Next Article in Journal
Investigating the Relationships between Taste Preferences and Beverage Intake in Preadolescents
Next Article in Special Issue
Alginate/Pectin Film Containing Extracts Isolated from Cranberry Pomace and Grape Seeds for the Preservation of Herring
Previous Article in Journal
Longitudinal Study for the Detection and Quantification of Campylobacter spp. in Dairy Cows during Milking and in the Dairy Farm Environment
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Identification of Antibacterial Components and Modes in the Methanol-Phase Extract from a Herbal Plant Potentilla kleiniana Wight et Arn

1
Key Laboratory of Quality and Safety Risk Assessment for Aquatic Products on Storage and Preservation (Shanghai), Ministry of Agriculture and Rural Affairs of the People’s Republic of China, Shanghai 201306, China
2
College of Food Science and Technology, Shanghai Ocean University, Shanghai 201306, China
*
Author to whom correspondence should be addressed.
Foods 2023, 12(8), 1640; https://doi.org/10.3390/foods12081640
Submission received: 17 March 2023 / Revised: 6 April 2023 / Accepted: 7 April 2023 / Published: 13 April 2023

Abstract

:
The increase in bacterial resistance and the decline in the effectiveness of antimicrobial agents are challenging issues for the control of infectious diseases. Traditional Chinese herbal plants are potential sources of new or alternative medicine. Here, we identified antimicrobial components and action modes of the methanol-phase extract from an edible herb Potentilla kleiniana Wight et Arn, which had a 68.18% inhibition rate against 22 species of common pathogenic bacteria. The extract was purified using preparative high-performance liquid chromatography (Prep-HPLC), and three separated fragments (Fragments 1–3) were obtained. Fragment 1 significantly elevated cell surface hydrophobicity and membrane permeability but reduced membrane fluidity, disrupting the cell integrity of the Gram-negative and Gram-positive pathogens tested (p < 0.05). Sixty-six compounds in Fragment 1 were identified using Ultra-HPLC and mass spectrometry (UHPLC-MS). The identified oxymorphone (6.29%) and rutin (6.29%) were predominant in Fragment 1. Multiple cellular metabolic pathways were altered by Fragment 1, such as the repressed ABC transporters, protein translation, and energy supply in two representative Gram-negative and Gram-positive strains (p < 0.05). Overall, this study demonstrates that Fragment 1 from P. kleiniana Wight et Arn is a promising candidate for antibacterial medicine and food preservatives.

1. Introduction

Infectious diseases caused by pathogenic bacteria continue to be a global concern for public health, causing millions of deaths worldwide per year [1]. Since the introduction of sulfonamides in 1933, a large number of antibiotics have been applied in clinics [2]. Nevertheless, in recent decades, the overuse and/or misuse of antibiotics have accelerated the spread of antibiotic-resistant bacteria, leading to ineffective drug treatment [3]. It was estimated that at least 700,000 people worldwide die each year due to antimicrobial resistance [4].
Pharmacophagous plants are recognized as a rich source of phytochemicals with antimicrobial potential [5]. Phytocompounds extracted from such plants are long known for their therapeutic uses, and characterized by safety and low toxicity [6]. The application of herbal products may be a better choice for the extensive and imprudent use of synthetic antibiotics [7]. For example, In China, approximately 34,984 native higher plant species have been recorded [8]. Of these, the herbal plant Potentilla kleiniana Wight et Arn was first recorded in the earliest pharmaceutical book “Divine Farmer’s Classic of Materia Medica” during the Warring States period (475–221 B.C.) in China. It belongs to the phylum of Angiospermae, the class of Dicotyledoneae, the order of Rosales Bercht. and J. Presl, and the family of Rosaceae Juss. P. kleiniana Wight et Arn is widely distributed in China, and many Asian countries such as Japan, India, Malaysia, Indonesia, and North Korea. Its whole plant has been used as a traditional Chinese medicine to treat fever, arthritis, malaria, insect and snake bites, hepatitis, and traumatic injury [9]. Recently, Zhou et al. identified bioactive components from P. kleiniana Wight et Arn with anti-human immunodeficiency virus-1 (HIV-1) protease activity [10]. Liu et al. developed an efficient method for the rapid screening and separation of α-glucosidase inhibitors from P. kleiniana Wight et Arn [11]. Li et al. [12] found antihyperglycemic and antioxidant effect of the total flavones of P. kleiniana Wight et Arn in streptozotocin induced diabetic rats, which may be helpful in the prevention of diabetic complications associated with oxidative stress [12]. However, to the best of our knowledge, there are few studies so far in the current literature on antibacterial activity of P. kleiniana Wight et Arn. Tao et al. [9] reported that total flavonoids from P. kleiniana Wight et Arn (TFP) inhibited biofilm formation and virulence factor production in methicillin-resistant Staphylococcus aureus (MRSA). The TFP also damaged cell membrane integrity of Pseudomonas aeruginosa. These results supported potential application of the TFP as a novel natural bioactive preservative in food processing [13]. Song et al. also reported that bioactive components extracted from P. kleiniana Wight et Arn showed antibacterial effects against S. aureus, Candida albicans, P. aeruginosa, and Escherichia coli, but not against the mold Aspergillus niger [14].
To further exploit bioactive nature products in P. kleiniana Wight et Arn, in the present study, we extracted bacteriostatic components in P. kleiniana Wight et Arn using the methanol and chloroform method [15,16]. Antimicrobial action modes of the methanol-phase extract were further investigated. The results of this study provide useful data for potential pharmaceutical application of P. kleiniana Wight et Arn against the common pathogenic bacteria.

2. Results and Discussion

2.1. Antibacterial Activity of Crude Extracts from P. kleiniana Wight et Arn

Antibacterial substances in the fresh P. kleiniana Wight et Arn were extracted using the methanol and chloroform method [15,16]. The water loss rate of the fresh plant sample was 94.12% after freeze-drying treatment of the sample. The extraction rates of the methanol-phase and chloroform-phase crude extracts were 31.13% and 25.43%, respectively. As shown in Table 1, the chloroform-phase extract from P. kleiniana Wight et Arn had a 50.00% inhibition rate, which inhibited one species of Gram-positive bacterium S. aureus, and 10 species of Gram-negative bacteria, including Bacillus cereus A1-1, B. cereus A2-2, Enterobacter cloacae ATCC13047, Salmonella typhimurium ATCC15611, Shigella dysenteriae CMCC51252, Shigella flexneri CMCC51572, Shigella sonnei ATCC25931, Vibrio cholerae Q10-54, Vibrio mimicus bio-56759, Vibrio parahemolyticus ATCC33847, V. parahemolyticus B3-13, V. parahemolyticus B5-29, V. parahemolyticus B9-35, V. parahemolyticus A1-1, and Vibrio vulnificus ATCC27562 (Table 1).
Of note, the methanol-phase crude extract from P. kleiniana Wight et Arn inhibited the growth of 15 bacterial species, including one species of Gram-positive S. aureus, and 14 species of Gram-negative bacteria, P. aeruginosa ATCC9027, S. typhimurium ATCC15611, S. dysenteriae CMCC51252, S. flexneri CMCC51572, S. flexneri CMCC51574, S. sonnei ATCC25931, V. alginolyticus ATCC17749, V. cholerae Q10-54, V. fluvialis ATCC33809, V. mimicus bio-56759, V. parahemolyticus ATCC17802, and V. vulnificus ATCC27562, which showed a 68.18% inhibition rate (Table 1, Figure 1).
In this study, the methanol and chloroform extract method exhibited a broader antibacterial spectrum, consistent with our previous reports [15,16]. Previous studies also reported effective extraction of bioactive compounds from P. kleiniana Wight et Arn. For example, Tao et al. [13] extracted TFP in P. kleiniana Wight et Arn using an ethanol-water solution, and the obtained extract was further partitioned using petroleum ethers, chloroform and ethyl acetate. The extracted TFP inhibited survival and virulence of P. aeruginosa, and MRSA. Song et al. [14] extracted bioactive compounds from P. kleiniana Wight et Arn using ethanol and ethyl acetate, and the obtained extract showed antibacterial activity against P. aeruginosa, S. aureus, C. albicans, and E. coli. The difference in bioactive compounds extracted from P. kleiniana Wight et Arn using the different methods may explain the distinct antibacterial profiles between this study and the previous reports [13,14].
We further determined minimum inhibitory concentrations (MICs) of the crude extracts from P. kleiniana Wight et Arn, and the results are shown in Table 1. The MICs of the chloroform-phase extract ranged from 12.5 mg/mL to 50 mg/mL against the eleven species of the bacteria. Notably, for the methanol-phase extract, the MICs were between 1.56 mg/mL and 50 mg/mL against the fifteen bacterial species. Of these, the growth of B. cereus A2-2 and V. parahemolyticus ATCC17802 was the most strongly repressed by the methanol-phase extract with the MICs of 1.56 mg/mL, followed by V. alginolyticus ATCC17749, V. mimicus bio-56759, V. parahemolyticus B3-13, V. parahemolyticus B5-29, V. parahemolyticus B9-35, and V. parahemolyticus A1-1 with MICs of 3.13 mg/mL. In addition, the growth of B. cereus A1-1, P. aeruginosa ATCC9027, S. typhimurium ATCC15611, S. flexneri CMCC51572, S. aureus ATCC8095, and V. parahemolyticus B4-10 was also inhibited by the methanol-phase extract with lower MICs (6.25 mg/mL). Of these pathogens, for example, V. alginolyticus is a foodborne marine Vibrio that can cause gastroenteritis, otitis media, otitis externa, and septicemia in humans [17]. V. mimicus can also cause gastroenteritis in humans due to contaminated fish consumption and seafood [18]. P. aeruginosa is an opportunistic pathogen and can cause serious infections, especially in patients with compromised immune systems [19].
Recently, Song et al. [14] reported that the ethyl acetate extract of P. kleiniana Wight et Arn inhibited E. coli, P. aeruginosa, and C. albicans, with MICs of 5 mg/mL, 2.5 mg/mL, and 5 mg/mL, respectively. Tao et al. reported the MIC value of the TFP against MRSA was 20 μg/mL [9].
These results indicated that the methanol-phase crude extract had a higher inhibition rate (68.18%), showing a more broad inhibitory profile with much lower MICs (1.56–50 mg/mL) against the pathogens tested, as compared to the chloroform-phase crude extract (50.00%; 12.5–50 mg/mL). Thus, the methanol-phase crude extract was chosen for further analysis in this study.

2.2. Purification of the Methanol-Phase Crude Extract from P. kleiniana Wight et Arn

Based on the obtained results, a large amount of the methanol-phase crude from P. kleiniana Wight et Arn was prepared and further purified using Prep-HPLC analysis. As shown in Figure S1, three separated fragments (designated Fragments 1–3) were observed via scanning at OD211 for 12 min, including Fragment 1 (2.45 min), Fragment 2 (6.75 min), and Fragment 3 (9.83 min). The main peak of the methanol-phase crude was observed to occur at 2.45 min, wherein the absorption peak of Fragment 1 reached its maximum.
The three single fragments were subjected for antibacterial activity analysis. Fragment 1 had strong inhibitory effects on V. parahemolyticus ATCC17802, V. parahemolyticus B5-29, V. parahemolyticus B9-35, V. parahemolyticus B3-13, and V. parahemolyticus B4-10. In addition, the growth of the other six strains was also effectively repressed, including B. cereus A2-2, V. parahemolyticus A1-1, S. flexneri CMCC51572, S. aureus ATCC25923, S. aureus ATCC8095, and S. aureus ATCC6538 (Table 2). Of these, V. parahaemolyticus is a Gram-negative halophilic bacterium that can cause diseases in marine animals, leading to huge economic losses to the aquaculture. V. parahaemolyticus can also cause gastrointestinal infections and other health complications in humans [20]. B. cereus is a Gram-positive foodborne pathogen that can cause diarrhea and emesis [21]. S. flexneri is a Gram-negative intracellular pathogen that invades colonic cells and causes bloody diarrhea in humans [22]. S. aureus is a Gram-positive opportunistic pathogen leading to food poisoning as well as human and animal infectious diseases [23,24].
We also determined MICs of Fragment 1 against the four species of pathogenic bacteria (Table 2). The synergistic effect may explain the observed MICs of Fragment 1 (6.25–50 mg/mL), as compared to the methanol-phase extract from P. kleiniana Wight et Arn. Among the Gram-negative pathogens, V. parahemolyticus ATCC17802 and V. parahemolyticus B5-29 were the most sensitive strains to Fragment 1, with MICs of 6.25 mg/mL. For the Gram-positive pathogen, the growth of S. aureus ATCC8095 and S. aureus ATCC25923 was also effectively repressed, with MICs of 6.25 mg/mL and 12.5 mg/mL, respectively.
Conversely, the other two peaks (Fragments 2 and 3) showed weak or no antibacterial activity. To further investigate possible antibacterial modes of Fragment 1, the two Gram-negative strains V. parahemolyticus ATCC17802 and V. parahemolyticus B5-29, and two Gram-positive stains S. aureus ATCC8095 and S. aureus ATCC25923 were chosen for the further analysis in this study.

2.3. Bacterial Cell Surface Hydrophobicity, Membrane Fluidity and Permeability Changes Triggered by Fragment 1 from P. kleiniana Wight et Arn

2.3.1. Cell Surface Hydrophobicity

Cell surface hydrophobicity is an important cellular biophysical parameter that affects cell surface interactions and cell–cell communication [25]. In this study, the hexadecane was used as a probe to assess cell surface hydrophobicity change. The difference between before and after the absorbance value of bacterial fluid can indicate the change of hydrophobicity, and the larger the difference, the more hydrophobicity of the surface [26]. The cell surface hydrophobicity of the four experimental groups (1× MIC of Fragment 1) was significantly increased, as compared to the control groups (p < 0.05) (Figure 2A). For instance, after being treated with Fragment 1 for 2 h, bacterial cell surface hydrophobicity was significantly increased, including V. parahaemolyticus B5-29 (8.62%, 1.42-fold), V. parahaemolyticus ATCC17802 (8.27%, 1.50-fold), S. aureus ATCC25923 (10.34%, 1.24-fold), and S. aureus ATCC8095 (12.20%, 1.19-fold) (p < 0.05). Increasing treatment time, the cell surface hydrophobicity was further increased. After the 4 h treatment, the cell surface hydrophobicity was the most significantly increased (11.97%, 1.97-fold) in the V. parahaemolyticus B5-29 treatment group. The highest increase (15.96%, 2.63-fold) was also observed in V. parahaemolyticus B5-29, after treatment for 6 h. The results indicated that Fragment 1 from P. kleiniana Wight et Arn can significantly increase the cell surface hydrophobicity of both Gram-negative V. parahemolyticus and Gram-positive S. aureus pathogens.

2.3.2. Cell Membrane Fluidity

Cell membrane is a natural barrier to prevent extracellular substances from freely entering the cell [27]. In this study, as shown in Figure 2B, when compared to the control groups, the membrane fluidity of V. parahaemolyticus B5-29, S. aureus ATCC25923, and S. aureus ATCC8095 did not change significantly after treatment with Fragment 1 (1× MIC) for 2 h and 4 h. However, a significant decrease (1.16-fold, 1.25-fold, and 1.24-fold) was observed in these three treatment groups after treatment for 6 h, respectively (p < 0.05). In addition, a significant decrease in cell membrane fluidity was only found in V. parahaemolyticus ATCC17802 after treatment for 4 h (1.16-fold) and 6 h (1.24-fold), respectively (p < 0.05). These results indicated that Fragment 1 from P. kleiniana Wight et Arn can significantly reduce the cell membrane fluidity of both Gram-negative V. parahemolyticus and Gram-positive S. aureus pathogens.

2.3.3. Cell Membrane Permeability

β-galactosidase is a macromolecular protein naturally found in the interior of cells that can hydrolyze the substrate o-nitrophenyl-β-D-galactopyranosi (ONPG) to galactose and o-nitrophenol in yellow. If the inner membrane of bacterial cells is damaged, ONPG will quickly enter the cell [28]. In this study, the ONPG was used as a probe to assess whether the bacterial inner membrane is damaged. As illustrated in Figure 3D, the inner cell membrane permeability of S. aureus ATCC8095 did not change significantly after treatment with Fragment 1 (1× MIC) from P. kleiniana Wight et Arn for 2 h (p > 0.05); conversely, significant increases were observed in V. parahaemolyticus B5-29, V. parahaemolyticus ATCC17802, and S. aureus ATCC25923 treatment groups (1.15-fold, 1.18-fold, and 1.04-fold), respectively (p < 0.05). After being treated for 4 h, the highest increase was found in V. parahaemolyticus B5-29 (1.22-fold). After treatment for 6 h, significant increases were also observed in V. parahaemolyticus B5-29, V. parahaemolyticus ATCC17802, S. aureus ATCC25923, and S. aureus ATCC8095 (1.20-fold, 1.17-fold, 1.07-fold, and 1.08-fold), respectively (p < 0.05). These results indicated that Fragment 1 from P. kleiniana Wight et Arn can significantly increase the inner cell membrane permeability of both Gram-negative V. parahemolyticus and Gram-positive S. aureus pathogens.
Outer membrane permeability was assessed by measuring the uptake of a hydrophobic fluorescent probe N-phenyl-1-naphthylamine (NPN) [29]. The outer membrane permeability increased significantly in the four treatment groups, after being treated with Fragment 1 for 2 h (1.38-fold to 1.66-fold) (p < 0.01), and 4 h (1.77-fold to 2.72-fold), respectively (p < 0.001) (Figure 2C). The highest increase was found in V. parahaemolyticus ATCC17802 (2.70-fold), after treatment for 6 h. These results indicated that Fragment 1 from P. kleiniana Wight et Arn can significantly increase the outer cell membrane permeability of the Gram-negative V. parahemolyticus and Gram-positive S. aureus pathogens. Recently, Tao et al. also reported that the TFP from P. kleiniana Wight et Arn increased cell membrane permeability of MRSA [13].
Taken together, the results of this study demonstrated that Fragment 1 (1× MIC) from P. kleiniana Wight et Arn can significantly increase the cell surface hydrophobicity and membrane permeability, but decreases the cell membrane fluidity of both Gram-negative V. parahemolyticus and Gram-positive S. aureus pathogens. Antibacterial compounds (e.g., flavonoids) in Fragment 1 from P. kleiniana Wight et Arn may have interacted with lipid components of the bacterial cell membrane. The disorder in lipid chains resulted in changed permeability and fluidity of the bacterial cell membrane [30]. The compounds may also have interacted with the bacterial cell surface proteins, leading to the altered nanomechanical properties, which consequently changed cell surface hydrophobicity and fluidity [31]. The two common pathogens V. parahemolyticus and S. aureus were chosen for further analysis in this study. The former is the leading sea foodborne pathogen worldwide [20], while the latter leads to food poisoning, as well as human and animal infections [23].

2.4. Bacterial Cell Surface Structure Changes Triggered by Fragment 1 from P. kleiniana Wight et Arn

Based on the obtained results in this study, the representative Gram-negative V. parahaemolyticus ATCC17802 and Gram-positive S. aureus ATCC25923 strains were chosen for further scanning electron microscope (SEM) analysis. As shown in Figure 4, the cells of V. parahaemolyticus ATCC17802 were intact in shape with a flat surface, showing a typical rod-like structure, while those of S. aureus ATCC25923 were also intact and clear, showing a typical spherical structure. In remarkable contrast to the control groups, the bacterial morphological structures were altered to varying degrees in the treatment groups triggered by Fragment 1 (1× MIC) for different times.
For the Gram-negative V. parahaemolyticus ATCC17802, its cell surface was slightly shrunken after being treated with Fragment 1 for 2 h. After 4 h of treatment, the cell surface was more wrinkled and was slightly depressed, the cell membrane was folded and some contents were exuded. After 6 h of the treatment, the cells were severely deformed and crumpled, with a large amount of content leaked.
For the Gram-positive S. aureus ATCC25923, its cell surface was rough and slightly wrinkled, but certain cells were depressed, with a small amount of content leaked after the treatment for 2 h. Upon the increased treatment time (4 h), more cells were obviously wrinkled and deformed with the irregularly spherical, and more content leaked out. The cell morphological structure was seriously damaged after being treated for 6 h.
These results demonstrated that Fragment 1 (1× MIC) from P. kleiniana Wight et Arn can severely damage the cell surface structure of both Gram-negative V. parahaemolyticus and Gram-positive S. aureus after treatment for 6 h.

2.5. Identification of Potential Antibacterial Compounds in Fragment 1 from P. kleiniana Wight et Arn

Potential antibacterial components in Fragment 1 from P. kleiniana Wight et Arn were further identified using UHPLC-MS analysis. As shown in Table 3, a total of 66 different compounds were identified. The highest relative percentage of the compounds was D-maltose (6.77%), followed by oxymorphone (6.29%), rutin (6.29%), D-proline (5.41%), and L-proline (5.41%). In addition, alkaloids, flavonoids, phenols, sesquiterpenoids, fatty acyls, and organic acids were also detected (Table 3).
Highly concentrated sugar solutions, such as the D-maltose identified in this study, are known to be effective antimicrobial agents [32]. Previous research has indicated that the antibacterial activity of phenanthrenes and derivatives, such as the oxymorphone identified in this study, was primarily related to the destruction of the bacterial cell wall structure [33]. Plant extracts contain a large number of bioactive compounds, mainly polyphenols including flavonoids and phenolic compounds. Flavonoids, such as the rutin identified in this study, could exert antibacterial activity via damaging the cytoplasmic membrane, inhibiting energy metabolism and synthesis of nucleic acids [34]. Tao et al. also reported the major compounds of the TFP were 3-O-methylducheside A, naringenin, rutin and quercetin [9,13]. Phenols, such as the p-octopamine identified in this study, are potent antibacterial agents against both Gram-positive and Gram-negative bacteria via the disruption of the bacterial membrane, leading to bacterial lysis and leakage of intracellular contents [35]. Indole alkaloids, such as the indole identified in this study, possess not only intriguing structural features but also biological/pharmacological activities e.g., antimicrobial activity [36]. Additionally, amino acids and its derivatives, such as the D-proline, L-proline, glutamic acid, 5-aminovaleric acid, lysine, pipecolic acid, and L-valine identified in this study, are a kind of antibacterial agent with the advantages of being not easily drug-resistant, and having low toxicity or harmless metabolites [37].

2.6. Differential Transcriptomes Triggered by Fragment 1 from P. kleiniana Wight et Arn

To obtain the genome-wide gene expression changes triggered by Fragment 1 from P. kleiniana Wight et Arn, we determined transcriptomes of the Gram-negative V. parahaemolyticus ATCC17802 and the Gram-positive S. aureus ATCC25923 pathogens treated with Fragment 1 (1× MIC) for 6 h using the Illumina RNA sequencing technology. A complete list of differently expressed genes (DEGs) in the two strains are available in the National Center for Biotechnology Information (NCBI) SRA database under the accession number PRJNA906658.

2.6.1. The Major Changed Metabolic Pathways in V. parahaemolyticus ATCC17802

Approximately 13.07% (580 of 4436 genes) of V. parahaemolyticus ATCC17802 genes were differentially expressed in the treatment group, as compared to the control group. Of these, 238 DEGs showed higher transcriptional levels (fold change ≥ 2.0), whereas 342 DEGs were significantly down-regulated (fold change ≤ 0.5) (p < 0.05). Sixteen significantly altered metabolic pathways were identified in V. parahaemolyticus ATCC 17802, including the citrate cycle; glyoxylate and dicarboxylate metabolism; fatty acid degradation; glycine, serine, and threonine metabolism; oxidative phosphorylation; pyruvate metabolism; propanoate metabolism; beta-Lactam resistance; ABC transporters; two-component system; alanine, aspartate, and glutamate metabolism; phosphotransferase system (PTS); butanoate metabolism; lysine degradation; quorum sensing (QS); and nitrogen metabolism (Figure 5, Table 4).
In the citrate cycle, all the DEGs (n = 14) were significantly repressed (0.146-fold to 0.35-fold) (p < 0.05) in V. parahaemolyticus ATCC17802 after treatment by Fragment 1 from P. kleiniana Wight et Arn. For instance, the DEGs (sucABCD, WU75_19785 and WU75_19790, WU75_19795, and WU75_19800), encoding a 2-oxoglutarate dehydrogenase, a dihydrolipoamide succinyltransferase, and succinyl-CoA synthetase subunits alpha and beta, respectively, were highly inhibited (0.146-fold, 0.133-fold, 0.134-fold, and 0.16-fold) (p < 0.05). Moreover, the DEGs (sdhABCD, WU75_19775, WU75_19780, WU75_19765, and WU75_19770) encoding a succinate dehydrogenase were also highly repressed (0.144-fold to 0.199-fold) (p < 0.05), which links two essential energy-producing processes, the citrate cycle and oxidative phosphorylation [38]. The inhibited key enzymes in the citrate cycle highlighted inactive energy production in V. parahaemolyticus ATCC17802 triggered by Fragment 1.
In the propanoate metabolism, all the DEGs (n = 2) were significantly inhibited (0.402-fold to 0.435-fold) in the V. parahaemolyticus ATCC17802 treatment group (p < 0.05). For example, the DEG (prpC, WU75_15770) encoding a 2-methylcitrate synthase was significantly inhibited (0.435-fold) (p < 0.05). It has been reported that the strategic inhibition of organic acid catabolism in P. aeruginosa through inhibition of PrpC activity may be a potent mechanism to halt the growth of this pathogen [39].
In the glyoxylate and dicarboxylate metabolism, five of the six DEGs were significantly repressed (0.129-fold to 0.277-fold) (p < 0.05). For instance, the DEGs (aceAB, WU75_19150, WU75_19145, and WU75_00290), encoding an isocitrate lyase and a malate synthase of the glyoxylate shunt (GS) carbon cycle, were significantly inhibited (0.315-fold to 0.370-fold) (p < 0.05). The GS could avoid unnecessary reactive oxygen species (ROS) generation by bypassing nicotinamide adenine dinucleotide (NADH) production, and respiration, eventually helping cells to survive in harsh conditions [40,41].
In the glycine, serine, and threonine metabolism, all the DEGs (n = 15) were significantly inhibited (0.113-fold to 0.495-fold) in V. parahaemolyticus ATCC17802 (p < 0.05). For example, the DEGs (ectBAC, WU75_16140, WU75_16145, and WU75_16135), encoding a diaminobutyrate-2-oxoglutarate aminotransferase, a 2% 2C4-diaminobutyric acid acetyltransferase, and an ectoine synthase, which are involved in the synthesis of ectoine that is commonly found in halophilic and halotolerant microorganisms to maintain cell osmotic balance [42]. Additionally, in the alanine, aspartate, and glutamate metabolism, ten of the thirteen DEGs were significantly down-regulated (0.037-fold to 0.466-fold) in V. parahaemolyticus ATCC17802 as well (p < 0.05). Conversely, the DEGs (ansAB, WU75_20915, and WU75_01110) were up-regulated (2.141-fold and 2.718-fold) (p < 0.05), which encoded a cytoplasmic asparaginase I and a L-asparaginase II. The asparaginase I is required for bacterial growth on asparagine as the sole nitrogen source [43], while asparaginases are important in maintaining nitrogen balance and the levels of amino acids within cells [43]. These results indicated that the amino acid synthesis was inhibited in V. parahaemolyticus ATCC17802 mediated by Fragment 1.
For the ABC transporters, 29 of the 35 DEGs were significantly down-regulated (0.106-fold to 0.491-fold) in V. parahaemolyticus ATCC17802 (p < 0.05). Of these, the DEGs (proVXW, WU75_10380, WU75_10390, and WU75_10385), encoding a choline ABC transporter ATP-binding protein, a choline ABC transporter substrate-binding protein, and a choline ABC transporter permease subunit that are responsible for the choline transport, were all significantly repressed (0.106-fold to 0.138-fold). The DEGs (oppABCDF, WU75_12765, WU75_12770, WU75_12775, WU75_12780, and WU75_12785) encoding a peptide ABC transporter substrate-binding protein, an oligopeptide transporter permease, a peptide ABC transporter permease, an oligopeptide transporter ATP-binding component, and a peptide ABC transporter ATP-binding protein, respectively, were all highly repressed (0.172-fold and 0.214-fold). Additionally, the DEGs (yejABE, WU75_13090, WU75_07210, WU75_07220, and WU75_07215) encoding a diguanylate cyclase, an ABC transporter permease subunit, and a peptide ABC transporter permease, respectively, were highly repressed as well (0.151-fold and 0.220-fold). The ABC transporter YejABEF is required for resistance to antimicrobial peptides and virulence of Brucella melitensis [44]. These results indicated that the inhibited ABC transporters likely led to the repressed substance transport and harmful substances discharged in V. parahaemolyticus ATCC17802.
In the oxidative phosphorylation, nine of the thirteen DEGs were significantly down-regulated in V. parahaemolyticus ATCC17802 (0.195-fold to 0.478-fold) (p < 0.05). Oxidative phosphorylation is a major metabolic pathway to obtain energy required for cell growth and proliferation [45] (Huang et al., 2019). For instance, the DEGs (ccoNOQ, WU75_14575, WU75_14570, and WU75_14565) were significantly inhibited (0.228-fold to 0.475-fold) (p < 0.05), which regulated the bacterial adhesion in environmental stresses in V. alginolyticus [45].
In the QS, most DEGs (n = 9) were significantly inhibited (0.109-fold to 0.484-fold) (p < 0.05), e.g., cytochrome c (WU75_06010), cytochrome B (WU75_06015), and peptidase S41 (WU75_14570). For instance, the cytochrome c mediates electron-transfer in the respiratory chain and acts as a detoxifying agent to dispose of reactive oxygen species (ROS) [46].
In contrast, in the PTS, nine of the eleven DEGs were significantly up-regulated (2.36-fold to 6.946-fold) in the V. parahaemolyticus ATCC17802 treatment group (p < 0.05). Of these, the DEGs (fruA, WU75_14960; ulaA, WU75_00450), encoding a PTS fructose transporter subunit IIBC and a PTS beta-glucoside transporter subunit IIBC, respectively, were highly up-regulated (5.096-fold and 6.946-fold) (p < 0.05).
In the nitrogen metabolism, most of the DEGs (n = 4) were significantly up-regulated (2.286-fold to 63.107-fold) (p < 0.05). Remarkably, the DEG (hcp, WU75_08850) encoding a hydroxylamine reductase was strongly up-regulated (63.107-fold) (p < 0.05), and is involved in the processes of scavenging hydroxylamine with NO detoxification [47].
In the two-component system, 19 DEGs were significantly inhibited (0.186-fold to 0.491-fold), whereas 9 DEGs were significantly enhanced (2.068-fold to 26.5-fold) (p < 0.05). The two-component system is one of the primary pathways by which bacteria adapt to environmental stresses [48]. For instance, the DEGs (cpxAR, WU75_18570, and WU75_18575) encoding a two-component sensor protein and a transcriptional regulator were strongly up-regulated (10.981-fold and 26.500-fold) (p < 0.05). The CpxAR is a key modulator of capsule export that facilitates Actinobacillus pleuropneumoniae survival in the host [49]. It also regulates cell membrane permeability and efflux pump activity and induces multidrug resistance (MDR) in Salmonella enteritidis [50].
Additionally, in the beta-lactam resistance, all the DEGs (acrAB, WU75_09925, WU75_09315, and WU75_09310) were strongly up-regulated (6.699-fold to 40.366-fold) in the V. parahaemolyticus ATCC17802 treatment group (p < 0.05), which encoded a multidrug efflux resistance nodulation division (RND) transporter periplasmic adaptor subunit and a multidrug transporter. The RND family efflux pumps, including the major pump AcrAB-TolC, are important mediators of intrinsic and evolved antibiotic resistance [51].
Taken together, these results indicated that Fragment 1 from P. kleiniana Wight et Arn can significantly change sixteen metabolic pathways in the Gram-negative V. parahaemolyticus ATCC17802, which consequently led to repressed substance transporting, energy production, and protein translation, but enhanced stringent response, and harmful substance discharging, and thereby cell death.

2.6.2. The Major Changed Metabolic Pathways in S. aureus ATCC25923

Approximately 7.3% (196 of 2672 genes) of S. aureus ATCC25923 genes were differentially expressed in the treatment group, as compared to the control group. Of these, 156 DEGs showed higher transcriptional levels (fold changes ≥ 2.0), whereas 40 DEGs were significantly down-regulated (fold changes ≤ 0.5) (p < 0.05). Based on the comparative transcriptomic analysis, seven significantly altered metabolic pathways were identified in S. aureus ATCC25923, including the two-component system; nitrogen metabolism; riboflavin metabolism; arginine and proline metabolism; atrazine degradation; alanine, aspartate and glutamate metabolism; and pyrimidine metabolism (Figure 6, Table 5).
In the arginine and proline metabolism, all the DEGs (n = 4) were significantly down-regulated at the transcription levels (0.109-fold to 0.461-fold) in S. aureus ATCC25923 (p < 0.05). The arginine metabolism converts L-arginine to urea and L-ornithine, which are further metabolized into proline and polyamides that drive collagen synthesis and bioenergetic pathways critical for cell proliferation, respectively [52]. For instance, the DEG (rocF, KQ76_11235) encoding an arginase was significantly down-regulated (0.461-fold) (p < 0.05), and was associated with the ability of Helicobacter pylori to establish chronic infections [53].
All the DEGs (n = 4) in the riboflavin metabolism were also significantly inhibited (ribBADEH, 0.3734-fold to 0.480-fold) (p < 0.05). In this pathway, the redox cofactors flavin mononucleotide and flavin adenine dinucleotide and their precursor riboflavin play important roles in many cellular processes, such as respiration, DNA repair, biosyntheses of heme groups, cofactors and nucleotides, fatty acid beta-oxidation, and bioluminescence [54].
Bacteria use two-component signal transduction systems to elicit adaptive responses to environmental changes [55]. In this study, seven DEGs in the two-component system were significantly up-regulated (2.117-fold to 28.924-fold) in S. aureus ATCC25923 (p < 0.05). For instance, the DEGs (agrB, KQ76_10520; and graS, KQ76_03245) encoding histidine kinases were significantly up-regulated by 2.565-fold and 2.989-fold, respectively (p < 0.05). The accessory gene regulator (agr) quorum-sensing system contributes to its pathogenicity of S. aureus [56]. GraS, the sensor histidine kinase of the GraXRS system, has been suggested to directly activate the response regulator ArlR [53]. Loss of the ArlR alone impairs the ability of S. aureus to respond to host-imposed manganese starvation and glucose limitation [57].
Interestingly, expression of all the DEGs (n = 7) in the nitrogen metabolism was significantly increased at the transcription level (3.529-fold to 10.404-fold) in S. aureus ATCC25923 (p < 0.05). The seven DEGs (nirBD, narHIJZT) were all involved in nitrate reduction [58,59,60]. Of these, the NirD (KQ76_12515) was a small subunit of cytoplasmic NADH-dependent nitrite reductase complex NirBD [61,62]. Over-expression of nirD limits RelA-dependent accumulation of guanosine 5′-triphosphate 3′-diphosphate ((p)ppGpp) in vivo and can prevent activation of the stringent response during amino acid starvation in E. coli [62].
In the alanine, aspartate, and glutamate metabolism, two DEGs (carBA, KQ76_05770 and KQ76_05765) encoding carbamoyl phosphate synthase were significantly up-regulated (2.154-fold and 3.084-fold) in S. aureus ATCC25923 (p < 0.05). The interface residues located near the CarB region of carboxy phosphate synthetic domain plays a key role in carbamoyl phosphate synthetase, aspartate transcarbamoylase, and dihydroorotase (CAD) complex regulation in the pyrimidine biosynthesis [63]. Correspondingly, in the pyrimidine metabolism, four DEGs (pyrBCR, KQ76_05755, KQ76_05760, and KQ76_05745) were also significantly up-regulated (2.968-fold to 3.213-fold) (p < 0.05), and encoded an aspartate carbamoyltransferase, a dihydroorotase, and a phosphoribosyl transferase, respectively. The pyrimidines are involved in the synthesis of DNA, RNA, lipids, and carbohydrates. The pyrimidine metabolism is involved in the synthesis, degradation, salvage, interconversion, and transport of these compounds [64].
Taken together, these results indicate that Fragment 1 from P. kleiniana Wight et Arn can significantly influence seven metabolic pathways in the Gran-positive S. aureus ATCC25923. Of these, the two-component system, alanine, aspartate and glutamate metabolism, and nitrogen metabolism were also changed in the Gram-negative V. parahaemolyticus ATCC17802, which led to the enhanced regulation of stringent response in the two pathogens. On the other hand, we also found distinct transcriptomic profiles between the Gram-positive and Gram-negative pathogens triggered by Fragment 1. For example, consistent with the results obtained from the cell structure analysis, V. parahaemolyticus ATCC17802 was more sensitive to Fragment 1 treatment, as more metabolic pathways were altered, such as the citrate cycle, glyoxylate and dicarboxylate metabolism, fatty acid degradation, glycine, serine and threonine metabolism, oxidative phosphorylation, pyruvate metabolism, propanoate metabolism, beta-lactam resistance, ABC transporters, PTS, butanoate metabolism, lysine degradation, and QS, which resulted in cell destruction and even death.
In addition, to validate the transcriptome data, we tested 16 representative DEGs (Table S1) via reverse transcription real time-quantitative PCR (RT-qPCR) analysis, and the resulting data were generally correlated with those yielded from the transcriptome analysis (Table S2).

3. Materials and Methods

3.1. Bacterial Strains and Culture Conditions

The bacterial strains and culture media used in this study are listed in Table S3. Vibrio strains and non-Vibrio strains were incubated as described in our recent studies [15,16,65].

3.2. Extraction of Bioactive Substances from P. kleiniana Wight et Arn

Fresh P. kleiniana Wight et Arn was purchased from the Qian Shan Zhen Pin shop in Guiyang City (26°36′5.01″ N, 106°41′19.90″ E), Guizhou Province, China, in October of 2021. Bioactive substances were extracted from the samples using the methanol and chloroform method described in our recent reports [15,16,66]. Briefly, aliquot of a 500 g of the whole plant sample was lyophilized, pulverised, powded, sonicated, and then filtered and collected for the secondary extraction. The methanol and chloroform phases were separated and then concentrated using the Rotary Evaporator (IKA, Staufen, Germany) [15,16].

3.3. Antimicrobial Susceptibility Assay

The susceptibility of the bacterial strains (Table S3) to the extracts from P. kleiniana Wight et Arn were determined according to the standard method issued by the Clinical and Laboratory Standards Institute, USA (CLSI, M100-S23, 2018). The antibacterial activity was defined as described previously [15,16]. Broth dilution testing (microdilution) (CLSI, M100-S18, 2018) was used to determine MICs of the extracts. The MIC was defined as described previously [15,16].

3.4. Prep-HPLC Analysis

Aliquots of the extracted samples (10 mg/mL) were resolved, centrifuged, filtered, and subjected for the Prep-HPLC Analysis, using Waters 2707 (Waters, Milford, MA, USA) linked with UPLC Sunfire C18 column (5 µm, 10 × 250 mm) (Waters, Milford, MA, USA) with the same parameters and elution conditions described in our recent reports [15,16].

3.5. UHPLC–MS Analysis

The UHPLC–MS analysis was conducted using the EXIONLC System (Sciex, Framingham, MA, USA) by Shanghai Hoogen Biotech, Shanghai, China [67].

3.6. Bacterial Cell Surface Hydrophobicity and Membrane Fluidity Assays

The cell surface hydrophobicity was measured according to the method of Cui et al. [68]. The cell membrane fluidity was measured according to the method of Kuhry et al. [69], using the 1,6-diphenyl-1,3,5-hexatriene (DPH, Sangon, Shanghai, China).

3.7. Cell Membrane Permeability Analysis

Cell outer membrane permeability was measured according to the method of Wang et al. [70], with the NPN solution (Sangon, Shanghai, China). The inner membrane permeability was measured according to the method of Huang et al. [71], with the ONPG solution (Sangon, Shanghai, China).

3.8. Scanning Electron Microscope (SEM) Assay

The preparation of the samples for the SEM analysis was performed using the method described in our recent reports [15,16,72]. The samples were observed using the Scanning Electron Microscope (Tescan Mira 3 XH, Tescan, Brno, Czech Republic, 5.0 kV, 30,000×).

3.9. Illumina RNA Sequencing

The bacterial cell culture at the mid-LGP was treated with Fragment 1 (1× MIC) from P. kleiniana Wight et Arn for 6 h, and then collected via centrifugation for the total RNA preparation [15,16,72]. Three independently prepared RNA samples for each strain were subjected for the Illumina RNA sequencing analysis, using Illumina HiSeq 2500 platform (Illumina, Santiago, CA, USA) [72].

3.10. RT-qPCR Assay

The RT-qPCR assay was performed according to the method described in our recent reports [15,16,72]. The oligonucleotide primers were designed (Table S1), and synthesized via Sangon (Shanghai, China).

3.11. Data Analysis

The DEGs were analyzed as described in our recent reports [15,16,72]. All tests were carried out in triplicate. The data were analyzed using the SPSS statistical analysis software version 17.0 (SPSS Inc., Armonk, NY, USA). One-way analysis of variance (ANOVA) was performed using the least-significant difference (LSD) method and homogeneity of variance test. There was no significant difference between the control and the treatment groups if the generalized p-values were more than 0.05; conversely, there was significant difference if p-values were less than 0.05.

4. Conclusions

In this study, the methanol-phase extract from P. kleiniana Wight et Arn showed an inhibition rate of 68.18% against 22 species of common pathogenic bacteria. The methanol-phase extraction inhibited the growth of one species of Gram-positive S. aureus, and 14 species of Gram-negative bacteria, including B. cereus, E. cloacae, E. coli, P. aeruginosa, S. typhimurium 1, S. dysenteriae, S. flexneri, S. flexneri, S. sonnei, V. alginolyticus, V. cholerae, V. fluvialis, V. mimicus, V. parahemolyticus, and V. vulnificus strains. This extract was further purified using the Prep-HPLC, and three separated fragments were obtained. Fragment 1 significantly increased bacterial cell surface hydrophobicity and membrane permeability and decreased membrane fluidity, disrupting the cell integrity of the Gram-positive and Gram-negative bacteria such as S. aureus ATCC25923, S. aureus ATCC8095, V. parahaemolyticus ATCC17802, and V. parahaemolyticus B5-29. The MIC values of Fragment 1 ranged from 6.25 mg/mL to 50 mg/mL. A total of 66 different compounds in Fragment 1 were identified. The highest relative percentage of the compounds was D-maltose (6.77%), followed by oxymorphone (6.29%), rutin (6.29%), D-proline (5.41%), and L-proline (5.41%). Highly concentrated sugar solutions, such as the D-maltose identified in Fragment 1, are known to be effective antimicrobial agents. The identified oxymorphone and rutin could exert antibacterial activity via damaging the bacterial cell wall and cytoplasmic membrane, respectively. Multiple cellular metabolic pathways altered by Fragment 1 in the representative Gram-negative V. parahaemolyticus ATCC17802 and Gram-positive S. aureus ATCC25923 pathogens after treatment with Fragment 1 (1× MIC) for 6 h (p < 0.05). These results indicated that the energy supply and protein translation of the tested strains was inhibited, the signal transduction was blocked, and the ability to pump foreign harmful substances was reduced, leading to cell death. Overall, the results of this study demonstrate that Fragment 1 from P. kleiniana Wight et Arn is a promising candidate for antibacterial medicine and food preservatives.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/foods12081640/s1, Table S1: The oligonucleotide primers designed and used in the RT-qPCR assay; Table S2: The relative expression of representative DEGs by the RT-qPCR assay; Table S3: The bacterial strains and media used in this study; Figure S1: The Prep−HPLC diagram of purifying the methanol-phase crude extract from P. kleiniana Wight et Arn.

Author Contributions

Y.T.: major experiments, data curation, and writing—original draft; P.Y.: writing—review and editing; L.C.: funding acquisition, conceptualization, and writing—review and editing. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by Shanghai Municipal Science and Technology Commission, grant number 17050502200, and National Natural Science Foundation of China, grant number 31671946.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data is contained within the article or Supplementary Materials. The complete lists of DEGs in the two strains are available in the NCBI SRA database (https://submit.ncbi.nlm.nih.gov/subs/bioproject/, accessed on 29 November 2022) under the accession number PRJNA906658.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Bueno, E.; Pinedo, V.; Cava, F. Adaptation of Vibrio cholerae to hypoxic environments. Front. Microbiol. 2020, 11, 739. [Google Scholar] [CrossRef]
  2. Stocco, G.; Lucafò, M.; Decorti, G. Pharmacogenomics of antibiotics. Int. J. Mol. Sci. 2020, 21, 5975. [Google Scholar] [CrossRef]
  3. Bombaywala, S.; Mandpe, A.; Paliya, S.; Kumar, S. Antibiotic resistance in the environment: A critical insight on its occurrence, fate, and eco-toxicity. Environ. Sci. Pollut. Res. Int. 2021, 28, 24889–24916. [Google Scholar] [CrossRef]
  4. Mancuso, G.; Midiri, A.; Gerace, E.; Biondo, C. Bacterial antibiotic resistance: The most critical pathogens. Pathogens 2021, 10, 1310. [Google Scholar] [CrossRef] [PubMed]
  5. Gomes, F.; Rodrigues, M.E.; Martins, N.; Ferreira, I.C.F.R.; Henriques, M. Phenolic plant extracts versus penicillin G: In vitro susceptibility of Staphylococcus aureus isolated from bovine mastitis. Pharmaceuticals 2019, 12, 128. [Google Scholar] [CrossRef] [Green Version]
  6. Thomford, N.E.; Senthebane, D.A.; Rowe, A.; Munro, D.; Seele, P.; Maroyi, A.; Dzobo, K. Natural products for drug discovery in the 21st century: Innovations for novel drug discovery. Int. J. Mol. Sci. 2018, 19, 1578. [Google Scholar] [CrossRef] [Green Version]
  7. Chandra, G.; Mukherjee, D.; Ray, A.S.; Chatterjee, S.; Bhattacharjee, I. Phytoextracts as antibacterials: A review. Curr. Drug Discov. Technol. 2020, 17, 523–533. [Google Scholar] [CrossRef] [PubMed]
  8. Volis, S. Securing a future for China’s plant biodiversity through an integrated conservation approach. Plant Divers. 2018, 40, 91–105. [Google Scholar] [CrossRef] [PubMed]
  9. Tao, J.; Yan, S.; Zhou, C.; Liu, Q.; Zhu, H.; Wen, Z. Total flavonoids from Potentilla kleiniana Wight et Arn inhibits biofilm formation and virulence factors production in methicillin-resistant Staphylococcus aureus (MRSA). J. Ethnopharmacol. 2021, 279, 114383. [Google Scholar] [CrossRef]
  10. Zhou, Y.Q.; Li, S.M.; Wei, X.; Yang, X.; Xiao, J.W.; Pan, B.W.; Xie, S.X.; Zhou, Y.; Yang, J.; Wei, Y. Identification and quantitative analysis of bioactive components from Potentilla kleiniana Wight et Arn with anti HIV-1 proteases activity. Nat. Prod. Res. 2022, 1–4. [Google Scholar] [CrossRef] [PubMed]
  11. Liu, M.; Huang, X.; Liu, Q.; Li, X.; Chen, M.; Zhu, Y.; Chen, X. Separation of α-glucosidase inhibitors from Potentilla kleiniana Wight et Arn using solvent and flow-rate gradient high-speed counter-current chromatography target-guided by ultrafiltration HPLC-MS screening. Phytochem. Anal. 2019, 30, 661–668. [Google Scholar] [CrossRef]
  12. Li, S.; Tan, J.; Zeng, J.; Wu, X.W.X.; Zhang, J. Antihyperglycemic and antioxidant effect of the total flavones of Potentilla kleiniana Wight et Arn. in streptozotocin induced diabetic rats. Pak. J. Pharm. Sci. 2017, 30, 171–178. [Google Scholar]
  13. Tao, J.; Yan, S.; Wang, H.; Zhao, L.; Zhu, H.; Wen, Z. Antimicrobial and antibiofilm effects of total flavonoids from Potentilla kleiniana Wight et Arn on Pseudomonas aeruginosa and its potential application to stainless steel surfaces. LWT-Food Sci. Technol. 2022, 154, 112631. [Google Scholar] [CrossRef]
  14. Xuan, S.H.; Hong, I.K.; Lee, Y.J.; Kim, J.W.; Park, S.N. Biological activities and chemical components of Potentilla kleiniana Wight & Arn. Nat. Prod. Res. 2020, 34, 3262–3266. [Google Scholar] [CrossRef] [PubMed]
  15. Liu, Y.; Yang, L.; Liu, P.; Jin, Y.; Qin, S.; Chen, L. Identification of antibacterial components in the methanol-phase extract from edible herbaceous plant Rumex madaio Makino and their antibacterial action modes. Molecules 2022, 27, 660. [Google Scholar] [CrossRef]
  16. Fu, J.; Wang, Y.; Sun, M.; Xu, Y.; Chen, L. Antibacterial activity and components of the methanol-phase extract from rhizomes of pharmacophagous plant Alpinia officinarum Hance. Molecules 2022, 27, 4308. [Google Scholar] [CrossRef]
  17. Wang, J.; Ding, Q.; Yang, Q.; Fan, H.; Yu, G.; Liu, F.; Bello, B.K.; Zhang, X.; Zhang, T.; Dong, J.; et al. Vibrio alginolyticus triggers inflammatory response in mouse peritoneal macrophages via activation of NLRP3 inflammasome. Front. Cell. Infect. Microbiol. 2021, 11, 769777. [Google Scholar] [CrossRef] [PubMed]
  18. Hernández-Robles, M.F.; Natividad-Bonifacio, I.; Álvarez-Contreras, A.K.; Tercero-Alburo, J.J.; Quiñones-Ramírez, E.I.; Vázquez-Salinas, C. Characterization of potential virulence factors of Vibrio mimicus isolated from fishery products and water. Int. J. Microbiol. 2021, 2021, 8397930. [Google Scholar] [CrossRef]
  19. Dey, R.; Rieger, A.M.; Stephens, C.; Ashbolt, N.J. Interactions of Pseudomonas aeruginosa with Acanthamoeba polyphaga observed by imaging flow cytometry. Cytom. Part A 2019, 95, 555–564. [Google Scholar] [CrossRef]
  20. Liu, J.; Qin, K.; Wu, C.; Fu, K.; Yu, X.; Zhou, L. De Novo sequencing provides insights into the pathogenicity of foodborne Vibrio parahaemolyticus. Front. Cell. Infect. Microbiol. 2021, 11, 652957. [Google Scholar] [CrossRef] [PubMed]
  21. Huang, Y.; Flint, S.H.; Palmer, J.S. Bacillus cereus spores and toxins—The potential role of biofilms. Food Microbiol. 2020, 90, 103493. [Google Scholar] [CrossRef]
  22. Ojha, R.; Dittmar, A.A.; Severin, G.B.; Koestler, B.J. Shigella flexneri diguanylate cyclases regulate virulence. J. Bacteriol. 2021, 203, e0024221. [Google Scholar] [CrossRef] [PubMed]
  23. Li, H.; Tang, T.; Stegger, M.; Dalsgaard, A.; Liu, T.; Leisner, J.J. Characterization of antimicrobial-resistant Staphylococcus aureus from retail foods in Beijing, China. Food Microbiol. 2021, 93, 103603. [Google Scholar] [CrossRef] [PubMed]
  24. Guo, Y.; Song, G.; Sun, M.; Wang, J.; Wang, Y. Prevalence and therapies of antibiotic-resistance in Staphylococcus aureus. Front. Cell. Infect. Microbiol. 2020, 10, 107. [Google Scholar] [CrossRef] [Green Version]
  25. Danchik, C.; Casadevall, A. Role of cell surface hydrophobicity in the pathogenesis of medically-significant fungi. Front. Cell. Infect. Microbiol. 2020, 10, 594973. [Google Scholar] [CrossRef] [PubMed]
  26. Soundharrajan, I.; Kim, D.; Kuppusamy, P.; Muthusamy, K.; Lee, H.J.; Choi, K.C. Probiotic and triticale silage fermentation potential of Pediococcus pentosaceus and Lactobacillus brevis and their impacts on pathogenic bacteria. Microorganisms 2019, 7, 318. [Google Scholar] [CrossRef] [Green Version]
  27. Yu, X.; Sha, L.; Liu, Q.; Zhao, Y.; Fang, H.; Cao, Y.; Zhao, J. Recent advances in cell membrane camouflage-based biosensing application. Biosens. Bioelectron. 2021, 194, 113623. [Google Scholar] [CrossRef]
  28. Zhang, M.; Yu, Y.; Lian, L.; Li, W.; Ren, J.; Liang, Y.; Xue, F.; Tang, F.; Zhu, X.; Ling, J.; et al. Functional mechanism of antimicrobial peptide bomidin and its safety for Macrobrachium rosenbergii. Probiotics Antimicrob. Proteins 2022, 14, 169–179. [Google Scholar] [CrossRef]
  29. Bojkovic, J.; Richie, D.L.; Six, D.A.; Rath, C.M.; Sawyer, W.S.; Hu, Q.; Dean, C.R. Characterization of an acinetobacter baumannii lptD deletion strain: Permeability defects and response to inhibition of lipopolysaccharide and fatty acid biosynthesis. J. Bacteriol. 2015, 198, 731–741. [Google Scholar] [CrossRef] [Green Version]
  30. Veiko, A.G.; Olchowik-Grabarek, E.; Sekowski, S.; Roszkowska, A.; Lapshina, E.A.; Dobrzynska, I.; Zamaraeva, M.; Zavodnik, I.B. Antimicrobial activity of quercetin, naringenin and catechin: Flavonoids inhibit Staphylococcus aureus-induced hemolysis and modify membranes of bacteria and erythrocytes. Molecules 2023, 28, 1252. [Google Scholar] [CrossRef]
  31. Zdybicka-Barabas, A.; Stączek, S.; Pawlikowska-Pawlęga, B.; Mak, P.; Luchowski, R.; Skrzypiec, K.; Mendyk, E.; Wydrych, J.; Gruszecki, W.I.; Cytryńska, M. Studies on the interactions of neutral Galleria mellonella cecropin D with living bacterial cells. Amino Acids 2019, 51, 175–191. [Google Scholar] [CrossRef] [Green Version]
  32. Mizzi, L.; Maniscalco, D.; Gaspari, S.; Chatzitzika, C.; Gatt, R.; Valdramidis, V.P. Assessing the individual microbial inhibitory capacity of different sugars against pathogens commonly found in food systems. Lett. Appl. Microbiol. 2020, 71, 251–258. [Google Scholar] [CrossRef]
  33. Fan, X.; Kong, D.; He, S.; Chen, J.; Jiang, Y.; Ma, Z.; Feng, J.; Yan, H. Phenanthrene derivatives from asarum heterotropoides showed excellent antibacterial activity against phytopathogenic bacteria. J. Agric. Food Chem. 2021, 69, 14520–14529. [Google Scholar] [CrossRef]
  34. Tan, Z.; Deng, J.; Ye, Q.; Zhang, Z. The antibacterial activity of natural-derived flavonoids. Curr. Top. Med. Chem. 2022, 22, 1009–1019. [Google Scholar] [CrossRef]
  35. Kachur, K.; Suntres, Z. The antibacterial properties of phenolic isomers, carvacrol and thymol. Crit. Rev. Food Sci. Nutr. 2020, 60, 3042–3053. [Google Scholar] [CrossRef]
  36. Wibowo, J.T.; Ahmadi, P.; Rahmawati, S.I.; Bayu, A.; Putra, M.Y.; Kijjoa, A. marine-derived indole alkaloids and their biological and pharmacological activities. Mar. Drugs 2021, 20, 3. [Google Scholar] [CrossRef]
  37. Li, H.; Li, Y.; Wang, Y.; Liu, L.; Dong, H.; Satoh, T. Membrane-active amino acid-coupled polyetheramine derivatives with high selectivity and broad-spectrum antibacterial activity. Acta Biomater. 2022, 142, 136–148. [Google Scholar] [CrossRef] [PubMed]
  38. Sharma, P.; Maklashina, E.; Cecchini, G.; Iverson, T.M. The roles of SDHAF2 and dicarboxylate in covalent flavinylation of SDHA, the human complex II flavoprotein. Proc. Natl. Acad. Sci. USA 2020, 117, 23548–23556. [Google Scholar] [CrossRef] [PubMed]
  39. Dolan, S.K.; Wijaya, A.; Kohlstedt, M.; Gläser, L.; Brear, P.; Silva-Rocha, R.; Wittmann, C.; Welch, M. Systems-wide dissection of organic acid assimilation in Pseudomonas aeruginosa Reveals a Novel Path to Underground Metabolism. Mbio 2022, 13, e0254122. [Google Scholar] [CrossRef]
  40. Park, C.; Shin, B.; Park, W. Alternative fate of glyoxylate during acetate and hexadecane metabolism in Acinetobacter oleivorans DR1. Sci. Rep. 2019, 9, 14402. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  41. Durall, C.; Kukil, K.; Hawkes, J.A.; Albergati, A.; Lindblad, P.; Lindberg, P. Production of succinate by engineered strains of Synechocystis PCC 6803 overexpressing phosphoenolpyruvate carboxylase and a glyoxylate shunt. Microb. Cell Factories 2021, 20, 39. [Google Scholar] [CrossRef]
  42. Zhang, H.; Liang, Z.; Zhao, M.; Ma, Y.; Luo, Z.; Li, S.; Xu, H. Metabolic engineering of Escherichia coli for ectoine production with a fermentation strategy of supplementing the amino donor. Front. Bioeng. Biotechnol. 2022, 10, 824859. [Google Scholar] [CrossRef] [PubMed]
  43. Yun, M.K.; Nourse, A.; White, S.W.; Rock, C.O.; Heath, R.J. Crystal structure and allosteric regulation of the cytoplasmic Escherichia coli L-asparaginase I. J. Mol. Biol. 2007, 369, 794–811. [Google Scholar] [CrossRef] [Green Version]
  44. Wang, Z.; Bie, P.; Cheng, J.; Lu, L.; Cui, B.; Wu, Q. The ABC transporter YejABEF is required for resistance to antimicrobial peptides and the virulence of Brucella melitensis. Sci. Rep. 2016, 6, 31876. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Huang, L.; Huang, L.; Zhao, L.; Qin, Y.; Su, Y.; Yan, Q. The regulation of oxidative phosphorylation pathway on Vibrio alginolyticus adhesion under adversities. Microbiologyopen 2019, 8, e00805. [Google Scholar] [CrossRef] [Green Version]
  46. Santucci, R.; Sinibaldi, F.; Cozza, P.; Polticelli, F.; Fiorucci, L. Cytochrome c: An extreme multifunctional protein with a key role in cell fate. Int. J. Biol. Macromol. 2019, 136, 1237–1246. [Google Scholar] [CrossRef]
  47. Maza-Márquez, P.; Lee, M.D.; Detweiler, A.M.; Bebout, B.M. Millimeter-scale vertical partitioning of nitrogen cycling in hypersaline mats reveals prominence of genes encoding multi-heme and prismane proteins. ISME J. 2022, 16, 1119–1129. [Google Scholar] [CrossRef] [PubMed]
  48. Tierney, A.R.; Rather, P.N. Roles of two-component regulatory systems in antibiotic resistance. Future Microbiol. 2019, 14, 533–552. [Google Scholar] [CrossRef]
  49. Liu, F.; Yao, Q.; Huang, J.; Wan, J.; Xie, T.; Gao, X.; Sun, D.; Zhang, F.; Bei, W.; Lei, L. The two-component system CpxA/CpxR is critical for full virulence in Actinobacillus pleuropneumoniae. Front. Microbiol. 2022, 13, 1029426. [Google Scholar] [CrossRef]
  50. Hu, M.; Huang, X.; Xu, X.; Zhang, Z.; He, S.; Zhu, J.; Liu, H.; Shi, X. Characterization of the role of two-component systems in antibiotic resistance formation in Salmonella enterica Serovar Enteritidis. mSphere 2022, 7, e0038322. [Google Scholar] [CrossRef]
  51. Blair, J.M.A.; Siasat, P.; McNeil, H.E.; Colclough, A.; Ricci, V.; Lawler, A.J.; Abdalaal, H.; Buckner, M.M.C.; Baylay, A.; Busby, S.J.; et al. EnvR is a potent repressor of acrAB transcription in Salmonella. J. Antimicrob. Chemother. 2022, 78, 133–140. [Google Scholar] [CrossRef]
  52. Oberlies, J.; Watzl, C.; Giese, T.; Luckner, C.; Kropf, P.; Müller, I. Regulation of NK cell function by human granulocyte arginase. J. Immunol. 2009, 182, 5259–5267. [Google Scholar] [CrossRef] [Green Version]
  53. Kim, S.H.; Sierra, R.A.; McGee, D.J.; Zabaleta, J. Transcriptional profiling of gastric epithelial cells infected with wild type or arginase-deficient Helicobacter pylori. BMC Microbiol. 2012, 12, 175. [Google Scholar] [CrossRef] [Green Version]
  54. Yurgel, S.N.; Johnson, S.A.; Rice, J.; Sa, N.; Bailes, C.; Baumgartner, J.; Pitzer, J.E.; Roop, R.M., II; Roje, S. A novel formamidase is required for riboflavin biosynthesis in invasive bacteria. J. Biol. Chem. 2022, 298, 102377. [Google Scholar] [CrossRef]
  55. Lai, R.Z.; Parkinson, J.S. Monitoring two-component sensor kinases with a chemotaxis signal readout. Methods Mol. Biol. 2018, 1729, 127–135. [Google Scholar] [CrossRef]
  56. Tan, L.; Huang, Y.; Shang, W.; Yang, Y.; Peng, H.; Hu, Z.; Wang, Y.; Rao, Y.; Hu, Q.; Rao, X.; et al. Accessory gene regulator (agr) allelic variants in cognate Staphylococcus aureus strain display similar phenotypes. Front. Microbiol. 2022, 13, 700894. [Google Scholar] [CrossRef]
  57. Párraga Solórzano, P.K.; Shupe, A.C.; Kehl-Fie, T.E. The sensor histidine kinase ArlS is necessary for Staphylococcus aureus to activate ArlR in response to nutrient availability. J. Bacteriol. 2021, 203, e0042221. [Google Scholar] [CrossRef]
  58. Loiseau, L.; Vergnes, A.; Ezraty, B. Methionine oxidation under anaerobic conditions in Escherichia coli. Mol. Microbiol. 2022, 118, 387–402. [Google Scholar] [CrossRef]
  59. Fang, J.; Yan, L.; Tan, M.; Li, G.; Liang, Y.; Li, K. Nitrogen removal characteristics of a marine denitrifying Pseudomonas stutzeri BBW831 and a simplified strategy for improving the denitrification performance under stressful conditions. Mar. Biotechnol. 2023, 25, 109–122. [Google Scholar] [CrossRef]
  60. Alvarez, L.; Sanchez-Hevia, D.; Sánchez, M.; Berenguer, J. A new family of nitrate/nitrite transporters involved in denitrification. Int. Microbiol. 2019, 22, 19–28. [Google Scholar] [CrossRef] [Green Version]
  61. Harborne, N.R.; Griffiths, L.; Busby, S.J.; Cole, J.A. Transcriptional control, translation and function of the products of the five open reading frames of the Escherichia coli nir operon. Mol. Microbiol. 1992, 6, 2805–2813. [Google Scholar] [CrossRef]
  62. Léger, L.; Byrne, D.; Guiraud, P.; Germain, E.; Maisonneuve, E. NirD curtails the stringent response by inhibiting RelA activity in Escherichia coli. Elife 2021, 10, e64092. [Google Scholar] [CrossRef] [PubMed]
  63. Kanagarajan, S.; Dhamodharan, P.; Mutharasappan, N.; Choubey, S.K.; Jayaprakash, P.; Biswal, J.; Jeyaraman, J. Structural insights on binding mechanism of CAD complexes (CPSase, ATCase and DHOase). J. Biomol. Struct. Dyn. 2021, 39, 3144–3157. [Google Scholar] [CrossRef] [PubMed]
  64. Garavito, M.F.; Narváez-Ortiz, H.Y.; Zimmermann, B.H. Pyrimidine metabolism: Dynamic and versatile pathways in pathogens and cellular development. J. Genet. Genom. 2015, 42, 195–205. [Google Scholar] [CrossRef] [PubMed]
  65. Xu, M.; Fu, H.; Chen, D.; Shao, Z.; Zhu, J.; Alali, W.Q.; Chen, L. Simple visualized detection method of virulence-associated genes of Vibrio cholerae by loop-mediated isothermal amplification. Front. Microbiol. 2019, 10, 2899. [Google Scholar] [CrossRef]
  66. Wang, Y.; Chen, L.; Pandak, W.M.; Heuman, D.; Hylemon, P.B.; Ren, S. High Glucose Induces Lipid Accumulation via 25-Hydroxycholesterol DNA-CpG Methylation. iScience 2020, 23, 101102. [Google Scholar] [CrossRef]
  67. Shan, X.; Fu, J.; Li, X.; Peng, X.; Chen, L. Comparative proteomics and secretomics revealed virulence, and coresistance-related factors in non O1/O139 Vibrio cholerae recovered from 16 species of consumable aquatic animals. J. Proteom. 2022, 251, 104408. [Google Scholar] [CrossRef]
  68. Cui, J.; Hölzl, G.; Karmainski, T.; Tiso, T.; Kubicki, S.; Thies, S.; Blank, L.M.; Jaeger, K.E.; Dörmann, P. The glycine-glucolipid of Alcanivorax borkumensis is resident to the bacterial cell wall. Appl. Environ. Microbiol. 2022, 88, e0112622. [Google Scholar] [CrossRef]
  69. Kuhry, J.G.; Duportail, G.; Bronner, C.; Laustriat, G. Plasma membrane fluidity measurements on whole living cells by fluorescence anisotropy of trimethylammoniumdiphenylhexatriene. Biochim. Biophys. Acta. 1985, 845, 60–67. [Google Scholar] [CrossRef]
  70. Wang, Z.; Qin, Q.; Zheng, Y.; Li, F.; Zhao, Y.; Chen, G.Q. Engineering the permeability of Halomonas bluephagenesis enhanced its chassis properties. Metab. Eng. 2021, 67, 53–66. [Google Scholar] [CrossRef]
  71. Huang, B.; Liu, X.; Li, Z.; Zheng, Y.; Wai Kwok Yeung, K.; Cui, Z.; Liang, Y.; Zhu, S.; Wu, S. Rapid bacteria capturing and killing by AgNPs/N-CD@ZnO hybrids strengthened photo-responsive xerogel for rapid healing of bacteria-infected wounds. Chem. Eng. J. 2021, 414, 128805. [Google Scholar] [CrossRef]
  72. Yang, L.; Wang, Y.; Yu, P.; Ren, S.; Zhu, Z.; Jin, Y.; Yan, J.; Peng, X.; Chen, L. Prophage-related gene VpaChn25_0724 contributes to cell membrane integrity and growth of Vibrio parahaemolyticus CHN25. Front. Cell. Infect. Microbiol. 2020, 10, 595709. [Google Scholar] [CrossRef]
Figure 1. Inhibition activity of the methanol-phase crude extract from P. kleiniana Wight et Arn against the four representative bacterial strains. (A-1D-1) V. parahemolyticus B5-29, V. parahemolyticus ATCC17802, S. aureus ATCC25923, and S. aureus ATCC8095, respectively. (A-2D-2) corresponding negative controls, respectively.
Figure 1. Inhibition activity of the methanol-phase crude extract from P. kleiniana Wight et Arn against the four representative bacterial strains. (A-1D-1) V. parahemolyticus B5-29, V. parahemolyticus ATCC17802, S. aureus ATCC25923, and S. aureus ATCC8095, respectively. (A-2D-2) corresponding negative controls, respectively.
Foods 12 01640 g001
Figure 2. Effects of Fragment 1 (1× MIC) from P. kleiniana Wight et Arn on cell surface hydrophobicity, membrane fluidity and outer membrane permeability of the four bacterial strains. (AC) cell surface hydrophobicity, membrane fluidity, and outer membrane permeability, respectively. *: p < 0.05; **: p < 0.01; and ***: p < 0.001.
Figure 2. Effects of Fragment 1 (1× MIC) from P. kleiniana Wight et Arn on cell surface hydrophobicity, membrane fluidity and outer membrane permeability of the four bacterial strains. (AC) cell surface hydrophobicity, membrane fluidity, and outer membrane permeability, respectively. *: p < 0.05; **: p < 0.01; and ***: p < 0.001.
Foods 12 01640 g002
Figure 3. Effects of Fragment 1 (1× MIC) from P. kleiniana Wight et Arn on the bacterial inner cell membrane permeability. (AD) V. parahaemolyticus B5-29, V. parahaemolyticus ATCC17802, S. aureus ATCC25923, and S. aureus ATCC8095, respectively. The treatment groups were overall significantly different from the control groups (p < 0.05), except the S. aureus ATCC8095 group treated for 2 h (D).
Figure 3. Effects of Fragment 1 (1× MIC) from P. kleiniana Wight et Arn on the bacterial inner cell membrane permeability. (AD) V. parahaemolyticus B5-29, V. parahaemolyticus ATCC17802, S. aureus ATCC25923, and S. aureus ATCC8095, respectively. The treatment groups were overall significantly different from the control groups (p < 0.05), except the S. aureus ATCC8095 group treated for 2 h (D).
Foods 12 01640 g003
Figure 4. The SEM observation of cell surface structure of the two bacterial strains treated with the 1× MIC of Fragment 1 for different times. (A): V. parahaemolyticus ATCC17802; (B): S. aureus ATCC 25923.
Figure 4. The SEM observation of cell surface structure of the two bacterial strains treated with the 1× MIC of Fragment 1 for different times. (A): V. parahaemolyticus ATCC17802; (B): S. aureus ATCC 25923.
Foods 12 01640 g004
Figure 5. The major changed metabolic pathways in V. parahaemolyticus ATCC 17802 mediated by Fragment 1 from P. kleiniana Wight et Arn. (A) The Volcano plot of the DEGs. (B) The significantly altered metabolic pathways in the bacterium. Different colors represented significant levels of the enriched genes.
Figure 5. The major changed metabolic pathways in V. parahaemolyticus ATCC 17802 mediated by Fragment 1 from P. kleiniana Wight et Arn. (A) The Volcano plot of the DEGs. (B) The significantly altered metabolic pathways in the bacterium. Different colors represented significant levels of the enriched genes.
Foods 12 01640 g005
Figure 6. The major changed metabolic pathways in S. aureus ATCC25923 triggered by Fragment 1 from P. kleiniana Wight et Arn. (A) The Volcano plot of the DGEs. (B) The significantly altered metabolic pathways in the bacterium.
Figure 6. The major changed metabolic pathways in S. aureus ATCC25923 triggered by Fragment 1 from P. kleiniana Wight et Arn. (A) The Volcano plot of the DGEs. (B) The significantly altered metabolic pathways in the bacterium.
Foods 12 01640 g006
Table 1. Antibacterial activity of crude extracts from P. kleiniana Wight et Arn.
Table 1. Antibacterial activity of crude extracts from P. kleiniana Wight et Arn.
StrainInhibition Zone (Diameter, mm)MIC (mg/mL)
CPEMPECPEMPE
Aeromonas hydrophila ATCC35654----
Bacillus cereus A1-17.03 ± 0.0110.54 ± 0.48506.25
Bacillus cereus A2-27.11 ± 0.0210.54 ± 0.75501.56
Enterobacter cloacae ATCC130477.00 ± 0.117.11 ± 0.265050
Enterobacter cloacae C1-1----
Escherichia coli ATCC8739-7.62 ± 0.37-25
Escherichia coli ATCC25922----
Escherichia coli K12-7.51 ± 0.29-25
Enterobacter sakazakii CMCC45401----
Klebsiella pneumoniae 8-2-10-8----
Klebsiella pneumoniae 8-2-1-14----
Pseudomonas aeruginosa ATCC9027-10.51 ± 0.41-6.25
Pseudomonas aeruginosa ATCC27853-8.14 ± 0.32-25
Salmonella enterica subsp. enterica (ex Kauffmann and Edwards) Le Minor and Popoff serovar Choleraesuis ATCC13312----
Salmonella paratyphi-A CMCC50093----
Salmonellaenterica subsp. enterica (ex Kauffmann and Edwards) Le Minor and Popoff serovar Vellore ATCC156117.09 ± 0.0910.11 ± 0.61506.25
Salmonella E1-1----
Shigella dysenteriae CMCC512527.02 ± 0.119.29 ± 0.515012.5
Shigella flexneri CMCC515727.82 ± 0.2010.17 ± 0.20256.25
Shigella flexneri ATCC12022----
Shigella flexneri CMCC51574-9.17 ± 0.21-12.5
Shigella sonnei ATCC259317.00 ± 0.118.19 ± 0.515025
Shigella sonnet CMCC51592----
Staphylococcus aureus ATCC259237.03 ± 0.149.41 ± 0.275012.5
Staphylococcus aureus ATCC80957.07 ± 0.1510.15 ± 0.24506.25
Staphylococcus aureus ATCC292137.78 ± 0.109.21 ± 0.012512.5
Staphylococcus aureus ATCC65387.62 ± 0.619.55 ± 0.372512.5
Staphylococcus aureus D1-17.11 ± 0.257.00 ± 0.515050
Vibrio alginolyticus ATCC17749-10.11 ± 0.24-3.13
Vibrio alginolyticus ATCC33787----
Vibrio cholerae GIM1.449-7.00 ± 0.14-50
Vibrio cholerae Q10-547.22 ± 0.107.02 ± 0.215050
Vibrio fluvialis ATCC33809-7.12 ± 0.03-50
Vibrio harvey ATCC BAA-1117----
Vibrio harveyi ATCC33842----
Vibrio mimicus bio-567597.21 ± 0.4111.00 ± 0.32253.13
Vibrio parahemolyticus ATCC17802-10.67 ± 1.21-1.56
Vibrio parahemolyticus ATCC338478.63 ± 0.247.14 ± 0.1212.550
Vibrio parahemolyticus B3-137.17 ± 0.2912.33 ± 0.65503.13
Vibrio parahemolyticus B4-10-11.26 ± 0.34-6.25
Vibrio parahemolyticus B5-297.17 ± 0.0413.77 ± 0.85503.13
Vibrio parahemolyticus B9-357.20 ± 0.0913.15 ± 0.44253.13
Vibrio parahemolyticus A1-17.13 ± 0.1510.35 ± 0.58503.13
Vibrio vulnificus ATCC275627.65 ± 0.447.01 ± 0.232550
Note: CPE: chloroform-phase extract. MPE: methanol-phase extract. -: no bacteriostasis activity. Inhibition zone: diameter includes the disk diameter (6 mm). MIC: minimum inhibitory concentration. Values were means ± standard deviation (S.D.) of three parallel measurements.
Table 2. Antibacterial activity of Fragment 1 of the methanol-phase extract from P. kleiniana Wight et Arn.
Table 2. Antibacterial activity of Fragment 1 of the methanol-phase extract from P. kleiniana Wight et Arn.
StrainInhibition Zone (Diameter, mm)MIC (mg/mL)
B. cereus A2-28.03 ± 0.456.25
S. flexneri CMCC515727.50 ± 0.506.25
S. aureus ATCC259238.03 ± 0.4012.5
S. aureus ATCC80959.53 ± 0.356.25
S. aureus ATCC65387.10 ± 0.3650.0
V. parahemolyticus ATCC1780210.31 ± 0.626.25
V. parahemolyticus A1-18.57 ± 0.6025.0
V. parahemolyticus B3-1310.37 ± 0.326.25
V. parahemolyticus B4-1010.30 ± 0.5012.5
V. parahemolyticus B5-2911.30 ± 0.266.25
V. parahemolyticus B9-3511.27 ± 0.4012.5
Table 3. Compounds identified in Fragment 1 from P. kleiniana Wight et Arn via UHPLC–MS analysis.
Table 3. Compounds identified in Fragment 1 from P. kleiniana Wight et Arn via UHPLC–MS analysis.
Peak
No.
Identified CompoundCompound NatureRt (min)FormulaExact MassPeak Area (%)
1D-MaltoseCarbohydrates0.76C12H22O11342.11626.77%
2OxymorphonePhenanthrenes and derivatives11.18C17H19NO4301.13146.29%
3RutinFlavonoids12.99C27H30O16281.08996.29%
4D-ProlineAmino acid and derivatives0.76C5H9NO2115.06335.41%
5L-ProlineAmino acid and derivatives0.73C5H9NO2115.06335.41%
6L-Glutamic acidAmino acid and derivatives0.66C5H9NO4147.05325.20%
7SucroseCarbohydrates0.89C12H22O11342.11623.62%
8CynarosideFlavonoids12.98C21H20O11282.1623.37%
9PiperlonguminineAlkaloids10.57C16H19NO3273.13653.21%
105-Aminovaleric acidAmino acid and derivatives1.11C5H11NO2117.0793.12%
11D-GlutamineCarboxylic acids and derivatives0.66C5H10N2O3146.06912.99%
12L-LysineAmino acid and derivatives0.64C6H14N2O2146.10552.99%
13p-OctopaminePhenols3.84C8H11NO2153.0792.96%
14Oleic acidFatty acyls13.03C18H34O2282.25592.91%
15IsoquercitrinFlavonoids10.58C21H20O12274.19332.44%
16L-Pipecolic acidAmino acid and derivatives0.69C6H11NO2129.0792.31%
17Moracin CPhenols0.67C19H18O4129.04262.31%
18KojibioseFatty acyls0.72C12H22O11342.11622.22%
19Gluconic acidCarbohydrates0.69C6H12O7196.05831.97%
20BetaineAlkaloids1.06C5H11NO2117.0791.51%
21L-ValineAmino acid and derivatives0.93C5H11NO2117.0791.49%
22D-alpha-Aminobutyric acidCarboxylic acids and derivatives0.65C4H9NO2103.06331.46%
23cis-Aconitic acidOrganic acids and derivatives1.46C6H6O6174.01641.34%
24LactuloseOrganooxygen compounds0.77C12H22O11342.11621.33%
25TuranoseFatty acyls0.79C12H22O11342.11621.33%
26L-Pipecolic acidAmino acid and derivatives1.47C6H11NO2129.0791.15%
27DL-NorvalineAmino acid and derivatives1.05C5H11NO2117.0791.11%
28L-AsparagineAmino acid and derivatives0.64C4H8N2O3132.05351.11%
29Malic acidHydroxy acids and derivatives0.8C4H6O5134.02150.90%
30TrigonellineAlkaloids0.82C7H7NO2137.04770.90%
31AcetamideAlkaloids13.95C2H5NO59.037110.88%
32Beta-D-fructose 2-phosphateOrganooxygen compounds0.75C6H13O9P260.02970.77%
3322-DehydroclerosterolSteroids12.59C29H46O410.35490.76%
34ArtemisininSesquiterpenoids13.02C15H22O5282.14670.72%
35Kaempferol-3-O-rutinosideflavonoids6.29C27H30O15594.15850.54%
36L-HomoserineAmino acid and derivatives0.67C4H9NO3119.05820.52%
37L-ThreonineAmino acid and derivatives0.64C4H9NO3119.05820.50%
38Palmitic acidLipids12.92C16H32O2256.24020.49%
39O-AcetylethanolamineAlkaloids0.67C4H9NO2103.06330.46%
40Galactose 1-phosphateOrganooxygen compounds0.65C6H13O9P260.02970.46%
41Glucose 1-phosphateOrganooxygen compounds13C6H13O9P260.02970.45%
42Adenosine 5′-monophosphateNucleotide and its derivates1.38C10H14N5O7P347.06310.43%
43L-ArginineAmino acid and derivatives0.6C6H14N4O2174.11170.43%
44MaltotrioseOrganooxygen compounds1.23C18H32O16504.1690.40%
45IndoleAlkaloids3.82C8H7N117.05780.38%
46D-Glucose 6-phosphateCarbohydrates0.65C6H13O9P260.02970.37%
47D-Aspartic acidAlkaloids0.76C4H7NO4133.03750.36%
48Vitexin rhamnosideFlavonoids6.78C27H30O14578.16360.35%
49L-Aspartic acidAmino acid and derivatives0.63C4H7NO4133.03750.33%
50MaltolFlavonoids0.9C6H6O3126.03170.33%
51AstragalinFlavonoids6.52C21H20O11448.10060.32%
523-Hydroxy-3-methylpentane-1,5-dioic acidAmino acid and derivatives2.32C6H10O5162.05280.31%
53CampesterolSteroids and steroid derivatives12.18C28H48O400.37050.30%
54L-OrnithineAmino acid and derivatives0.55C5H12N2O2132.08990.30%
55AdenosineNucleotide and its derivates2.58C10H13N5O4267.09680.29%
56VidarabinePurine nucleosides2.28C10H13N5O4267.09680.27%
57Nicotinic acidNicotinic acid derivatives0.73C6H5NO2123.0320.27%
58Pelargonidin-3-O-glucosideFlavonoids1.11C21H20O10100.05240.26%
59L-CitrulineAmino acid and derivatives0.66C6H13N3O3175.09570.26%
60Diallyl disulfideMiscellaneous0.68C6H10S2146.02240.26%
61SarracineAlkaloids13.14C18H27NO5337.18890.22%
62N-AcetylputrescinePhenolamides1.79C6H14N2O130.11060.22%
63Salicylic acidOrganic acid7.06C7H6O3138.03170.22%
645-MethylcytosineNucleotide and its derivates2.26C5H7N3O125.05890.21%
65Ellagic acidPhenols6.12C14H6O8302.00630.21%
66IsodiospyrinQuinones11.28C22H14O6374.0790.21%
Table 4. The major altered metabolic pathways in V. parahaemolyticus ATCC17802.
Table 4. The major altered metabolic pathways in V. parahaemolyticus ATCC17802.
Metabolic PathwayGene IDGene NameFold ChangeGene Description
Citrate cycleWU75_19785sucA0.1462-oxoglutarate dehydrogenase
WU75_07425pckA0.465Phosphoenolpyruvate carboxykinase
WU75_19790sucB0.133Dihydrolipoamide succinyltransferase
WU75_11550acnB0.143Bifunctional aconitate hydratase 2/2-methylisocitrate dehydratase
WU75_19795sucC0.134Succinyl-CoA synthetase subunit beta
WU75_19800sucD0.16Succinyl-CoA synthetase subunit alpha
WU75_19770sdhD0.199Succinate dehydrogenase
WU75_19780sdhB0.157Succinate dehydrogenase
WU75_19765sdhC0.182Succinate dehydrogenase
WU75_13785fumA0.497Fumarate hydratase
WU75_09605icd0.179Isocitrate dehydrogenase
WU75_19775sdhA0.144Succinate dehydrogenase
WU75_06430mdh0.177Malate dehydrogenase
WU75_16530lpd0.35Dihydrolipoamide dehydrogenase
Glyoxylate and dicarboxylate metabolismWU75_19760gltA0.129Type II citrate synthase
WU75_19150aceA0.37Isocitrate lyase
WU75_19145aceB0.352Malate synthase
WU75_00290aceB0.315Malate synthase
WU75_10840phbB0.2773-ketoacyl-ACP reductase
WU75_03265katE2.389Catalase
Fatty acid degradationWU75_22235fadB0.151Multifunctional fatty acid oxidation complex subunit alpha
WU75_08655fadE0.184Acyl-CoA dehydrogenase
WU75_20175fadJ0.204Multifunctional fatty acid oxidation complex subunit alpha
WU75_22230fadA0.2083-ketoacyl-CoA thiolase
WU75_20180fadA0.3053-ketoacyl-CoA thiolase
WU75_10835atoB0.433Acetyl-CoA acetyltransferase
WU75_10445atoB0.445Acetyl-CoA acetyltransferase
WU75_12560fadE0.452Acyl-CoA dehydrogenase
WU75_19885fadD0.493Long-chain fatty acid—CoA ligase
Glycine, serine and threonine metabolismWU75_14910gcvP0.113Glycine dehydrogenase
WU75_14915gcvH0.127Glycine cleavage system protein H
WU75_10395betA0.162Choline dehydrogenase
WU75_14930gcvT0.184Glycine cleavage system protein T
WU75_16130lysC0.187Aspartate kinase
WU75_14920glyA0.203Serine hydroxymethyltransferase
WU75_16140ectB0.222Diaminobutyrate-2-oxoglutarate aminotransferase
WU75_16145ectA0.246L-2,4-diaminobutyric acid acetyltransferase
WU75_10400betB0.259Betaine-aldehyde dehydrogenase
WU75_00565sdaA0.264Serine dehydratase
WU75_16135ectC0.27Ectoine synthase
WU75_02030trpB0.397Tryptophan synthase subunit beta
WU75_05755thrC0.429Threonine synthase
WU75_05760thrB0.47Serine kinase
WU75_05330glxK0.495Glycerate kinase
Oxidative phosphorylationWU75_06010petC0.195Cytochrome C
WU75_06015petB0.209Cytochrome B
WU75_14570ccoO0.228Peptidase S41
WU75_14575ccoN0.272Cbb3-type cytochrome c oxidase subunit I
WU75_14560ccoP0.301Cytochrome Cbb3
WU75_06485ppa0.339Inorganic pyrophosphatase
WU75_06020petA0.442Ubiquinol-cytochrome C reductase
WU75_14565ccoQ0.475Cytochrome C oxidase
WU75_02240cyoC0.478Cytochrome o ubiquinol oxidase subunit III
WU75_19125ppk22.159Polyphosphate kinase
WU75_09420cydA3.637Cytochrome d terminal oxidase subunit 1
WU75_09415cydB4.11Cytochrome d ubiquinol oxidase subunit 2
WU75_09410cydX5.362Membrane protein
Pyruvate metabolismWU75_01940yiaY0.171Alcohol dehydrogenase
WU75_03655lldD0.276Lactate dehydrogenase
WU75_22155dld0.322Lactate dehydrogenase
WU75_16665oadA0.324Oxaloacetate decarboxylase
WU75_16060aldB0.397Aldehyde dehydrogenase
WU75_20855gloA2.451Lactoylglutathione lyase
WU75_12805pta8.464Phosphate acetyltransferase
WU75_02150ackA8.851Acetate kinase
WU75_12810ackA10.365Acetate kinase
WU75_09685pflD12.853Pyruvate formate-lyase
WU75_00810gloA13.536Glyoxalase
Propanoate metabolismWU75_15760prpF0.4023-methylitaconate isomerase
WU75_15770prpC0.435Methylcitrate synthase
beta-Lactam resistanceWU75_09315acrA6.699Hemolysin D
WU75_09310acrB8.911Multidrug transporter
WU75_09925acrA40.366Hemolysin D
ABC transportersWU75_10385proW0.106ABC transporter permease
WU75_16175proX0.116Glycine/betaine ABC transporter substrate-binding protein
WU75_10390proX0.122Glycine/betaine ABC transporter substrate-binding protein
WU75_12775oppC0.133Peptide ABC transporter permease
WU75_10380proV0.138ABC transporter ATP-binding protein
WU75_09655aotM0.143Amino acid ABC transporter permease
WU75_09665aotJ0.144Nickel transporter
WU75_13090yejA0.151Diguanylate cyclase
WU75_12770oppB0.164Oligopeptide transporter permease
WU75_12780oppD0.172Oligopeptide transporter ATP-binding component
WU75_09660aotQ0.176ABC transporter
WU75_16170proW0.199Glycine/betaine ABC transporter permease
WU75_08085oppA0.201Peptide ABC transporter substrate-binding protein
WU75_07210yejA0.204Diguanylate cyclase
WU75_12765oppA0.214Peptide ABC transporter substrate-binding protein
WU75_07220yejB0.22Hypothetical protein
WU75_07215yejE0.221Peptide ABC transporter permease
WU75_09670aotP0.228Amino acid transporter
WU75_12785oppF0.228Peptide ABC transporter ATP-binding protein
WU75_04720oppA0.341Peptide ABC transporter substrate-binding protein
WU75_16165proV0.343Glycine/betaine ABC transporter ATP-binding protein
WU75_14765aapQ0.377Amino acid ABC transporter permease
WU75_03180malE0.4Sugar ABC transporter substrate-binding protein
WU75_14775aapP0.405ABC transporter ATP-binding protein
WU75_04605vcaM0.406Multidrug ABC transporter ATP-binding protein
WU75_14055mdlB0.411Multidrug ABC transporter ATP-binding protein
WU75_10275rbsD0.438D-ribose pyranase
WU75_05845btuF0.487Vitamin B12-binding protein
WU75_14760aapJ0.491Amino acid ABC transporter substrate-binding protein
WU75_03185malK2.175Maltose/maltodextrin transporter ATP-binding protein
WU75_19815znuA2.204Zinc ABC transporter substrate-binding protein
WU75_19810znuC2.491Zinc ABC transporter ATPase
WU75_02265artP2.617Arginine ABC transporter ATP-binding protein
WU75_19805znuB2.666Membrane protein
WU75_00425macB14.353Macrolide transporter
Two-component systemWU75_07480glnG0.186Nitrogen regulation protein NR(I)
WU75_13735mcp0.218Chemotaxis protein
WU75_15795tctB0.237TctB
WU75_21750dctD0.288C4-dicarboxylate ABC transporter
WU75_13155ttrB0.314Fe-4S ferredoxin
WU75_21770dctP0.31C4-dicarboxylate ABC transporter
WU75_01920mcp0.32Chemotaxis protein
WU75_21745dctB0.352ATPase
WU75_10200phoA0.353Alkaline phosphatase
WU75_21765dctQ0.368C4-dicarboxylate ABC transporter permease
WU75_00210dctD0.406C4-dicarboxylate ABC transporter
WU75_16210qseC0.423Histidine kinase
WU75_23015fliC0.435Flagellin
WU75_07100mcp0.453Chemotaxis protein
WU75_13380crp0.457Transcriptional regulator
WU75_09825mcp0.471Chemotaxis protein
WU75_16525hapR0.477LuxR family transcriptional regulator
WU75_15800tctA0.485Tripartite tricarboxylate transporter TctA
WU75_14800mcp0.491Chemotaxis protein
WU75_06085tolC2.068Outer membrane channel protein
WU75_15630dcuB2.125C4-dicarboxylate transporter
WU75_06045degP2.148Serine endoprotease DegQ
WU75_04355mcp2.163Chemotaxis protein
WU75_10915luxQ3.377ATPase
WU75_22175mcp4.001Chemotaxis protein
WU75_02450pfeR4.828Transcriptional regulator
WU75_18570cpxA10.981Two-component sensor protein
WU75_18575cpxR26.5Transcriptional regulator
Alanine, aspartate and glutamate metabolismWU75_06265glmS0.037Glucosamine-fructose-6-phosphate Aminotransferase
WU75_07465glnA0.123Glutamine synthetase
WU75_04655putA0.145Pyrroline-5-carboxylate dehydrogenase
WU75_14680-0.286NAD-glutamate dehydrogenase
WU75_05875carB0.343Carbamoyl phosphate synthase large subunit
WU75_05820gltB0.414Glutamate synthase
WU75_05825gltD0.44Glutamate synthase
WU75_05880carA0.46Carbamoyl phosphate synthase small subunit
WU75_18095pyrI0.462Aspartate carbamoyltransferase regulatory subunit
WU75_18090pyrB0.466Aspartate carbamoyltransferase catalytic subunit
WU75_20915ansA2.141Cytoplasmic asparaginase I
WU75_01110ansB2.718L-asparaginase II
WU75_18550aspA7.015Aspartate ammonia-lyase
PTSWU75_03285ptsN0.462PTS fructose transporter subunit IIA
WU75_12990ptsG0.5PTS glucose transporter subunit IIBC
WU75_17910celC2.36Molecular chaperone TorD
WU75_14970fruB2.451Bifunctional PTS system fructose-Specific transporter subunit IIA/HPr protein
WU75_19555ptsH3.973PTS sugar transporter
WU75_00455ulaB3.977PTS ascorbate transporter subunit IIB
WU75_19550ptsI4.075Phosphoenolpyruvate-protein Phosphotransferase
WU75_00460cmtB4.118PTS system mannitol-specific Transporter subunit IIA
WU75_01640cmtB4.539PTS mannitol transporter subunit IIA
WU75_14960fruA5.096PTS fructose transporter subunit IIBC
WU75_00450ulaA6.946PTS beta-glucoside transporter subunit IIBC
Butanoate metabolismWU75_01985acsA0.334Acetoacetyl-CoA synthetase
WU75_10825phaC0.336Poly(3-hydroxyalkanoate) synthetase
Lysine degradationWU75_21960ldcC7.207Lysine decarboxylase LdcC
QSWU75_07805-0.109Cytochrome C
WU75_07800-0.181ABC transporter permease
WU75_07795-0.202ABC transporter permease
WU75_07810ddpD0.216ABC transporter ATP-binding protein
WU75_11620-0.218Peptide ABC transporter permease
WU75_11630-0.233Peptide ABC transporter substrate-binding protein
WU75_11625-0.261Peptide ABC transporter permease
WU75_11610ddpF0.358Chemotaxis protein
WU75_11615ddpD0.484Sugar ABC transporter ATP-binding protein
WU75_21410aphA2.288Transcriptional regulator
Nitrogen metabolismWU75_00760ncd20.2762-nitropropane dioxygenase
WU75_10810napA2.286Nitrate reductase
WU75_15655nirD3.934Nitrite reductase
WU75_10815napB6.27Nitrate reductase
WU75_08850hcp63.107Hydroxylamine reductase
Table 5. The major altered metabolic pathways in S. aureus ATCC25923.
Table 5. The major altered metabolic pathways in S. aureus ATCC25923.
Metabolic PathwayGene IDGene NameFold ChangeGene Description
Two-component systemKQ76_00500-0.373Capsular biosynthesis protein
KQ76_00560wecC0.490UDP-N-acetyl-D-mannosamine dehydrogenase
KQ76_12475nreC2.117Nitrate respiration regulation response regulator NreC
KQ76_12480nreB2.276Nitrate respiration regulation sensor histidine kinase NreB
KQ76_12485nreA2.433Nitrate respiration regulation accessory nitrate sensor NreA
KQ76_10520agrB2.565Histidine kinase
KQ76_03245graS2.989Histidine kinase
KQ76_10785kdpF5.371ATPase
KQ76_04230dltC28.924Alanine-phosphoribitol ligase
Nitrogen metabolismKQ76_12490narI3.529Nitrate reductase
KQ76_12515nirD4.199Nitrite reductase
KQ76_12520nirB5.060Nitrite reductase
KQ76_12460narT6.376Nitrate transporter NarT
KQ76_12500narH5.799Nitrate reductase
KQ76_12505narZ8.442Nitrate reductase
KQ76_12495narJ10.404Nitrate reductase
Riboflavin metabolismKQ76_09200ribE0.373Riboflavin synthase subunit alpha
KQ76_09195ribBA0.413GTP cyclohydrolase
KQ76_09205ribD0.430Diaminohydroxyphosphoribosylaminopyrimidine deaminase
KQ76_09190ribH0.4806,7-dimethyl-8-ribityllumazine synthase
Arginine and proline metabolismKQ76_09185fadM0.109Proline dehydrogenase
KQ76_00580-0.218Aldehyde dehydrogenase
KQ76_13360-0.3031-pyrroline-5-carboxylate dehydrogenase
KQ76_11235rocF0.461Arginase
Atrazine degradationKQ76_11915ureC0.406Urease subunit alpha
KQ76_11910ureB0.412Urease subunit beta
Alanine, aspartate and glutamate metabolismKQ76_13360-0.3031-pyrroline-5-carboxylate dehydrogenase
KQ76_05770carB2.158Carbamoyl phosphate synthase large subunit
KQ76_05765carA3.084Carbamoyl phosphate synthase small subunit
Pyrimidine metabolismKQ76_05745pyrR2.968Phosphoribosyl transferase
KQ76_05760pyrC3.115Dihydroorotase
KQ76_05755pyrB3.213Aspartate carbamoyltransferase
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

Tang, Y.; Yu, P.; Chen, L. Identification of Antibacterial Components and Modes in the Methanol-Phase Extract from a Herbal Plant Potentilla kleiniana Wight et Arn. Foods 2023, 12, 1640. https://doi.org/10.3390/foods12081640

AMA Style

Tang Y, Yu P, Chen L. Identification of Antibacterial Components and Modes in the Methanol-Phase Extract from a Herbal Plant Potentilla kleiniana Wight et Arn. Foods. 2023; 12(8):1640. https://doi.org/10.3390/foods12081640

Chicago/Turabian Style

Tang, Yingping, Pan Yu, and Lanming Chen. 2023. "Identification of Antibacterial Components and Modes in the Methanol-Phase Extract from a Herbal Plant Potentilla kleiniana Wight et Arn" Foods 12, no. 8: 1640. https://doi.org/10.3390/foods12081640

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