumelis, S. Karkabounas, K. Charalabopoulos, A. Badeka, S. K. Hadjikakou. J. Enzyme Inhib. Med. Chem. 4, 592 (2011).10.3109/14756366.2010.529807Search in Google Scholar PubMed

[29] M. P. Chrysouli, C. N. Banti, N. Kourkoumelis, N. Panayiotou, G. S. Markopoulos, A. J. Tasiopoulos, S. K. Hadjikakou. J. Inorg. Biochem. 179, 107 (2018).10.1016/j.jinorgbio.2017.11.004Search in Google Scholar PubMed

[30] F. Radenkovic, O. Holland, J. J. Vanderlelie, A.V. Perkins. Biochem. Pharmacol. 146, 42 (2017).10.1016/j.bcp.2017.09.009Search in Google Scholar PubMed

[31] D. B. Shpakovsky, C. N. Banti, E. M. Mukhatova, Y. A. Gracheva, V. P. Osipova, N. T. Berberova, D. V. Albov, T. A. Antonenko, L. A. Aslanov, E. R. Milaeva, S. K. Hadjikakou. Dalton Trans. 43, 6880 (2014).10.1039/C3DT53469CSearch in Google Scholar

[32] D. B. Shpakovsky, A. A. Shtil, E. V. Kharitonashvili, V. Y. Tyurin, T. A. Antonenko, A. A. Nazarov, V. P. Osipova, N. T. Berberova, L. S. Foteeva, C. Schmidt, I. Ott, E. R. Milaeva. Metallomics 10, 406 (2018).10.1039/C7MT00286FSearch in Google Scholar

[33] E. Denisov (Ed.), Handbook of Antioxidants, CRC Press, Boca Raton (2018).Search in Google Scholar

[34] R. Amorati, A. Baschieri, L. Valgimigli. J. Chem. MS. ID. 6369358 (2017).10.1155/2017/6369358Search in Google Scholar

[35] Z. Huyut, S. Beydemir, I. Gulcin. Biochem. Res. Int. MS. ID. 7616791 (2017).10.1155/2017/7616791Search in Google Scholar PubMed PubMed Central

[36] I. Belostotskaja, E. Burlakova, V. Misin, G. Nikiforov, N. Khrapova, V. Shtol’ko. Chem. Chem. Technol. 7, 457 (2013).10.23939/chcht07.04.457Search in Google Scholar

[37] D. B. Shpakovsky, C. N. Banti, G. Beaulieu-Houle, N. Kourkoumelis, M. Manoli, M. J. Manos, A. J. Tasiopoulos, S. K. Hadjikakou, E. R. Milaeva, K. Charalabopoulos, T. Bakas, I. S. Butler, N. Hadjiliadis. Dalton Trans. 41, 14568 (2012).10.1039/c2dt31527kSearch in Google Scholar PubMed

[38] E. R. Milaeva. Curr. Top. Med. Chem. 11, 2703 (2011).10.2174/156802611798040741Search in Google Scholar PubMed

[39] C. Pellerito, L. Nagy, L. Pellerito, A. Szorcsik. J. Organomet. Chem. 691, 1733 (2006).10.1016/j.jorganchem.2005.12.025Search in Google Scholar

[40] E. R. Milaeva, D. B. Shpakovsky, V. P. Dyadchenko, A. I. Gryzlov, Yu. A. Gracheva, T. A. Antonenko, M. J. Parulava, D. V. Albov, L. A. Aslanov, L. G. Dubova, P. N. Shevtsov, M. E. Neganova, E. F. Shevtsova. Polyhedron 127, 512 (2017).10.1016/j.poly.2016.08.037Search in Google Scholar

[41] R. Apak, K. Guclu, M. Ozyurek, S. E. Karademir. J. Agric. Food Chem. 52, 7970 (2004).10.1021/jf048741xSearch in Google Scholar PubMed

[42] T. A. Antonenko, D. B. Shpakovsky, M. A. Vorobyov, Y. A. Gracheva, E. V. Kharitonashvili, L. G. Dubova, E. F. Shevtsova, V. A. Tafeenko, L. A. Aslanov, A. G. Iksanova, Y. G. Shtyrlin, E. R. Milaeva. Appl. Organomet. Chem. 32, e4381 (2018).10.1002/aoc.4381Search in Google Scholar

[43] W. Brand-Williams, M. E. Cuvelier, C. Berset. Food Sci. Technol. 28, 25 (1995).10.1016/S0023-6438(95)80008-5Search in Google Scholar

[44] I. Kubo, N. Masuoka, T. J. Ha, K. Tsujimoto. Food Chem. 99, 555 (2006).10.1016/j.foodchem.2005.08.023Search in Google Scholar

[45] Y. C. Chung, S. J. Chen, C. K. Hsu, C. T. Chang, S. T. Chou. Food Chem. 91, 419 (2005).10.1016/j.foodchem.2004.06.022Search in Google Scholar

[46] I. Gulcin, Z. Huyut, M. Elmastas, H. Y. Aboul-Enein. Arab. J. Chem. 3, 43 (2010).10.1016/j.arabjc.2009.12.008Search in Google Scholar

[47] M. E. Neganova, V. A. Blik, S. G. Klochkov, N. E. Chepurnova, E. F. Shevtsova. Neurochem. J. 5, 208 (2011).10.1134/S1819712411030056Search in Google Scholar

[48] V. S. Petrosyan, Y. A. Gracheva, V. Y. Tyurin, E. V. Grigor’ev, E. R. Milaeva, L. Pellerito. Russ. J. Org. Chem. 39, 353 (2003).10.1023/A:1025585513286Search in Google Scholar

[49] S.-H. Lee, X. W. Meng, K. S. Flatten, D. A. Loegering, S. H. Kaufmann. Cell Death Differ. 20, 64 (2012).10.1038/cdd.2012.93Search in Google Scholar PubMed PubMed Central

[50] A. Dziedzic, R. Kubina, A. Kabała-Dzik, M. Tanasiewicz. Evid. Based Complement Alternat Med. MS. ID. 6793456 (2017).10.1155/2017/6793456Search in Google Scholar PubMed PubMed Central

[51] E. R. Milaeva, D. B. Shpakovsky, Yu. A. Gracheva, T. A. Antonenko, D. I. Osolodkin, V. A. Palyulin, P. N. Shevtsov, M. E. Neganova, D. V. Vinogradova, E. F. Shevtsova. J. Organomet. Chem. 782, 96 (2015).10.1016/j.jorganchem.2014.12.013Search in Google Scholar

Published Online: 2020-05-06
Published in Print: 2020-09-25

©2020 IUPAC & De Gruyter. This work is licensed under a Creative Commons Attribution-NonCommercial-NoDerivatives 4.0 International License. For more information, please visit: http://creativecommons.org/licenses/by-nc-nd/4.0/

Downloaded on 3.6.2024 from https://www.degruyter.com/document/doi/10.1515/pac-2019-1209/html
Scroll to top button