Next Article in Journal
How Do Overseas Economic and Trade Cooperation Zones along the Belt and Road Affect the Economic Growth of Host Countries?
Next Article in Special Issue
Potential of Ornamental Trees to Remediate Trace Metal Contaminated Soils for Environmental Safety and Urban Green Space Development
Previous Article in Journal
Assessing the Feasibility of Rooftop Rainwater Harvesting for Food Production in Northwestern Arizona on the Hualapai Indian Reservation
Previous Article in Special Issue
Enhancement of Wheat Growth by UV-Mutagenesis of Potential Chromium Tolerant Bacillus sp. Strain SR-2-1/1
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Genotoxicity of Synthetic Food Colors on Nitrogen-Fixing Bacteria in Agricultural Lands Irrigated with Wastewater of Corresponding Industries

1
Department of Botany, Government College University Lahore, Lahore 54000, Pakistan
2
Department of Environmental Sciences, University of Veterinary and Animal Sciences, Lahore 54000, Pakistan
3
Department of Zoology, Government College University Lahore, Lahore 54000, Pakistan
4
Department of Botany, University of Sargodha, Sargodha 40100, Pakistan
5
Department of Plant Biology and Soil Science, University of Vigo, 36310 Vigo, Spain
6
Department of Statistics, Government College University Lahore, Lahore 54000, Pakistan
7
Department of Environmental Engineering and Management, Chaoyang University of Technology, Taichung 413310, Taiwan
*
Authors to whom correspondence should be addressed.
Sustainability 2023, 15(4), 2897; https://doi.org/10.3390/su15042897
Submission received: 18 October 2022 / Revised: 19 January 2023 / Accepted: 27 January 2023 / Published: 6 February 2023

Abstract

:
Food colors are considered the most important component of foodstuff for enhancing the aesthetic appeal of the products. The rapid increase in population raised the demand for food materials, while wastewater from as-related processing industries is used for irrigation. This study was conducted to examine the genotoxicity of industrial wastewater on the plant growth-promoting rhizobacteria (PGPR). Three predominantly used synthetic food colors, including Azorubine E-122, Tartrazine E-102 and Allura Red AC E-129, were used during this project. Rhizobacteria were isolated from agricultural soils and treated with various concentrations of Azorubine E-122, Tartrazine E-102 and Allura red E-129 for a 24 and 48 h duration. DNA extraction and quantification were performed through a modified CTAB method, spectrophotometry and agarose gel electrophoresis. A comet assay was used to check DNA damage. According to the results, all the food colors had caused significant damage to DNA depending upon the concentration and exposure time. The extent of DNA damage caused by Azorubine E-122 was relatively greater compared with the other colors, so the fragmentation rate of 86% and 92% was obtained at 1.25% concentration for 24 and 48 h, respectively. The current results have revealed the devastation capacity of food colors by accelerating distortion risk to soil micro-flora, hence the fertility of the soil.

1. Introduction

Food colors are an important sensory property in the food production and beverage industries. The global synthetic dye market in 2022 was forecasted to be USD 15.46 billion. Population growth and increasing demand for food products have exerted pressure on food processing industries for the past several decades [1]. Synthetic food colors are widely used in all food products to sustain the sensory characteristics of the food products, which vanish in processing, and also in order to maintain the preferred appearance [2]. Natural food colors are less stable and also are not toxic in nature compared with synthetic food colors [3]. Most of the synthetic colorants are azo compounds in nature, which have been added to various edible items, imparting the latest color and also making them palatable and attractive [4]. The most selling and daily need-based products, such as colas or carbonated drinks, would not be brown without color additives, and similarly, the margarine would not be yellow, and mint ice cream also would not be green. In candy manufacturing, synthetic food colors are more frequently used as compared to natural food colors. Synthetic food colors are considered to be the most important part of processed food. They are usually water-soluble chemical substances and are directly used in the items without further processing. The coloring ability of synthetic dyes is because of the azo group [5]. Azo dyes are commonly synthetic aromatic compounds, having a functional azo group (N꞊N) [6]. Most food dyes (azo dyes) are hydrophilic in nature [7]. In Pakistan, there are eighteen synthetic and five natural permitted food colorants [8]. The United States permitted only seven food colors, including Fast Red (which was forbidden in India), whereas the European Union permitted sixteen synthetic food colors. Iran and Australia legalized thirteen synthetic food colors [9]. Synthetic food colors have a potential effect on the health of living organisms. Most synthetic dyes used in conventional toxicity experiments showed a high rate of toxic effects. The majority of food-borne diseases resulted from the intake of prohibited textile colors [10,11]. They cause DNA damage, skin diseases, cancer of gastrointestinal tracts, colon cancer, and disorders of the liver, kidney and abdomen. Synthetic food colors have explicit genotoxicity and cytotoxicity impacts on humans, such as the impacts of sunset yellow and tartrazine on DNA, by increasing the level of DNA damage and the percentage of apoptosis. The adverse effects of synthetic food colors have also been studied on reactive oxygen species (ROS) [12]. The food color Allura red E-129 (azo dyes) and other vigorous synthetic food colors are the major reason for gut, oral, cardiovascular and neurodegenerative disorders [13]. Because these azo dyes reduce the aromatic amines by the intestinal micro-flora, that is why they cause different carcinogenic diseases in the human body. Epidemiologic research on synthetic food colors has been conducted to identify the neurobehavioral outcomes in both humans (children) and animals. In another finding, azo dye (synthetic food color) azorubine/carmoisine was used as an investigating component; in the results, it showed a high rate of toxicology in animal body weight, hematology, organ weights and biochemistry. It was also reported that the expression of BCL-x and PARP genes was intensively increased, whereas the expression of the p53 gene declined in the mouse liver tissues treated with carmoisine. Consequently, carmoisine was reported to cause renal failure, hepatotoxicity and a suspected reason for liver oncogenesis [14].
Azo dye, such as Red 3 (azorubine/carmoisine), was reported to cause cancer in animals, and in addition to Red 40, Yellow 5 and Yellow 6 have been found to be contaminated with benzidine and other carcinogenic compounds. Whereas Blue 1, Red 40, Yellow 5 and Yellow 6 also resulted in causing hypersensitivity reactions [15]. While having adverse effects on living organisms, synthetic dyes are widely used in textile, food, and pharmaceutical industries and cause significant negative effects on the environment and human health when discharged without prior treatment. The untreated effluents released by the industries contain a diverse range of organic pollutants [16]. Industries that release their wastewater without degrading the free azo dyes in it can cause serious ecotoxicological threats as well as toxic effects on living organisms. Farmers in developing countries used untreated industrial wastewater as a primary source of irrigation, which had a negative impact on crop germination rates and soil quality and ended in prompting genotoxicity in nitrogen-fixing bacteria by causing a reduction in the soil fertility when it reached the soil horizons [17,18]. Several degradation techniques or methods of dye compounds have been introduced to overcome the toxicity concern of contaminants discharged from different industrial activities. Among them, photocatalytic technologies are revealed to be promising candidates because they play an important role in breaking down different hazardous compounds into innocuous products [19,20,21]. Synthetic food dyes are abundantly present in industrial effluents, and when they are discharged into the environment (water and soil), they reduce light penetration and inhibit the performance and growth of algae, aquatic plants and agricultural crops. Synthetic dyes ingested by living organisms can be metabolized in their bodies into toxic intermediates, causing adverse effects in their bodies [22]. Plant growth-promoting rhizobacteria (PGPR) are the rhizosphere bacteria that have the power to raise the availability of nitrogen concentration in the soil atmosphere [23]. Bacterial oxidoreductase enzymes play an important role in the degradation of synthetic food colors. Bacterial metabolism permits them to exploit complex xenobiotic compounds of the dyestuff as a substrate. During this procedure, they are degraded to minute complex metabolites [24]. Rhizobacteria are more in the rhizosphere of plants than in the adjacent soil, and this perfusion depends on the crop species [25]. The process of nitrogen fixation is among the most important biological processes. Biological nitrogen fixation assists in improving soil fertility and crop production [26]. Rhizobacteria are useful microorganisms and can apply as bioinoculants and also for studying the nitrogen fixation process. They are sensitive to acidic pH, high salts and temperature [27]. It has beneficial effects on crop growth and yields through the biosynthesis of biologically active substances, stimulation of rhizospheric microbes and by producing phytopathogenic inhibitors [28]. Rhizobacteria have the capacity to continuously grow and quickly fix large amounts of nitrogen and are able to alter atmospheric nitrogen into ammonia, which is consumed by plants. Moreover, nitrogen plays an explicit role in plant growth as it is a fundamental plant nutrient and is broadly used as a nitrogen fertilizer to improve the yield of agriculturally important crops. In addition to nitrogen fertilizer, other chemical fertilizers also play an important role in increasing the economic yield of crops [29]. In addition, the efficiency of PGPR has been found to decrease with increased nitrogen levels [30]. The finest grouping was verified with NH₄Cl (Ammonium chloride) at 0.1 g/L, whereas the action of copper in rhizobacteria was found to be toxic, even in small amounts [31]. The rhizobacteria may suffer due to extreme amounts of nitrates and the acidic environment created because of chemical fertilizers and toxic industrial effluents [32]. It is investigated that the result of herbicide 2, 4–D and its products, p-chlorophenoxy-acetic acid and p-chlorophenol were exploited by Azotobacter croococcum as a carbon source, which eventually stimulates the nitrogenase enzyme. The organophosphorus insecticides profenofos and chlorpyrifos had contrary effects on the aerobic nitrogen fixers and lessened their nitrogen-fixing ability. The intake and absorption of some toxic heavy metals by aquatic flora and soil can cause detrimental health issues in animals and, ultimately, in humans via the food chain [33]. In this study, three synthetic food colors, such as Azorubine E-122, Tartrazine E-102 and Allura red AC E-129, were selected to study the rate of genotoxicity on the genome of PGPR. The food color Azorubine E-122, commonly known as carmoisine or Cl food Red 3 [34], has a structural formula in which two subunits of naphthalene were present (C₂₀H₁₂N₂Na₂O₇S₂). The food color Tartrazine E-102 was a synthetic lemon-yellow azo dye, also known as “Cl food yellow 4” with a chemical formula (C₁₆H₉N₄Na₃O₉S₂) [35], and the synthetic food color Allura red AC E-129, known as “CL Food Red 17”, was a synthetic red azo color, having the chemical formula C₁₈H₁₄N₂Na₂O₈S₂, as shown in Figure 1 [35]. Although it was commonly delivered as a sodium salt, it can normally be used as potassium and calcium salts.
Subsequently, this study revealed how synthetic food colors, which were released as industrial effluent by food processing industries, cause DNA damage in rhizobacteria. Consequently, the damage of PGPR (rhizobacteria) results in inhibiting the biological process of nitrogen fixation and also the performance of rhizobacteria in the production and regulation of compounds and stress hormonal status. As a result, via entering the food chain, these adverse chemicals cause various serious disorders in the human body.

2. Materials and Methods

2.1. Sample Collection and Preparation

Soil samples were collected by using sterilized equipment from contamination-free fields more than 60 km away from urban settlements and industrial setups, along with rhizosphere soil surface and subsurface about 10–15 cm in depth, and along with the roots of the plant in a sterilized polybag. For isolation of bacteria from roots, 3–5 g of fresh roots were washed and, for surface sterilization, dipped in 5% NaOCl for one minute. After sterilization with double-distilled water was performed three times, the root sample was ground by using a sterilized mortar and pestle. For the preparation of diluent/peptone salt solution, 0.5 g of peptone was taken in a large glass bottle along with 4.25 g NaCl and 500 mL distilled water, and then the solution was well-shaken and autoclaved. A measurement of 225 mL of peptone solution was added into the stomacher bag and mixed into the solution to homogenize [36].
Bacterial growth media was prepared by following the protocol designed in [37] with some modifications (2.5 g of tryptone, 1.25 g of yeast extract and 2.5 g of NaCl added into 250 mL of distilled water). A total of 10 mL of LB broth (Luria Bertani) was added to each screw-capped test tube. One mL of prepared samples from both soil and root was in each test tube. After this, the enrichment broth was incubated at 37 °C for 24 h. Synthetic food colorants, such as Azorubine E-122 and Tartrazine E-102, were purchased from Essence.pk Company (important colors), whereas Allura red AC was bought from a local commercial brand (Italiano).

2.2. Morphological Identification of Bacterial Isolates

Colony morphology, shape, size, color and growth pattern were noted after 24 h of growth on LB (Luria Bertani) medium, as previously described [38]. The size of cells was observed by light microscopy, and the gram staining was completed by following the protocol in [39]. Several steps of biochemical tests were also performed to characterize bacteria by following the procedure in [40], such as the KOH solubility test, performed as previously described [41].

2.3. Application of Dilution of Synthetic Food Colors

Five dilutions were made from each synthetic food color (Azorubine E-122, Tartrazine E-102 and Allura red AC E-129). Precisely 0.25, 0.50, 0.75, 1.0 and 1.25 g [1] of Azorubine E-122, Tartrazine E-102 and Allura red E-129 were taken in each jar along with 100 mL of dH2O and stirred to homogeneously dissolve the color. After dilution preparation, two mL of each diluted synthetic food color was applied to bacterial cultures. The experiment was conducted in two setups for each synthetic color at 24 and 48 h. Both setups contained the same concentration of dilutions, 0.25%, 0.50%, 0.75%, 1.0% and 1.25%, along with the control. After the application of dilutions, the bacterial cultures were incubated at 37 ℃ for 24 and 48 h.

2.4. Genomic DNA Extraction

The genomic DNA of rhizobacteria was extracted by the CTAB extraction method [42] with some modifications.

2.5. Quantification of Genomic DNA

Quantification of rhizobacterial DNA was measured after being treated with Azorubine E-122, Tartrazine E-102 and Allura red AC E-129 for 24 and 48 h by a Single Beam Scanning UV-VIS Spectrophotometer-2373 [43]. Agarose gel electrophoresis was used to quantify the presence of genomic DNA [44].

2.6. Comet Assay

The comet assay was accomplished according to the described protocol [45], with some modifications. Precisely 1% of NMPA (Normal Melting Point Agarose) was prepared in distilled water. After this step, the solution was placed into a water bath to maintain the temperature at 65 °C. The slides were washed with 70% methanol to make them contamination free. For labeling, paper tape was used on one side of the slide. In the next step, the slides were dipped in NMPA and placed on a clean surface overnight to dry completely. A total of 50 µL of bacterial genomic DNA sample were dissolved into 50 µL of 1% LMPA (Low Melting Point Agarose prepared in PBS).
The samples were mixed homogeneously by pipetting them up and down, making sure to catch the speed at this stage because the agarose can be quickly solidified. After this step, we immediately added 50 µL of the sample on the slide and put a coverslip on it, and then placed the slides at 4 °C for 30 min. A lysis buffer cannot be applied if DNA extraction has been previously performed (which is known as an indirect method of the comet assay).
In the next step, single-gel electrophoresis was carried out to run the experimental slides. The tank was placed in the freezer before running the slides for almost 1 h, then chilled alkaline electrophoresis solution was added to the chilled tank (place the tank on ice to try to maintain the temperature because the most important thing in this procedure is temperature management). Then, the slides were placed on the tank, and the slides were run at 25 V, 300 mA for 30 min at 4 °C. The excess alkaline electrophoresis solution was removed with the help of blotting paper. The slides were washed with a neutralizing buffer for almost 10 min 3 times. The slides were placed in 70% ethanol for almost 5 min at room temperature.
The next step was the staining of the slides. To make sure the slides were completely dry, we placed a drop of Ethidium bromide (CAS No. 1239-45-8) on the gel attached to the slide and placed a cover slip on top of it.

Quantification and Visualization of DNA Damage

The slides were placed under a fluorescence microscope (Olympus BX50, Hamburg, Germany) at 10× magnification to take pictures of the stained slide. For image analysis using the Comet software (casplab, comet score, Comet Assay IV), the results of the comet assay were calculated by using Comet scoring software [46].

2.7. Statistical Analyses

A one-way ANOVA was carried out for the statistical examination of data with Microsoft Excel for analyzing the significant difference of DNA tails between three of the synthetic food colors.

3. Results

3.1. Quantification of DNA after Treatment with Azorubine E-122, Tartrazine E-102 and Allura Red AC E-129

Quantification of DNA samples after 24 and 48 h was performed by using a spectrophotometer. Readings were recorded at both A260 nm and A280 nm wavelengths to analyze the purity index of DNA in experimental samples. An absorbance of one unit at 260 nm corresponds to 50 μg genomic DNA per mL (A260 = 1 for 50 ug/mL; based on standard 1 cm path length), whereas the ratio of A₂₆₀/A₂₈₀ nm indicates sample purity as most of the results were above ≥1.8, which is a standard ratio for purified DNA as shown in Table 1.

3.2. Quantification of DNA through Agarose Gel Electrophoresis

Agarose gel electrophoresis was used to quantify the presence of genomic DNA of rhizobacteria by using 1% Agarose gel, in which 90 voltage is applied for 25 min to quantify the genomic DNA against 1 KB DNA ladder as in Figure 2. The presence of bands near the wells indicates the presence of genomic DNA.

3.3. Image Analysis by Using Comet Scoring Software

Comet slides (Azorubine E-122, Tartrazine E-102 and Allura red AC E-129) at both exposure time periods 24 and 48 h were visualized under a fluorescence microscope (Olympus BX50, Hamburg, Germany) at 10× magnification, and after visualization, images of comets were amplified by using the comet score software. In addition to measuring the amount of DNA damage, the ratio of fluorescence intensity was used inside the comet tail and the fluorescence intensity of the comet head. After putting the images in the software, other parameters, such as comet length, comet head radius, tail length and tail movement, were also measured. In Figure 3, the fragmented tail of comets showed adverse effects of synthetic food colors on the genome of rhizobacteria. The control of each dilution showed negative results. The frequent rate of fragmentation in every image indicates the rate of genotoxicity which was induced by three of the studied synthetic food colors.

3.4. Implementation of Comet Measurement to Evaluate the Rate of Genotoxicity on DNA Caused by Synthetic Food Colors

Single-cell gel electrophoresis or comet assay was used to measure the detection of DNA damage in the DNA cells of rhizobacteria. The total genotoxicity produced by Azorubine E-122, Tartrazine E-102 and Allura red AC E-129 was found to be dependent on concentration and exposure period. For measuring the significant units of a comet (comet length, head and tail length and movement), the comet score software was used. However, the result indicates that each of the three synthetic food colors causes significant genotoxicity. The most DNA damage was recorded in Azorubine E-122 rather than Tartrazine E-102 and Allura red AC E-129, as shown in Table 2.

3.5. Evaluation of DNA Damage through Statistical Analyses

Statistical analyses were used to evaluate significant differences in DNA tails among the three studied synthetic food colors at both exposure time periods (Azorubine E-122, Tartrazine E-102 and Allura red AC E-129). A one-way ANOVA was applied to find the rate of significance between the three studied synthetic food colors. In addition, the results showed that there is no significant difference between the DNA damage, which was caused by the application of synthetic colors (as the p-value of F statistics in both setups, 24 and 48 h, is higher than 0.05), as shown in Table 3. This means all three colors play a significant role in damaging the DNA of plant growth-promoting rhizobacteria (PGPR). However, the summary of the statistical analyses, in which the sum and average were calculated between two setups or groups (24 and 48 h), showed that most of the DNA damage resulted from Azorubine E-122 as compared with Tartrazine E-102 and Allura red AC E-129, as shown in Table 4.

4. Discussion

This study was designed to investigate the factor of genotoxicity on plant growth-promoting rhizobacteria (PGPR) caused by synthetic food dyes (azo dyes), which were released as industrial effluents. The extent of employing rhizobacteria in research experiments as microbial inoculants is because of their growth-promoting ability and also their nitrogen-fixating capabilities, such as synthesizing indole compounds and producing hydrolytic enzymes [47]. It was found that all three colors show a significant rate of DNA damage on the genome of rhizobacteria, but among them, an exceptional rate of genotoxicity was shown in Azorubine E-122 as compared with Tartrazine E-102 and Allura red AC E-129 because the recorded ratio of DNA tail in Azorubine E-122 was 86% and 93% at both 24 and 48 h. Tartrazine E-102 and Allura red AC E-129 also showed significant DNA damage, such as 65% and 79% and 73% and 91% at both 24 and 48 h setups. The rate of DNA damage in both Tartrazine and Allura red AC was reported less as compared with Azorubine. Thereby, it was recorded that all of the dilutions had caused genotoxicity, but higher concentrations of dilution caused some major injuries or toxicity. A similar pattern was previously observed in which genotoxicity impact was studied on meristematic cells root tips of Allium cepa, induced by food coloring dyes. The highest frequency of genotoxicity was reported at various stages of the cell cycle [48]. In another research work, Escherichia coli, Staphylococcus albus and Saccharomyces cerevisiae were used to study the toxicity rate of synthetic food colors, in which the rate of toxicity or damage was directly proportional to the rate of concentration applied to the test organism [49]. The purpose of the current study is to determine the genotoxicity of synthetic food colors on PGPR, which contributes a prominent role in soil fertility and increasing plant growth and plays an important role in nitrogen fixation [50]. However, in a previous research work, an azo dye (used as a commercial dye) known as CI Disperse Blue 291 (DB291) was used to investigate the rate of genotoxicity of DB291. Primary DNA damage in blood, kidney, and liver cells, BAX, BCL2, SMAD4 and TNFα gene expression in leukocytes were recorded. Animals that are treated with 50 mg/kg bodyweight had shown an amplified frequency of micronucleated polychromatic erythrocytes (MNPCEs), whereas, on the other hand, neither genetic alteration, primary DNA damage nor aberrations in gene expression were reported [51]. In addition, according to the present results, the selected synthetic dyes (E-122, E-102 and E-124) also showed a significant rate of damage in the genome of rhizobacteria, similar to the primary damage that had been recorded by CI disperse blue. Industrial effluents are commonly discharged in nearby wetlands, crop fields and water bodies which, as a result, cause low fertility, dermal disease, low yields, boost insect attack on crops and increase the salt content and salinity of the soil [52]. Increased soil salinity resulted in adversely affecting the plant nutrient balance, protein synthesis, energy metabolism and photosynthesis and disturbing the whole biological process of nitrogen fixation overall [53]. Industrial effluents contain several carcinogenic heavy metals, such as Mercury (Hg), Lead (Pb), Cadmium (Cd), Chromium (Cr), Nickel (Ni) and Thallium (TI) [54]. Different food additives, which were used as food colors, include heavy metals, such as Lead (Pb), Mercury (Hg) and arsenic [55]. Synthetic dyes are azo chemicals in nature, and the oxidative end products have carcinogenic and mutagenic effects, which produce modifications of DNA [52]. Scientific work on zebrafish embryos was used as an indicator to study the effect of food colorant carmoisine, which showed a significant decrease in height and eye diameter of the embryos, a rise in free oxygen radicals in apoptotic cells and lipid accumulation and caused serious malformations [53]. Although in collaboration with the above finding, the present results showed that most of the DNA damage was caused at higher concentrations of dilution, such as at 1.0% and 1.25%, and at greater exposure time periods. The numeric and structural variations in chromosomes have become a key factor in appraising the genotoxicity of pollutants, heavy metals, water samples and food chemicals [56].
Despite the fact that the present study also positively relates to a research finding in which different bacterial isolates obtained from the soil are aggregated by means of industrial wastewater (secondary source of aggregation), the results showed adsorption of Cd and Ni at different concentrations of single- and bi-metal systems and indicated that these heavy metals not only affect the soil flora but also affect the agricultural system [57]. Similar to the presence of benzidine, 3,3′-dimethoxybenzidineand 3, or 3′-dimethylbenzidine were reported in the industrial wastewater from where the soil samples were collected. These chemicals are classified as carcinogens of category [58]. The presence of these carcinogenic chemicals in soil inhibits the activity of PGPR and causes highly toxic effects on the biological process of nitrogen fixation. Notable results of the comet assay proved that synthetic food colors, which were discharged as industrial effluents, play an explicit role in destroying the soil flora, which in turn leads to distortion of the food chain. Several food additives, such as sodium benzoate and sodium sulfite, were found in chromosomal aberrations, colon cancer, etc. This may be due to the clastogenic activity of chemicals on DNA [59]. Moreover, some studies hold the opposite opinion that azorubine/carmoisine is not a standard genotoxicity assay [60], and in contrast to the present research work, it was found that not only Azorubine E-122 but also Tartrazine E-102 and Allura red AC E-129 cause significant DNA damage at all dilutions other than the control. However, in another research finding, it was reported that synthetic food colors cause health problems, such as cancers, reduced hemoglobin concentrations, mutations and allergic reactions. Several confectioners also contain unidentified colors, such as textile dyes [2]. In addition, different industrial wastewater contains synthetic dyes or food colors in which almost 70% of azo dyes contain one or more azo bonds (R₁–N–N–R₂) in their chemical structure [61]. Some synthetic food colors are reported to contain carcinogenic chemicals, such as β-naphthylamine (which is also known as 2-naphthylamine), carrying the amino group at position two [62,63]. The chemical structure of Azorubine, such as disodium 4-hydroxy-2-[(E)-(4-sulphonato-1-naphthy)diazenyl]naphthalene-1-sulfonate, contain β-naphthylamine in their chemical structure. A report was issued by the Department of Health and Senior Services in New Jersey [64] in which it was stated that 2-naphthylamine is a carcinogenic chemical that is basically used in the manufacturing of different azo dyes (synthetic food colors). According to the report, 2-naphthylamine is no longer produced for commercial use and is only available for research purposes. The presence of β-naphthylamine or 2-naphthylamine in the chemical composition of Azorubine or carmoisine makes it carcinogenic, which is considered to be the most hazardous environmental pollutant [65]. In support of this present work, three of the studied synthetic food colors showed significant DNA damage, whereas Azorubine E-122 showed slightly greater DNA damage as compared with Tartrazine E-102 and Allura red AC E-129.

5. Conclusions

Synthetic food colors, such as Azorubine E-122, Tartrazine E-102 and Allura red E-129, induce significant damage to DNA in rhizobacteria. The Azorubine E-122 has shown explicit genotoxicity in the plant growth-promoting rhizobacteria (PGPR). Moreover, this compound (Azorubine E-122) causes higher fragmentation in DNA tails, where the fragmentation rate is 86% at 1.25% concentration for 24 h exposure time period, and at 48 h, the fragmentation rate is 93% at 1.25% concentration of synthetic food color. Hence, these synthetic food colors are genotoxic chemical compounds. They are not only harmful to humankind but also to the environment in which they are released. The usage of synthetic coloring dyes should be prohibited, and natural food colors should be promoted at both industrial and commercial levels.

Author Contributions

Conceptualization, M.U.F.A. and H.-H.Y.; methodology, M.L.; software, S.M, U.H. and A.A.S.; validation, S.A., A.H. and M.T.; formal analysis, Z.I.K.; investigation, M.I.H.; resources, B. and M.N.A.; data curation, M.S.C. and A.E.; writing—original draft preparation, A.J.; writing—review and editing, A.J.; visualization, S.A.; supervision, S.M.; project administration, M.U.F.A.; funding acquisition, H.-H.Y. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

We gratefully acknowledge the support, assistance and engagement of the collaborators from the Department of Environmental Engineering and Management, Chaoyang University of Technology, Taiwan, and the Department of Botany, Government College University Lahore, Pakistan.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Khan, I.S.; Ali, M.N.; Hamid, R.; Ganie, S.A. Genotoxic effect of two commonly used food dyes metanil yellow and carmoisine using Allium cepa L. as indicator. Toxicology 2020, 7, 370–375. [Google Scholar] [CrossRef]
  2. Subhashish, D.; Nagababu, B.H. Applications of food color and bio-preservatives in the food and its effect on the human health. Food Chem. 2022, 1, 100019. [Google Scholar]
  3. Neves, M.I.L.; Silva, E.K.; Meireles, M.A.A. Natural blue food colorants: Consumer acceptance, current alternatives, trends, challenges, and future strategies. Food Sci. Technol. 2021, 112, 163173. [Google Scholar]
  4. Forgacs, E.; Cserháti, T.; Oros, G. Removal of synthetic dyes from wastewaters: A review. Environ. Int. 2004, 30, 953–971. [Google Scholar] [CrossRef] [PubMed]
  5. Benkhaya, S.; M’rabet, S.; El Harfi, A. Classifications, properties, recent synthesis and applications of azo dyes. Heliyon 2020, 6, e03271. [Google Scholar] [CrossRef] [PubMed]
  6. AbuKhader, M.; Dhanalekshmi, U.M.; Nazmi, A. Identification and Prevalence of Food Colors in Candies Commonly Consumed by Children in Muscat, Oman. IJNPND 2021, 11, 128. [Google Scholar] [CrossRef]
  7. Hallagan, J.B.; Allen, D.C.; Borzelleca, J.F. The safety and regulatory status of food, drug and cosmetics colour additives exempt from certification. FCT 1995, 33, 515–528. [Google Scholar] [CrossRef]
  8. Punjab Pure Food Regulation. Available online: https://foodscienceuniverse.com/wp-content/uploads/2020/10/Punjab-Pure-Food-Regulations-2018.pdf (accessed on 20 November 2022).
  9. European Parliament and Council Directive 94/36/EC. Off. J. Eur. Communities 1994, 10, 13–29.
  10. Kroger, M.; Meister, K.; Kava, R. Low-calorie Sweeteners and Other Sugar Substitutes: A Review of the Safety Issues. IFT 2006, 5, 27–47. [Google Scholar] [CrossRef]
  11. Lhotta, K.; Höfle, G.; Gasser, R.; Finkenstedt, G. Hypokalemia, hyperreninemia and osteoporosis in a patient ingesting large amounts of cider vinegar. Nephron 1998, 80, 242–243. [Google Scholar] [CrossRef]
  12. El-Borm, H.; Badawy, G.; Hassab El-Nabi, S.; El-Sherif, W.; Atallah, M. Toxicity of sunset yellow fcf and tartrazine dyes on dna and cell cycle of liver and kidneys of the chick embryo: The alleviative effects of curcumin. Egypt. J. Zool. 2020, 74, 43–55. [Google Scholar] [CrossRef]
  13. John, A.; Yang, H.-H.; Muhammad, S.; Khan, Z.I.; Yu, H.; Luqman, M.; Tofail, M.; Hussain, M.I.; Awan, M.U.F. Cross Talk between Synthetic Food Colors (Azo Dyes), Oral Flora, and Cardiovascular Disorders. Appl. Sci. 2022, 12, 7084. [Google Scholar] [CrossRef]
  14. Reza, M.S.A.; Hasan, M.M.; Kamruzzaman, M.; Hossain, M.I.; Zubair, M.A.; Bari, L.; Abedin, M.Z.; Reza, M.A.; Khalid-Bin-Ferdaus, K.M.; Haque, K.M.F.; et al. Study of a common azo food dye in mice model: Toxicity reports and its relation to carcinogenicity. Food Sci. Nutr. 2019, 7, 667–677. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Kobylewski, S.; Jacobson, M.F. Toxicology of food dyes. Int. J. Occup. Environ. Health 2012, 18, 220–246. [Google Scholar] [CrossRef] [PubMed]
  16. Oyeniran, D.O.; Sogbanmu, T.O.; Adesalu, T.A. Antibiotics, algal evaluations and subacute effects of abattoir wastewater on liver function enzymes, genetic and haematologic biomarkers in the freshwater fish, Clarias gariepinus. Ecotoxicol. Environ. Saf. 2021, 212, 111982. [Google Scholar] [CrossRef]
  17. Jiku, M.A.S.; Singha, A.; Faruquee, M.; Rahaman, M.A.; Alam, M.A.; Ehsanullah, M. Toxic wastewater status for irrigation usage at Gazipur and Savar industrial vicinity of Bangladesh. Acta Ecol. Sin. 2021, 41, 358–364. [Google Scholar] [CrossRef]
  18. Ao, S.; Zayed, T. Impact of sewer overflow on public health: A comprehensive scientometric analysis and systematic review. Environ. Res. 2022, 203, 111609. [Google Scholar]
  19. Padervand, M.; Mazloum, M.; Bargahi, A.; Arsalani, N. CQDs/BiOCl Photocatalysts for the Efficient Treatment of Congo Red Aqueous Solution under Visible Light. J. Nanostruct. 2021, 11, 790–801. [Google Scholar] [CrossRef]
  20. Padervand, M.; Rhimi, B.; Wang, C. One-pot synthesis of novel ternary Fe3N/Fe2O3/C3N4 photocatalyst for efficient removal of rhodamine B and CO2 reduction. J. Alloys Compd. 2021, 852, 156955. [Google Scholar] [CrossRef]
  21. Padervand, M.; Nasiri, F.; Hajiahmadi, S.; Bargahi, A.; Esmaeili, S.; Amini, M.; Nami, R.K.; Shahsavari, Z.; Karima, S. Ag@Ag2MoO4 decorated polyoxomolybdate/C3N4 nanostructures as highly efficient photocatalysts for the wastewater treatment and cancer cells killing under visible light. Inorg. Chem. Commun. 2022, 141, 109500. [Google Scholar] [CrossRef]
  22. Al-Tohamy, R.; Ali, S.S.; Li, F.; Okasha, K.M.; Mahmoud, Y.A.G.; Elsamahy, T.; Jiao, H.; Fu, Y.; Sun, J. A critical review on the treatment of dye-containing wastewater: Ecotoxicological and health concerns of textile dyes and possible remediation approaches for environmental safety. Ecotoxicol. Environ. Saf. 2022, 231, 113160. [Google Scholar] [CrossRef]
  23. Basu, A.; Prasad, P.; Das, S.N.; Kalam, S.; Sayyed, R.Z.; Reddy, M.S.; El Enshasy, H. Plant Growth Promoting Rhizobacteria (PGPR) as Green Bioinoculants: Recent Developments, Constraints, and Prospects. Sustainability 2021, 13, 1140. [Google Scholar] [CrossRef]
  24. Jamee, R.; Siddique, R. Biodegradation of Synthetic Dyes of Textile Effluent by Microorganisms: An Environmentally and Economically Sustainable Approach. Eur. J. Microbiol. Immunol. 2019, 9, 114–118. [Google Scholar] [CrossRef]
  25. Hayat, R.; Ali, S.; Amara, U.; Khalid, R.; Ahmed, I. Soil beneficial bacteria and their role in plant growth promotion: A review. Ann. Microbiol. 2010, 60, 579–598. [Google Scholar] [CrossRef]
  26. Wani, S.H.; Kumar, V.; Shriram, V.; Sah, S.K. Phytohormones and their metabolic engineering for abiotic stress tolerance in crop plants. Crop J. 2016, 4, 162–176. [Google Scholar] [CrossRef]
  27. Raffi, M.M.; Charyulu, P.B.B.N. Azospirillum-biofertilizer for sustainable cereal crop production: Current status. In Recent Developments in Applied Microbiology and Biochemistry; Academic Press: Cambridge, MA, USA, 2020; Volume 2, pp. 193–209. [Google Scholar]
  28. Santoyo, G.; Urtis-Flores, C.A.; Loeza-Lara, P.D.; Orozco-Mosqueda, M.; Glick, B.R. Rhizosphere colonization determinants by plant growth-promoting rhizobacteria (PGPR). Biology 2021, 10, 475. [Google Scholar] [CrossRef]
  29. El-Mageed, A.; Taia, A.; El-Mageed, A.; Shimaa, A.; El-Saadony, M.T.; Abdelaziz, S.; Abdou, N.M. Plant growth-promoting rhizobacteria improve growth, morph-physiological responses, water productivity, and yield of rice plants under full and deficit drip irrigation. Rice 2022, 15, 1–15. [Google Scholar] [CrossRef]
  30. Kaloterakis, N.; Van Delden, S.H.; Hartley, S.; De Deyn, G.B. Silicon application and plant growth promoting rhizobacteria consisting of six pure Bacillus species alleviate salinity stress in cucumber (Cucumis sativus L). Sci. Hortic. 2021, 288, 110383. [Google Scholar] [CrossRef]
  31. He, A.; Niu, S.; Yang, D.; Ren, W.; Zhao, L.; Sun, Y.; Meng, L.; Zhao, Q.; Paré, P.W.; Zhang, J. Two PGPR strains from the rhizosphere of Haloxylon ammodendron promoted growth and enhanced drought tolerance of ryegrass. Plant Physiol. Biochem. 2021, 161, 74–85. [Google Scholar] [CrossRef]
  32. Gupta, G.; Parihar, S.S.; Ahirwar, N.K.; Snehi, S.K.; Singh, V. Plant growth promoting rhizobacteria (PGPR): Current and future prospects for development of sustainable agriculture. J. Microb. Biochem. Technol. 2015, 7, 096–102. [Google Scholar]
  33. Yadav, A.N. Biodiversity and bioprospecting of extremophilic microbiomes for agro-environmental sustainability. J. Appl. Biol. 2021, 9, 1–6. [Google Scholar]
  34. Gupta, V.K.; Mittal, A.; Malviya, A.; Mittal, J. Adsorption of carmoisine A from wastewater using waste materials-bottom ash and deoiled soya. J. Colloid Interface Sci. 2009, 335, 24–33. [Google Scholar] [CrossRef]
  35. FAO. Tartrazine. Allura Red AC. Compendium of Food Additive Specifications. Joint FAO/WHO Expert Committee on Food Additives (JECFA), 82nd Meeting 2016. FAO JECFA Monographs 19. FAO/WHO 2016. Available online: http://www.fao.org (accessed on 12 September 2021).
  36. Barragán, B.E.; Costa, C.; Marquez, M.C. Biodegradation of azo dyes by bacteria inoculated on solid media. Dyes Pigm. 2007, 75, 73–81. [Google Scholar] [CrossRef]
  37. Maza, F.; Maldonado, J.; Vásquez-Dean, J.; Mandakovic, D.; Gaete, A.; Cambiazo, V.; González, M. Soil Bacterial Communities from the Chilean Andean Highlands: Taxonomic Composition and Culturability. Front. Bioeng. Biotechnol. 2019, 7, 10. [Google Scholar] [CrossRef] [PubMed]
  38. Somasegaran, P.; Hoben, H.J. Handbook for Rhizobia: Methods in Legume-Rhizobium Technology; Springer: New York, NY, USA, 1994. [Google Scholar]
  39. Vincent, J.M.; Humphrey, B. Taxonomically significant group antigens in Rhizobium. J. Gen. Microbiol. 1970, 63, 379–382. [Google Scholar] [CrossRef] [PubMed]
  40. Bergey, D.H.; Holt, J.G.; Noel, R.K. Bergey’s Manual of Systematic Bacteriology, 9th ed.; Williams & Wilkins: Baltimore, MD, USA, 1994; Volume 1, pp. 1935–2045. [Google Scholar]
  41. Kirsop, B.E.; Doyle, A. Maintenance of Microorganisms and Cultured Cells: A Manual of Laboratory Methods, 2nd ed.; Academic Press: London, UK, 1991. [Google Scholar]
  42. Shu, Y.; Wan-Ting, J.; Ya-Ning, Y.; Yu-Han, F. An Optimized CTAB Method for Genomic DNA Extraction from Freshly-picked Pinnae of Fern, Adiantum capillus-veneris L. Bio-Protocol 2018, 8, e2906. [Google Scholar] [CrossRef] [Green Version]
  43. Nzilibili, S.M.M.; Ekodiyanto, M.K.H.; Hardjanto, P.; Yudianto, A. Concentration and Purity DNA Spectrophotometer: Sodium Monofluorophosphate forensic impended effect. Egypt J. Forensic Sci. 2018, 8, 34. [Google Scholar] [CrossRef]
  44. Lee, P.Y.; Costumbrado, J.; Hsu, C.Y.; Kim, Y.H. Agarose gel electrophoresis for the separation of DNA fragments. J. Vis. Exp. 2012, 62, 3923. [Google Scholar] [CrossRef]
  45. Tice, R.R.; Agurell, E.; Anderson, D.; Burlinson, B.; Hartmann, A.; Kobayashi, H.; Iyamae, Y.; Rojas, E.; Ryu, J.C.; Sasaki, Y.F. Single cell gel/comet assay: Guidelines for in vitro and in vivo genetic toxicology testing. Environ. Mol. Mutagen. 2000, 35, 206–221. [Google Scholar] [CrossRef]
  46. Lovell, D.P. Statistical analysis of comet assay data. In The Comet Assay in Toxicology; Dhawan, A., Anderson, D., Eds.; RSC: Cambridge, UK, 2009; pp. 424–450. [Google Scholar]
  47. Raza, A.; Ejaz, S.; Saleem, M.S.; Hejnak, V.; Ahmad, F.; Ahmed, M.A.A.; Alotaibi, S.S.; El-Shehawi, M.A.; Alsubeie, M.S.; Zuan, A.T.K. Plant growth promoting rhizobacteria improve growth and yield related attributes of chili under low nitrogen availability. PLoS ONE 2021, 16, e0261468. [Google Scholar] [CrossRef]
  48. Sheetal, K.; Halady, P.; Swapna, B. Genotoxicity Induced by Food Coloring Dyes on Meristematic Cells Root Tips of Allium Cepa. Int. J. Trend Sci. Res. Dev. 2019, 3, 116–118. [Google Scholar] [CrossRef]
  49. Sadar, P.; Dande, P.; Kulkarni, N.; Pachori, R. Evaluation of toxicity of synthetic food colors on human normal flora and yeast. Int. J. Health Sci. Res. 2017, 7, 110–114. [Google Scholar]
  50. Gallart, M.; Paungfoo-Lonhienne, C.; Gonzalez, A.; Trueman, S.J. Nitrogen source influences the effect of plant growth-promoting rhizobacteria (Pgpr) on macadamia integrifolia. J. Agron. 2021, 11, 1064. [Google Scholar] [CrossRef]
  51. Fernandes, F.H.; de Aragão Umbuzeiro, G.; Fávero Salvadori, D.M. Genotoxicity of textile dye C.I. Disperse Blue 291 in mouse bone marrow. Mutat. Res. Genet. Toxicol. Environ. Mutagen. 2019, 837, 48–51. [Google Scholar] [CrossRef] [PubMed]
  52. Ma, X.; Pan, J.; Xue, X.; Zhang, J.; Guo, Q. A Bibliometric Review of Plant Growth-Promoting Rhizobacteria in Salt-Affected Soils. Agronomy 2022, 12, 2304. [Google Scholar] [CrossRef]
  53. Chequer, F.M.D.; de Paula Venâncio, V.; Bianchi, M.D.L.P.; Antunes, L.M.G. Genotoxic and mutagenic effects of erythrosine B, a xanthene food dye, on HepG2 cells. Food Chem. Toxicol. 2012, 50, 3447–3451. [Google Scholar] [CrossRef]
  54. Kinuthia, G.K.; Ngure, V.; Beti, D.; Lugalia, R.; Wangila, A.; Kamau, L. Levels of heavy metals in wastewater and soil samples from open drainage channels in Nairobi, Kenya: Community health implication. Sci. Rep. 2020, 10, 8434. [Google Scholar] [CrossRef]
  55. Ogimoto, M.; Uematsu, Y.; Suzuki, K.; Kabashima, J.; Nakazato, M. Shokuhin eiseigaku zasshi. J. Food Hyg. Soc. Jpn. 2009, 50, 256–260. [Google Scholar] [CrossRef]
  56. Kiziltan, T.; Baran, A.; Kankaynar, M.; Şenol, O.; Sulukan, E.; Yildirim, S.; Ceyhun, S.B. Effects of the food colorant carmoisine on zebrafish embryos at a wide range of concentrations. Arch. Toxicol. 2022, 96, 1089–1099. [Google Scholar] [CrossRef]
  57. Zeid, H.A.A.; El-Zayat, M.M.; Abdrabouh, A.E.S. Ecotoxicological impacts of industrial effluents on irrigation water quality, animal health and the role of calcium alginate in effluents treatment. Environ Monit Assess. 2022, 194, 1–25. [Google Scholar] [CrossRef]
  58. Othman, Y.A.; Al-Assaf, A.; Tadros, M.J.; Albalawneh, A. Heavy Metals and Microbes Accumulation in Soil and Food Crops Irrigated with Wastewater and the Potential Human Health Risk: A Metadata Analysis. Water 2021, 13, 3405. [Google Scholar] [CrossRef]
  59. Suryavathi, V.; Sharma, S.; Sharma, S.; Saxena, P.; Pandey, S.; Grover, R.; Kumar, S.; Sharma, K.P. Acute toxicity of textile dye wastewaters (untreated and treated) of Sanganer on male reproductive systems of albino rats and mice. Reprod. Toxicol. 2005, 19, 547–556. [Google Scholar] [CrossRef] [PubMed]
  60. Hobbs, C.A.; Saigo, K.; Koyanagi, M.; Hayashi, S.M. Magnesium stearate, a widely-used food additive, exhibits a lack of in vitro and in vivo genotoxic potential. Toxicology 2017, 4, 554–559. [Google Scholar] [CrossRef]
  61. Fernandes, A.; Pinto, B.; Bonardo, L.; Royo, B.; Robalo, M.P.; Martins, L.O. Wasteful Azo Dyes as a Source of Biologically Active Building Blocks. Front. Bioeng. Biotechnol. 2021, 9, 672436. [Google Scholar] [CrossRef]
  62. Biswas, M.M.; Taylor, K.E.; Bewtra, J.K.; Biswas, N. Enzymatic treatment of sulfonated aromatic amines generated from reductive degradation of reactive azo dyes. Water Environ. Res. 2007, 79, 351–356. [Google Scholar] [CrossRef] [PubMed]
  63. Tuormaa, T.E. The adverse effects of food additive on health. J. Orthomol. Med. 1994, 9, 225–243. [Google Scholar]
  64. New Jersey Department of Health and Senior Services. Hazardous Substances Fact Sheet [CAS Number: 91-59-8]. Revised 2004. Available online: https://nj.gov/health/eoh/rtkweb/documents/fs/1324.pdf (accessed on 12 December 2022).
  65. National Center for Biotechnology Information. PubChem Compound Summary for CID 7057, 2-Naphthylamine. 2022. Available online: https://pubchem.ncbi.nlm.nih.gov/compound/2-Naphthylamine (accessed on 31 December 2022).
Figure 1. Chemical structures of Azorubine E-122, Tartrazine E-102 and Allura red AC E-129.
Figure 1. Chemical structures of Azorubine E-122, Tartrazine E-102 and Allura red AC E-129.
Sustainability 15 02897 g001
Figure 2. Quantification of genomic DNA of plant growth-promoting rhizobacteria (PGPR) on Agarose gel electrophoresis. M represents the DNA ladder of 1 kb, and the extracted genomic DNA samples of rhizobacteria are from well A to D.
Figure 2. Quantification of genomic DNA of plant growth-promoting rhizobacteria (PGPR) on Agarose gel electrophoresis. M represents the DNA ladder of 1 kb, and the extracted genomic DNA samples of rhizobacteria are from well A to D.
Sustainability 15 02897 g002
Figure 3. Interpretation of comet scores by using the comet scoring software. Images from B1 to B5 and B6 to B10, along with controls, show the comets of Azorubine E-122 at both 24 and 48 h exposure time, whereas the comets of Tartrazine E-102 are shown from T1 to T5 and T6 to T10 and Allura red AC E-129 are shown from A1 to A5 and A6 to A10 at both 24 and 48 h exposure times. Dilution concentrations were 0.25, 0.50, 0.75, 1.0 and 1.25.
Figure 3. Interpretation of comet scores by using the comet scoring software. Images from B1 to B5 and B6 to B10, along with controls, show the comets of Azorubine E-122 at both 24 and 48 h exposure time, whereas the comets of Tartrazine E-102 are shown from T1 to T5 and T6 to T10 and Allura red AC E-129 are shown from A1 to A5 and A6 to A10 at both 24 and 48 h exposure times. Dilution concentrations were 0.25, 0.50, 0.75, 1.0 and 1.25.
Sustainability 15 02897 g003
Table 1. Quantification of DNA samples of three synthetic food colors, Azorubine E-122, Tartrazine E-102 and Allura red AC E-129, showing purity.
Table 1. Quantification of DNA samples of three synthetic food colors, Azorubine E-122, Tartrazine E-102 and Allura red AC E-129, showing purity.
TimeDilutionsAzorubine E-122Tartrazine E-102Allura Red AC E-129
Concentration
(μg/mL)
Ratio
260/280
Concentration
(μg/mL)
Ratio
260/280
Concentration
(μg/mL)
Ratio
260/280
24 hControl481.83521.89391.82
0.25%401.75361.4531.6
0.50%291.59401.83242.2
0.75%421.66521.75281.4
1.00%371.5481.63381.93
1.25%362.2331.66421.6
48 hControl341.81371.79601.8
0.25%401.6421.5251.5
0.50%301.5271.75531.63
0.75%411.22501.73441.57
1.00%501.96342.2342.2
1.25%441.99291.8491.75
Table 2. Comet measurement results determine the DNA damage, which indicates the size of the head and tail of comets at different concentrations of synthetic food color.
Table 2. Comet measurement results determine the DNA damage, which indicates the size of the head and tail of comets at different concentrations of synthetic food color.
TimeConcentration of DilutionAzorubine E-122Tartrazine E-102Allura Red AC E-129
Head LengthTail LengthComet LengthDNA HeadDNA TailHead LengthTail LengthComet LengthDNA HeadDNA TailHead LengthTail LengthComet LengthDNA HeadDNA Tail
24 h Control15318100−0.18.01.09.0100−0.1536.0100−0.1
0.25%1041210037.02.09.0831810312991
0.50%839461883141428614097146932
0.75%193554307131356650522024446141
1.00%102535317010253531702355787328
1.25%4151815865.0111636651133442873
48 h Control333.036100−0.1201.021100−0.114215100−0.1
0.25%515198813154.01988131333468120
0.50%16173336651415295447715212576
0.75%202343317019193821801019294556
1.00%31921119010122224771034432279
1.25%652578934.09.013227984552991
Table 3. Statistical analyses of genotoxicity variance (one-way ANOVA) between two groups of treatment.
Table 3. Statistical analyses of genotoxicity variance (one-way ANOVA) between two groups of treatment.
Groups of TreatmentSource of VariationSSdfMSFp-Value
24 hBetween groups10327780.7530.488
Within groups20,446151034
Total20,55017
48 hBetween groups103251.70.03790.963
Within groups20,446151363
Total20,55017
Table 4. Comparative statistical analyses of DNA damage caused by E-122, E-102 and E-129 at treatments 24 and 48 h.
Table 4. Comparative statistical analyses of DNA damage caused by E-122, E-102 and E-129 at treatments 24 and 48 h.
GroupsCountSumAverageVariance
E-122630851.41617
24 hE-102624240.4738.4
E-129617228.6747.5
E-122632754.61525
48 hE-102629448.91252
E-129631953.21311
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

John, A.; Luqman, M.; Muhammad, S.; Hanif, U.; Sardar, A.A.; Ali, S.; Hasnain, A.; Tufail, M.; Khan, Z.I.; Hussain, M.I.; et al. Genotoxicity of Synthetic Food Colors on Nitrogen-Fixing Bacteria in Agricultural Lands Irrigated with Wastewater of Corresponding Industries. Sustainability 2023, 15, 2897. https://doi.org/10.3390/su15042897

AMA Style

John A, Luqman M, Muhammad S, Hanif U, Sardar AA, Ali S, Hasnain A, Tufail M, Khan ZI, Hussain MI, et al. Genotoxicity of Synthetic Food Colors on Nitrogen-Fixing Bacteria in Agricultural Lands Irrigated with Wastewater of Corresponding Industries. Sustainability. 2023; 15(4):2897. https://doi.org/10.3390/su15042897

Chicago/Turabian Style

John, Arooba, Muhammad Luqman, Sohaib Muhammad, Uzma Hanif, Andleeb Anwar Sardar, Shaukat Ali, Ali Hasnain, Matiba Tufail, Zafar Iqbal Khan, Muhammad Iftikhar Hussain, and et al. 2023. "Genotoxicity of Synthetic Food Colors on Nitrogen-Fixing Bacteria in Agricultural Lands Irrigated with Wastewater of Corresponding Industries" Sustainability 15, no. 4: 2897. https://doi.org/10.3390/su15042897

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

Article Metrics

Back to TopTop