Next Article in Journal
Aqueous Two-Phase Flotation Combined with Gold Nanoparticle Colorimetry for Determination of Thiocyanate in Raw Milk
Next Article in Special Issue
Enhanced Antioxidant Extraction from Lonicerae japonicae Flos Based on a Novel Optimization Strategy with Tailored Deep Eutectic Solvents
Previous Article in Journal
Simultaneous Determination of Multi-Class Mushroom Toxins in Mushroom and Biological Liquid Samples Using LC-MS/MS
Previous Article in Special Issue
Natural Deep Eutectic Solvent-Based Ultrasound-Assisted Extraction of Flavonoids from Fagopyrum tataricum Bran
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Paeoniae Radix Alba and Network Pharmacology Approach for Osteoarthritis: A Review

1
College of Pharmacy, Ningxia Medical University, No. 1160 Sheng-Li Street, Yinchuan 750004, China
2
Key Laboratory of Ningxia Ethnomedicine Modernization, Ministry of Education, Ningxia Medical University, Yinchuan 750004, China
3
School of Chinese Materia Medica, Beijing University of Chinese Medicine, Beijing 102488, China
*
Authors to whom correspondence should be addressed.
Separations 2024, 11(6), 184; https://doi.org/10.3390/separations11060184
Submission received: 9 May 2024 / Revised: 3 June 2024 / Accepted: 7 June 2024 / Published: 12 June 2024

Abstract

:
Osteoarthritis (OA) is the most common type of arthritis and affects more than 240 million people worldwide; the most frequently affected areas are the hips, knees, feet, and hands. OA pathophysiology is multifactorial, involving genetic, developmental, metabolic, traumatic, and inflammation factors. Therefore, treatments able to address several path mechanisms can help control OA. Network pharmacology is developing as a next-generation research strategy to shift the paradigm of drug discovery from “one drug, one target” to “multi-component, multi-target”. In this paper, network pharmacology is employed to investigate the potential role of Paeoniae Radix Alba (PRA) in the treatment of OA. PRA is a natural product known for its protective effects against OA, which has recently drawn attention because of its ability to provide physiological benefits with fewer toxic effects. This review highlights the anti-inflammatory properties of PRA in treating OA. PRA can be used alone or in combination with conventional therapies to enhance their effectiveness and reduce side effects. The study also demonstrates the use of network pharmacology as a cost-effective and time-saving method for predicting therapeutic targets of PRA in OA treatment.

1. Introduction

Osteoarthritis (OA) is the top common joint disease and medical concern in the world, with loss of articular cartilage and changes in subchondral bone, menisci, synovium, ligaments, nerves, and periarticular muscle. OA can lead to joint pain, cartilage damage and degeneration, synovial inflammation, and swelling, which seriously affects the patient’s quality of life and is a health burden for families and society [1,2]. Currently, although Western medicine, such as oral therapy [3], intra-articular injection [4], rehabilitation exercise [5], and surgery [6], treatments for OA are available, they all focus on alleviating the clinical symptoms rather than treating the underlying causes. Therefore, it is critical to discover non-invasive treatments that are efficient in alleviating chronic pain, restoring joint function, and reducing side effects in OA patients. Given the complicated pathological mechanism of OA, including neovascularization, subjection to hypoxia, mechanical stress, and inflammation, novel drugs may need to target multiple pathways simultaneously [7,8].
For this reason, natural products with various chemical ingredients and numerous target characteristics have attracted wide attention as alternative therapies for OA. It is estimated that by 2050, there will be USD 5 trillion spent worldwide on medicinal plants, up from the current USD 14 billion yearly [9]. Traditional Chinese Medicine (TCM) has a history of 2300 years in treating various diseases with natural products with few adverse effects and has attracted more advertence in recent years. The natural product artemisinin being found to help treat malaria is a prime example. Radix Paeoniae Alba (aka PRA) is the dried root of Paeonia lactiflora Pall (PLP), a plant belonging to the Ranunculaceae family. The plant is primarily distributed in the temperate regions of the world, especially in South China, Korea, and Japan. PRA is a TCM with the functions of nourishing blood, preventing perspiration, regulating menstruation, extinguishing liver wind, and has astringent and analgesic characteristics. It is widely used in China and other Asian countries and was first listed as a medicine in Sheng Nong’s Herbal Classic [10,11]. The primary representative ingredients in PRA contain flavonoids [12], monoterpenes [13], and triterpenes [14]. Modern research has substantiated the analgesic, immunoregulatory, and anti-inflammatory properties of PRA [15,16]. Previous research shows that PRA alone or in combination with other herbs impacts OA [17]. Nevertheless, there is still a lack of adequate research focused on the treatment of OA by PRA [18,19], and its mechanism of action remains to be answered.
Utilizing improvements in bioinformatics and related databases, network pharmacology integrates technologies and contents from multiple disciplines, such as pharmacology, bioinformatics, and computer science, and discovers the interrelationship between drug–target–pathway–disease, which presents novel ideas and methods for the in-depth study of TCM.
In the present study, given the network pharmacology approach, we aimed to explore the primary bioactive constituents of PRA and predict their effective molecular targets and potential mechanisms in the treatment of OA by using network pharmacology, gene ontology (GO), and the Kyoto Encyclopedia of Genes and Genomes (KEGG) functional enrichment. Our investigation can help to understand the potential role of PRA in the treatment of OA and provide a reference for subsequent drug development and clinical application.

2. Materials and Methods

2.1. Screening and Target Prediction of Active Components of PRA

All of the candidate ingredients of PRA were retrieved using the Traditional Chinese Medicine Systems Pharmacology Database and Analysis Platform (https://old.tcmsp-e.com/tcmsp.php, accessed on 31 May 2022) [20], which is a platform based on the medicinal herbs repository comprehensive information network of chemicals, targets, and affiliated drug targets, including the pharmacokinetic properties of natural compounds like oral bioavailability (OB), drug-likeness (DL), intestinal epithelial permeability, blood–brain barrier penetrability, and water solubility. We applied candidate components that match both OB ≥ 30% and DL ≥ 0.18 as an evaluation criterion, using pharmacodynamics for further exploration.

2.2. Collection of PRA-Related Targets

The molecular structures of the 13 active components were evaluated and acquired in standard SMILES format from the PubChem database (https://pubchem.ncbi.nlm.nih.gov/, accessed on 31 May 2022) [21]. Chem Draw Ultra 8.0 was used to create the 2D chemical structures of potential candidate components. To gather the active components and possible targets of PRA, we searched the Swiss Target Prediction web tool (http://www.swisstargetprediction.ch/, accessed on 31 May 2022) to supplement predicted protein targets with the potential candidate compounds [22]. The specific information for the Swiss Target Prediction was used to extract the conventional SMILES, with the species “Homo sapiens” as the target organism. The target information was then gathered and collated using Microsoft Excel software. Simultaneously, the UniProt database (https://www.uniprot.org/, accessed on 31 May 2022) [23] was utilized to standardize the pharmacological target within each primary substance, which was limited to the “human species”.

2.3. Screening of Disease-Related Targets

The keyword “osteoarthritis” was searched from OMIM [24]. Gene Cards (https://www.genecards.org/, accessed on 31 May 2022) [25], Drug Bank (https://go.drugbank.com/, accessed on 31 May 2022) [26], and PharmGKB database (https://www.pharmgkb.org/, accessed on 31 May 2022) [27] are comprehensive databases to obtain OA-related target gene information. In addition, a Microsoft Excel program (Microsoft Office 2019) was applied to eliminate redundant genes to avoid duplicates.

2.4. Protein–Protein Interaction (PPI) Network Construction and Screening of Core Targets

The Veeny 2.1 overlap (https://bioinfogp.cnb.csic.es/tools/venny/, accessed on 31 May 2022) was created to acquire and evaluate the targets of potential active components in PRA and the disease-related gene of OA. The PPI network regulates all organisms’ appropriate functions. The STRING database (https://cn.stringdb.org/cgi/input?sessionId=b4jJyCOBi5KO&input_page_show_search=on, accessed on 31 May 2022) [28] was used to erect the PPI network to elucidate the interaction of therapeutic target genes and the central gene, with “Homo sapiens” being determined and presented in Cytoscape 3.9.1 software [29]. Only PPI data with an association highest score of more than 0.4 (medium confidence) were set for protein–protein interaction for topology analysis. The degree of a node in the PPI network refers to the number of other nodes with apparent connections. The degrees and total scores of pertinent protein nodes were exported and saved as TSV format files after unnecessary protein nodes were removed in a bid to obtain protein–protein interactions. The data were imported into Cytoscape 3.9.1 software for further analysis of PPI network topology. The central targets are the top 10 targets based on the values computed using network analysis (Cytoscape plugin). Therefore, a node’s importance in a PPI network increases with degree.

2.5. GO and KEGG Pathway Enrichment Analysis

GO terms are classified into 3 aspects, namely, biological process (BP), cellular component (CC), and molecular function (MF). GO and KEGG pathway enrichment analyses were performed in this investigation using the Metascape platform (https://metascape.org/gp/index.html#/main/step1, accessed on 31 May 2022) [30], which includes an extensive annotation function and gene annotation data that are updated regularly. For analysis of the results, the bioinformatic web platform (http://www.bioinformatics.com.cn/?p=1, accessed on 31 May 2022) showed the GO keywords and KEGG pathways sorted by the p-value and count for visual analysis [20].

2.6. Molecular Docking Imitation of Active Compound

Molecular docking technology is used in network pharmacology to verify compound targets, predict the molecules’ relationship of binding mode and affinity, and also research the interaction between molecules. It is a standard method to discover novel drugs [21]. The 2D structure of the small molecule ligands for the 8 representative compounds were first obtained from the PubChem online database (https://pubchem.ncbi.nlm.nih.gov/, accessed on 31 May 2022) as small molecule ligand files for molecular docking. The 3D structure file of the core target protein receptor was acquired from the PDB online database (https://www.rcsb.org, accessed on 31 May 2022). The ligand and protein obtained for molecular docking were prepared using AutoDock Vina software (http://vina.scripps.edu/). The core target protein’s crystal structure was processed, including removal of hydrogenation, modification of amino acids, optimization of energy, and adjustment of force field parameters, and then met the low-energy conformation of the ligand structure. Finally, the target structure was docked with the active component structure using the internal vina of PyRx-Virtual Screening software. The Affinity (kcal/mol) value represents the binding capacity of the representative ingredient and target; the lower the binding capacity, the more stable the binding between the ligand and the receptor.

3. Network Pharmacology: A Multi-Target and Multi-Component Strategy

In 2007, Andrew L. Hopkins made the initial proposal for the field of network pharmacology [22]. Polypharmacology and system biology are combined in network pharmacology to address the limited effectiveness of selective single-target medications [23]. Systems biology involves studying biological structures and processes by developing mathematical formulas and models that explain how a system operates in response to specific gene or protein changes. Network pharmacology, viewed as an emerging paradigm in pharmacological research, complements systems biology by offering insights into pharmacological and disease targets, along with their molecular pathways, from extensive datasets [24]. Nowadays, network pharmacology is increasingly used to study the overall pharmacological effects of drugs on diseases and their therapeutic mechanisms [25]. Various databases are essential for network research. Examples include Swiss Target and Pub Chem for active components, KEGG and DAVID for pathway enrichment, and STRING for drug molecule information. BIND and DIP are crucial for biomolecular interactions, while HPRD and BioGRID provide protein-related data. Moreover, OMIM and DisGeNET are valuable for disease target retrieval. To analyze these data, researchers often utilize Cytoscape for network creation and analysis [26]. Network pharmacology, with its multi-component, multi-target approach, has the potential to address the limitations of single-target therapies and facilitate the discovery of new medications for various diseases, including OA.

4. Extraction and Purification Process of Compounds

When terpenes are extracted from PRA, methanol and ethanol are the most often employed solvents. At present, the total glucosides of paeony (TGP) is thought to be PRA’s primary active ingredient. TGP is hydrophilic monoterpene glycosides typically extracted using water or alcohol. Extraction from PRA can be carried out through methods like ethanol/water reflux, ultrasonic, and microwave-assisted extraction. The initial TGP extract contains impurities such as tannins and carbohydrates, necessitating purification for high-purity TGP. Common purification methods include clarification, organic solvent extraction, and microporous resin purification [27,28]. Researchers have optimized the TGP extraction processes using techniques like response surface methods. For instance, optimal conditions are a solid ratio of 9 mL/g, two extractions with 85% ethanol, each lasting 100 min [29]. In addition, the most used technique for separating analytes from PRA is ultrasonication. Shen et al. extracted PRA for 40 min using an ultrasonic extraction method with 70% (v/v) ethanol, yielding a single standard material for eleven components [30]. In one investigation, 50% acetonitrile demonstrated the highest reflux-based extraction effectiveness of albiflorin and paeoniflorin when compared to solutions including water, 50% ethanol, 50% methanol, and acetonitrile at different concentrations. A 2-h reflux period was also required [31]. Wang et al. fractionated the preparative purification of paeoniflorin and albiflorin from the rhizome of Paeonia lactiflora Pall (PLP) using macroporous resins [32]. Paeoniflorin, albiflorin, benzoyl paeoniflorin, and benzoylalbiflorin were separated from the 50% ethanol extract of PRA using a D101 macroporous resin column [33]. At present, research has focused on developing greener and more efficient methods for extracting and purifying TGP from PRA. For instance, a method that has shown promise involves subcritical water extraction followed by purification using macroporous resin. In this optimal extraction process, the time duration is 22 min, the extraction temperature is 158 °C, the solid ratio is 21 mL/g, and the maximum extraction rate of TGP using subcritical water extraction can reach 7.08% [34].

5. Analysis of Chemical Compounds

Due to advancements in chemical component separation and analysis technologies, an increasing number of new components in PRA can now be separated and analyzed. With the aid of highly sensitive and selective MS detectors, LC-MS has emerged as a superior technique for analyzing trace compounds in compounds with low UV absorption or PRA, when compared to UV detectors. According to earlier studies [35], 26 major ingredients in Paeoniae Radix Rubra (PRR) were identified using LC coupled with DAD and time-of-flight (TOF) MS. Fingerprint analysis, a widely used method in the discriminate analysis of PRA, offers a comprehensive view of the chemical compositions of herbal medicines. Infrared spectroscopy (IR) provides physical properties and chemical composition information, facilitating a more thorough evaluation of TCM quality. Wang et al. reported the difference between PRA and PRR by IR [36]. Furthermore, various spectral techniques, including Raman spectroscopy and X-ray diffraction spectroscopy, have been utilized in the discrimination analysis of PRA. HPLC fingerprinting has emerged as a robust method for investigating chemical variances among species and geographical regions. In HPLC analyses of monoterpene glycosides and polyphenols, reverse-phase chromatography is commonly employed, with ODS (C18) often being the preferred stationary phase [27].

6. Anti-OA Effects

Inflammation is a sophisticated process arising from the host’s defense system response to different internal and external stimuli, which typically occur in various tissues and organs [37]. Chronic inflammation is regarded as a significant factor in various health conditions, such as rheumatoid arthritis [38] and asthma [39]. Therefore, suppressing inflammation is essential for treating diseases associated with chronic inflammation. The identification of new anti-inflammatory drugs from natural sources has increasingly garnered interest in the field of medical research. In recent decades, numerous publications have detailed the diverse biological activities of the Paeonia in both in vitro and in vivo experimental models. These activities include anti-inflammatory [40], antioxidant [41], antibacterial [42], antivirus [43], anti-tumor [44], central nervous system protective [45], and cardiovascular protective effects [46]. Currently, the anti-inflammatory properties of the genus Paeonia have been extensively researched.

6.1. Crude Extracts

In a rat model of arthritis induced by Complete Freund’s Adjuvant, PRA extract demonstrated a reduction in inflammation associated with arthritis. Administered at doses of 10 mg of PRA per gram of body weight, the swelling inhibition rates were 58.2%, 50.48%, and 35.67% at 3 h, 6 h, and 15 h, respectively. At the medium dosage, swelling inhibition percentages were 50.82%, 50.05%, and 35.39%, while at intermediate and low dosages, the inhibition percentages were 33.78%, 39.56%, and 28.66%, respectively. The decrease in phosphodiesterase (PDE) activity in neutrophils and possibly other inflammatory cells may play a role in mediating this anti-inflammatory impact [47].

6.2. Pure Compounds

Liao et al. conducted a series of studies of the anti-OA activity and mechanism of action of TGP [48]. In vivo and in vitro experiments demonstrated that TGP had a protective effect on cartilage. Treatment with TGP could induce the synthesis of critical elements in the cartilage extracellular matrix and downregulate the synthesis of degrading enzymes in the extracellular matrix. Regarding the underlying mechanisms, TGP inhibited the phosphorylation and nuclear translocation of p65 by regulating the nuclear factor-kappa B (NF-κB) signaling pathway. In addition, TGP could reduce the secretion of IL-1β, IL-6, and tumor necrosis factor-α (TNF-α). Moreover, it has a sustained effect on coupled subchondral bone remodeling through regulation of the OPG/RANKL/RANK pathway. Further investigation showed that TGP may protect articular cartilage by downregulating the NF-κB signaling pathway and may support coupled subchondral bone remodeling by regulating the OPG/RANKL/RANK signaling pathway in the DMM-induced KOA model of mouse for OA treatment. Another study suggests that paeoniflorin could inhibit IL-1β-induced MMP secretion via the NF-κB pathway in chondrocytes, exerting anti-OA effects [49]. Hu et al. found that paeoniflorin may exert its protective effect by inhibiting IL-1β-induced chondrocyte apoptosis by regulating the Bax/Bcl-2/caspase-3 signaling pathway [50]. In one study, paeoniflorin shows chondroprotective effects under IL-1β stress by regulating the circ-PREX1/miR-140-3p/WNT5B axis [51]. Furthermore, paeoniflorin inhibited the production of nitric oxide and prostaglandin E2 induced by IL-1β and the expression of inducible nitric oxide synthase and cyclooxygenase-2 in chondrocytes. The NF-κB transcription factors play a prominent role in various pathways and mediators influencing the occurrence of OA. IL-1β may activate NF-κB by triggering the degradation of IκBα, which leads to the translocation of NF-κB p65 from the cytoplasm to the nucleus, thereby regulating the expression of various inflammation-related genes. Treatment with paeoniflorin reduced the expression of NF-κB p65 and IκB-α degradation induced by IL-1β in OA chondrocytes. This therapeutic effect may occur through modulation of NF-κB signaling pathway [52].
Based on our research of the literature on published reviews [48,53], extracts and compounds of the PRA exhibit a broad spectrum of anti-OA activity and have multiple targets of action. Among the components of the extracts, monoterpene glycosides are the main active compounds. The plant extract TGP has been used clinically in the treatment of inflammatory diseases [54]. TGP has good efficacy and few side effects in rheumatoid arthritis patients but produces slow therapeutic effects [55]. Currently, TGP is often used in combination with other drugs to treat autoimmune diseases, but there are few pharmacological reports on this topic.

7. Coupling of Network Pharmacology with PRA in OA Diseases

7.1. Screening of Candidate Ingredients of PRA

The TCMSP was screened, and 85 components of PRA were identified. According to the filtering criteria of OB ≥ 30% and DL ≥ 0.18, 13 active compounds were identified that satisfied the criteria (Table 1). The 13 active constituents of PRA contained 11alpha,12alpha-epoxy-3beta-23-dihydroxy-30-norolean-20-en-28,12beta-olide, paeoniflorgenone, (3S,5R,8R,9R,10S,14S)-3,17-dihydroxy-4,4,8,10,14-pentamethyl-2,3,5,6,7,-9-hexahydro-1H-cyclopenta[a]phenanthrene-15,16-dione (DPHCD), lactiflorin, paeoniflorin, paeoniflorin_qt, albiflorin_qt, benzoyl paeoniflorin, Mairin, beta-sitosterol, sitosterol, kaempferol, and (+)-catechin, extracted through TCMSP.

7.2. Potential Target Genes and the PPI Network Map of PRA Therapy for OA

A total of 3493 OA target genes, excluding duplicates, were determined through investigations of the GeneCard, OMIM, Drug Bank, and PharmGKB databases. For the 13 active components of PRA, 707 target genes were retrieved from the TCMSP database and Swiss Target Prediction (Supplementary Table S1). Upon verification, duplicates were discarded, yielding 406 target genes. Using the Venny 2.1 program, OA target genes and PRA target genes were intersected to obtain 232 candidate targets (Figure 1A) (Table 2). The PPI network map was created by sequentially importing the 232 target genes into the STRING database (Figure 1B). The network was visualized and explored by evaluating centrality and other variables using the Cytoscape program. Following these guidelines, all of the targets were organized into circles.
The Cytoscape software (a Cytoscape plugin) discovered and listed the top 10 potential genes closely related to PRA and OA from PPI network results (Figure 1C). These genes involve RAC-alpha serine/threonine-protein kinase (AKT1), caspase-3 (CASP3), glyceraldehyde-3-phosphate dehydrogenase liver (GAPDH), intercellular adhesion molecule 1 (ICAM1), interleukin-6 (IL-6), MAP kinase ERK1 (MAPK3), matrix metalloproteinase 9 (MMP9), prostaglandin G/H synthase2 (PTGS2), tumor necrosis factor (TNF), and vascular endothelial growth factor A (VEGFA).

7.3. Construction and Analysis of the PRA–OA–Potential Target Gene Network

To construct a PRA–OA–potential representative gene network, we must further input the data of Supplementary Table S1 and Table 2 into Cytoscape software (version 3.9.1) (Figure 2). A total of 85 components of PRA were obtained from the TCMSP database. After screening, 13 active ingredients were finally selected, and the targets of 13 active components were predicted by the Swiss Target Prediction. We found 116 potential drug targets for 8 active components. The best 19 targets, PGR, AR, CYP19A1, ESR1, ESR2, AGTR1, PTGS2, SHBG, NR3C1, CHRM2, GABRA2, PTPN11, PTGS1, PTPN1, NOS2, AKR1B10, PLA2G1B, SRC, and KDR, were discovered to have a greater degree in this process, emphasizing their prominence in the network (degree ≥ 5) (Table 3).

7.4. GO enrichment and KEGG Pathway Analysis

The entrezIDs of the 232 target genes are presented in Table 2. Examinations of these genes’ functional enrichment were performed to dig into any prospective physiological functions in the therapy of OA. According to the results, they were enriched with 2301 GO-BP terms, 152 GO-CC terms, and 289 GO-MF terms from the Metascape database (p < 0.05). A total of 20 significantly enriched gene physiological activity catalogs were chosen to generate a histogram for subsequent investigation (Figure 3A, Supplementary Figure S1). Our BP results (Figure 3B) indicated that the function of PRA active ingredients in OA is mainly concentrated upon hormone response, regulation of defense responses, cellular responses to lipids, responses to the molecule of bacterial origin, responses to inorganic components, response to extracellular stimulus, positive regulation of responses to external stimuli, regulation of MAPK cascades, protein phosphorylation, regulation of systemic processes, response to decreased oxygen levels, positive regulation of cell migration, response to a steroid hormone, response to wounding, positive regulation of cell death, circulatory system process, regulation of hormone levels, response to xenobiotic stimulus, eicosane metabolic process, and rhythmic process. The MF items mainly contained protein serine/threonine/tyrosine kinase activity, nuclear receptor activity, oxidoreductase activity, protein tyrosine kinase activity, protein homodimerization activity, kinase binding, peptide binding, phosphatase binding, protein domain specific binding, carboxylic acid binding, MAP kinase activity, serine hydrolase activity, prostaglandin receptor activity, ligand-gated anion channel activity, cytokine receptor binding, non-membrane spanning protein tyrosine kinase activity, bile acid binding, heat shock protein binding, integrin binding, and virus receptor activity (Figure 3C). The CC items mainly included the membrane raft, receptor complex, side of membrane, dendrite, leading-edge membrane, vesicle lumen, neuron projection membrane, extracellular matrix, perinuclear region of cytoplasm, transcription regulator complex, nuclear envelope lumen, nuclear envelope, lytic vacuole, endocytic vesicle, cytoplasmic side of the membrane, dendritic spine, outer organelle membrane, apical part of cell, endoplasmic reticulum lumen, and glutamatergic synapse (Figure 3D). These results suggest that PRA is associated with the therapy of OA through multiple gene biological functions, which will further clarify that PRA is beneficial against diseases, including OA.
To further acquire a profound understanding of the mechanism of action of PRA in the treatment of OA, 209 signaling pathways were discovered by KEGG pathway enrichment analysis (p < 0.05) (Figure 4A,B). Scatterplots (Figure 4A) and histograms (Supplementary Figure S2) are listed using the top 20 crucial signaling pathways. As shown in Figure 4A, various signaling pathways are deeply connected with OA, including pathways in cancer, the VEGF signaling pathway, the chemokine signaling pathway, the NF-kappa B signaling pathway, and regulating pluripotency of stem cells signaling pathways (Figure 4B). According to the KEGG analysis results, 20 significant therapeutic pathways were contained in the targets shown in Table 4. Furthermore, the KEGG mapper was used to color targets red in the significant NF-kappa B signaling pathway, as shown in Figure 5.

7.5. Analysis of Molecular Docking Results

Molecular docking was executed to verify the binding ability of AKT1, CASP3, GAPDH, ICAM1, IL-6, MAPK3, MMP9, PTGS2, TNF, and VEGFA, with Mairin, be-ta-sitosterol, sitosterol, kaempferol, (+)-catechin, paeoniflorgenone, DPHCD, paeoniflorin, in line with network analysis and preliminary research. When the binding energy is small, it indicates that the binding degree of the ligand to the target protein is high. The molecular binding energy ranges from −6 to −10.3 kcal/mol, indicating that the binding energy results have good binding activity (Figure 6). The results in Figure 6 show that the representative components of PRA have a better binding activity to the core target, which, to a certain extent, verifies the reliability of the preliminary results of network pharmacology. Figure 7A–F shows the 3D binding map and hydrogen bond map of the optimal combination of 3 core components and their optimal target proteins in the active components of PRA. As shown in Figure 7A, kaempferol could form a structure of two hydrogen bonds with MET-522 and LEU-535 in PTGS2, and their bond lengths were 2.5 and 2.4. As shown in Figure 7B, (+)-catechin could form a hydrogen bond with GLN-461 and GLY-45 in PTGS2, and their bond lengths were 2.8 and 2.0. As shown in Figure 7C, DPHPD could form 1 hydrogen bond with GLN-372 in PTGS2 with a bond length of 2.4. As shown in Figure 7D, kaempferol could form 1, 2, and 3 hydrogen bonds with LEU-188, GLN-227, and ALA-189 in MMP9, with bond lengths of 2.7, 2.4, 2.1, 2.2, 2.5, and 3.5, respectively. As shown in Figure 7E, (+)-catechin could form a hydrogen bond with LEU-243 and LEU-188 in MMP9, respectively, and the bond lengths are 2.5, 2.2, and 2.9, respectively. As shown in Figure 7F, DPHCD could form 1 hydrogen bond with ALA-199 in MMP 9 with a bond length of 2.5.
In conclusion, there is an interaction between core genes with different bonds and key active ingredients. Our results show that the three key active ingredients in PRA can have a stronger binding force to PTGS2 and MMP9 to produce a therapeutic effect on OA.

8. Discussion

Osteoarthritis (OA) is a common degenerative joint disease that primarily affects 10% of men and 18% of women over 60 years of age; major signs include arthritis, synovial inflammation, joint edge, and underlying subchondral bone lesions, which poses a serious health threat and decreases the quality of life [56,57]. Currently, the representative drugs for the treatment of OA include selective COX-2 enzyme inhibitors, non-steroidal anti-inflammatory drugs (NSAIDs), glucocorticoids, viscosupplements, and other conventional medicines [58]. However, the choice of medical therapy for OA is still debatable because most of the medications used today are primarily intended to address patients’ pain, inflammatory symptoms, and other uncomfortable symptoms rather than prevent or treat the disease and have high toxicity and poor efficacy.
It is worth mentioning that Traditional Chinese Medicine (TCM) functions have been progressively implemented in the prevention and treatment of OA due to their significant tonifying of the liver and kidney, comforting of tendons, activation of blood circulation, dredging of collateral and alleviating pain, long-term efficacy, few adverse effects, and has been identified by the majority of patients [59]. The mechanism of OA is associated with numerous variables, including pro-inflammatory and anti-inflammatory cytokines and cytokine/receptor signaling pathways [60,61]. Radix Paeoniae Alba has a major advantage in treating OA as a significant blood-activating, pain-relieving medication, but further research into its precise molecular mechanism is needed [11]. In the present study, 13 candidate ingredients of PRA were found to play significant roles in the treatment of OA and to be associated with various proteins and signaling pathways, suggesting that these candidate constituents have a potential research value.
The 13 representative ingredients from PRA collection results were widely investigated and verified, demonstrating that they have positive effects on immunity and inflammation. However, there have been a few investigations on OA treatment by PRA, and most of them tended to concentrate on the isolated components from PRA, which required systematic insights and depth [62]. Unfortunately, massive constituents of Chinese herbal medicine have specific curative effects in experiments but are less than satisfactory in clinical applications [63].
Network pharmacology, which is based on the theoretical advancement of multidirectional pharmacology, systems biology, and omics, is a booming research approach for the design and development of new drugs with the development of systems biology technology. The multi-component, multi-target, and systematic regulation notions are embodied in the holistic perspective of TCM and the syndrome differentiation and therapy concept, which shares massive parallels with the network pharmacology research idea [64]. Based on many experimental data and clinical trial results, network pharmacology could be used to analyze the potential mechanism of clinical treatment and precaution of diseases, particularly appropriate for the characteristics of intricate components and multiple targets and mechanisms of TCM. Therefore, this study aims to unveil the anti-OA mechanism of PRA. In addition, the molecular docking pattern between the compound and the target indicated an excellent combination, further verified by the intrinsic relationship between PRA and OA.
According to the PPI network map of PRA therapy for OA, it was determined that paeoniflorgenone, DPHCD, beta-sitosterol, kaempferol, (+)-catechin, paeoniflorin, sitosterol, and Mairin were the most central potential therapeutic components of PRA anti-OA. Kaempferol has been recognized as an effective agent for alleviating the clinical symptoms of OA. Previous studies found that kaempferol inhibits interleukin-1β and, thus, inhibits cartilage damage by RANKL and MAOKs [65]. Zhuang and his colleague’s research revealed that kaempferol alleviates the interleukin-1β-induced inflammation in rat OA chondrocytes via suppression of NF-κB had significant anti-inflammatory and anti-arthritis effects [66]. In addition, kaempferol activates the PI3K/AKT/mTOR signaling pathway by downregulating miR-146 and inhibiting Decurion expression [67]. Cheng B C et al. pointed out that paeoniflorin can protect cartilage by targeting (IL)-1β in the NF-κB pathway [49]. However, no primary research on beta-sitosterol in the treatment of OA has been reported so far, and further research is worthy of following-up. Studies of other components have shown evidence that (+)-catechin normalizes the gene expression levels of inflammatory cytokines such as COX-2, TNF-α, IL-1β, IL-6, and NF-κB [68]. In summary, the 8 active components of PRA have different therapeutic effects on OA, involving a diverse range of cytokines and signaling pathways and warrant deep exploration.
We applied a PPI network (Figure 2); many of the 8 compounds can regulate many target genes in OA in PRA. These genes contain but are not limited to NR3C, CHRM2, GABRA2, PTPN11, PTGS1, PTPN1, and NOS2. This point completely reflected the characteristics of multi-target and multi-component Traditional Chinese Medicine. Additionally, the PPI results further show that the 232 target proteins are related and interact rather than exist independently [69]. As shown in Table 2, PGR, AR, CYP19A1, ESR1, ESR2, AGTR1, and PTGS2 were the important target genes in the PPI network.
We used GO and KEGG enrichment analysis to identify the 232 possible targets directly connected to the formation and progression of OA to anticipate the biological processes and signaling pathways that may underlie the therapeutic impact of PRA in OA. The 20 significantly enriched KEGG keywords that mainly were connected to pathways in cancer, VEGF signaling pathway, chemokine signaling pathway, NF-κB signaling pathway, FoxO signaling pathway, and regulating pluripotency of stem cells signaling pathways are depicted in Figure 4. Inflammation has a vital and close relationship with the pathogenesis of OA. The NF-κB pathway is a classic pathway highly related to OA. Some studies have revealed that several mouse models of OA in the NF-κB pathway were better-characterized signaling pathways loading in the animal model of OA [70,71]. The study showed that downregulating NF-κB signaling can prevent cartilage degeneration partially [72]. Inhibition of NF-κB Activity and chondrocyte pyroptosis may ameliorate cartilage degeneration and development in OA [73,74]. The HIF-1 signaling pathway contributes to many autoimmune diseases containing OA [75,76]. HIF-1 plays a fundamental role in chondrocyte survival and maintains the homeostatic conditions of articular cartilage, which regulate both autophagy and apoptosis [77]. Previous research has shown that activated or inhibitors of the HIF-1 signaling pathway can cure OA [78,79]. The study of the HIF-1 signaling pathway helps to reveal the pathogenesis of OA and reduces abnormal inflammation, providing potential targets. Angiogenesis occupies a pivotal position in OA. Vascular endothelial growth factor (VEGF) is a pro-inflammatory factor that directly affects vascular endothelial cells and early neovascularization, ultimately leading to joint inflammation in OA. In recent years, cumulative evidence has shown that angiogenesis plays a substantial part in synovitis, osteochondral channel formation, joint marginal osteophyte formation and joint pain in OA and is one of the crucial factors in the formation and aggravation of OA [80,81]. Recent studies have reported that increasing VEGF expression and promoting EPC angiogenesis in OA synovial fibroblasts (OASFs) by inhibiting miR-485-5p synthesis via PI3K and Akt signaling could be used to treat OA [82].
Additionally, inhibition of VEGF overexpression contributes to the recovery from OA [83]. Furthermore, we also found other pathways, such as pathways in cancer, lipid and atherosclerosis, chemical carcinogenesis–receptor activation, neuroactive ligand–receptor interaction, proteoglycans in cancer, cAMP signaling pathway, serotonergic synapse, and steroid hormone biosynthesis, indicating that PRA has potential applications in tumors, atherosclerosis, and other diseases. Based on the aforementioned multiple pathways, it is speculated that PRA cures OA by the inflammatory response, immune regulation, and other processes. In addition, to further explore the potential molecular mechanism of PRA in the treatment of OA, we screened 8 active compounds as ligands and conducted molecular docking studies on 10 targets closely related to OA. The results showed that the 2 potential target genes had a better binding to the 3 active compounds.
Traditional Chinese Medicine (TCM) stands as one of the crown jewels of Chinese culture, boasting a rich history of use across East Asia and Southeast Asia, and has been extensively employed since the dawn of ancient civilizations. With the rapid development of bioinformatics, systems biology, and multi-pharmacology, network-based drug discovery is considered a promising and cost-effective method for drug development. At present, a large number of pertinent databases and tools offer vital support for the field of network pharmacology in TCM. Databases frequently employed in the network pharmacology of TCM include TCMSP [84], drug information databases [85], PubChem [86], target interaction databases [87], and gene–disease association databases [88], which provide an understanding of the effects of herbs on diseases. The network pharmacology algorithm is a frequently employed network clustering technique that commences with a random node (drug, target, or disease), computes the similarity between that node and its neighboring nodes, and constructs a “drug–target–disease” network. In addition, the tools and databases employed in the field of TCM network pharmacology are founded on constraints imposed on ranking functions. These constraints take into account the smoothness of the network and the utilization of prior information to prioritize disease genes and deduce the associations among protein complexes. In addition to data acquisition and analysis, visualization is a crucial component of network pharmacology, rendering the network more intuitive. Cytoscape is an open-source platform designed for visualizing molecular interaction networks and biological pathways, while integrating these networks with annotations, gene expression profiles, and other data on gene states. Currently, most databases include web services that can transfer data between databases, serving as a conversion platform. These services are presented to users as web services integrated with various databases or browser plugins, ensuring their reliability. Currently, TCM has documented over 10,000 herbs used in more than 100,000 formulas [89]. However, research in network pharmacology for TCM is nascent and requires further development. Additionally, as data from TCM and clinical research accumulate, and with the interaction of various analytical and experimental methods, researchers can gain more substantial and authentic insights.

9. Conclusions and Future Perspective

The single target/single molecule, side effects, drug resistance, and decreased effectiveness of current medications continue to limit their ability to prevent OA and its progression. This is true even with the advancements in modern therapeutics. From the perspective of drug development, network-based drug discovery is seen as a viable and cost-effective approach, especially considering the rapid advancements in system biology, bioinformatics, and polypharmacology. PRA also has a wide variety of biological traits and structural polymorphism, which provides opportunities to find different lead entities. In the present study, we investigated the effective active ingredients and molecular mechanisms of PRA in the treatment of OA by the network pharmacology approach. The most significant potential 8 active ingredients of PRA in OA treatment were screened out that involves 232 target genes correlated with OA, among which AKT1, CASP3, GAPDH, ICAM1, IL-6, MAPK3, MMP9, PTGS2, TNF, VEGFA are identified as core target genes. Paeoniflorgenone, DPHCD, beta-sitosterol, kaempferol, (+)-catechin, paeoniflorin, sitosterol, and Mairin were identified as the essential phytochemicals treating OA. The signaling pathways of PRA in the treatment of OA mainly contain the HIF-1 signaling pathway, FoxO signaling pathway, VEGF signaling pathway, Necroptosis, and NF-κB signaling pathway. In addition, we verified the most positive effect on OA of the active components kaempferol, (+)-catechin, and DPHCD by molecular docking.
Despite the fact that PRA has demonstrated anti-inflammatory and pain-relieving properties for osteoarthritis in pre-clinical research, human study data is still lacking; therefore, conclusions cannot be made. Therefore, clinical trials are required to determine the best way to administer them to patients with osteoarthritis (OA) as well as the ideal dosage for pain and inflammation alleviation. Moreover, a lack of research calls for the quick development and application of innovative network approaches to challenge accepted theories about drug discovery and expand our understanding of action principles. Using network pharmacology with PRA can help open up new ways of addressing the limitations of the treatments that are now available to prevent OA.

10. Study Limitations

Although network pharmacology is a fast and effective method for predicting multiple drug targets in complex diseases, there are also some shortcomings; for example:
(1)
The ingredients in TCM are incredibly complex, and the ingredients that play a therapeutic role under the action of metabolic enzymes after absorption by the human body are not all the prototype ingredients of TCM; both OB and DL values of chemical ingredients are used to screen the chemical ingredients of active ingredients, so the tissue distribution of bioavailability of Chinese medicine cannot be quantified objectively.
(2)
Network pharmacology matches relevant information from public databases, but the existing databases lack a unified standard.
(3)
Obtaining disease targets by searching disease-related databases ignores the pathophysiological changes of diseases in the clinical development process and lacks objectivity.
(4)
The foundation of the therapeutic effect in Traditional Chinese Medicine does not necessarily lie in a single chemical compound but rather in the collective action of many chemical compounds. Consequently, research focused on the efficacy of chemical drugs targeting single pathways is not fully applicable to the study of the therapeutic effects of Traditional Chinese Medicine.
(5)
In this study, the number of components was not considered. However, the impact of both quantity and concentration on therapeutic efficacy cannot be disregarded. Only when a sufficient quantity of the drug reaches the target site can it effectively deliver its therapeutic effect.
(6)
In this study, the quantity of active ingredients in PRA was not taken into account, and the effect of the quantity of ingredients on the efficacy of the drug was overlooked.
In the present study, the interaction between the active ingredient of PRA and the biological function of the target is ignored, and the molecular binding between the active ingredient and the target of the non-intersection ingredient is seldom considered. In addition, PRA anti-OA has not been further validated by animals and cell models of corresponding OA.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/separations11060184/s1: Supplementary Table S1. The 706 target genes of PRA active ingredients. Supplementary Figure S1. The top 20 remarkably enriched GO analysis for biological function of potential target genes of PRA anti-OA. (A) GO-BP terms; (B) GO-CC terms; (C) GO-MF terms; (p < 0.05). Supplementary Figure S2. The top 20 remarkably enriched KEGG analysis for signaling pathway of potential target genes of PRA anti-OA; (p < 0.05).

Author Contributions

B.W.: Investigation, Methods, Drawing, Writing the first draft, drawing graphic abstracts, scatter plots, histograms, and organizing data. Y.Z. and C.B.: Helped with topic selection and reviewed and finalized the paper. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by grants from the National Natural Science Foundation of China (No. 81860755).

Data Availability Statement

The data are available for reproduction of results on request from the corresponding author.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. McAllister, M.; Chemaly, M.; Eakin, A.J.; Gibson, D.S.; McGilligan, V.E. NLRP3 as a potentially novel biomarker for the management of osteoarthritis. Osteoarthr. Cartil. 2018, 26, 612–619. [Google Scholar] [CrossRef]
  2. Nelson, A.E.; Schwartz, T.A.; Alvarez, C.; Golightly, Y.M. The prevalence of hand osteoarthritis: Comment on the article by Eaton et al. Arthritis Rheumatol. 2022, 74, 1861–1862. [Google Scholar] [CrossRef]
  3. Sok, D.; Raval, S.; McKinney, J.; Drissi, H.; Mason, A.; Mautner, K.; Kaiser, J.M.; Willett, N.J. NSAIDs Reduce Therapeutic Efficacy of Mesenchymal Stromal Cell Therapy in a Rodent Model of Posttraumatic Osteoarthritis. Am. J. Sports Med. 2022, 50, 1389–1398. [Google Scholar] [CrossRef]
  4. Zhang, T.; Chen, S.; Dou, H.; Liu, Q.; Shu, G.; Lin, J.; Zhang, W.; Peng, G.; Zhong, Z.; Fu, H. Novel glucosamine-loaded thermosensitive hydrogels based on poloxamers for osteoarthritis therapy by intra-articular injection. Mater. Sci. Eng. C Mater. Biol. Appl. 2021, 118, 111352. [Google Scholar] [CrossRef]
  5. Knoop, J.; Dekker, J.; van Dongen, J.M.; van der Leeden, M.; de Rooij, M.; Peter, W.F.; de Joode, W.; van Bodegom-Vos, L.; Lopuhaä, N.; Bennell, K.L.; et al. Stratified exercise therapy does not improve outcomes compared with usual exercise therapy in people with knee osteoarthritis (OCTOPuS study): A cluster randomised trial. J. Physiother. 2022, 68, 182–190. [Google Scholar] [CrossRef]
  6. Padilla-Martinez, J.P.; Lewis, W.; Ortega-Martinez, A.; Franco, W. Intrinsic fluorescence and mechanical testing of articular cartilage in human patients with osteoarthritis. J. Biophotonics 2018, 11, e201600269. [Google Scholar] [CrossRef]
  7. Talaie, R.; Torkian, P.; Clayton, A.; Wallace, S.; Cheung, H.; Chalian, M.; Golzarian, J. Emerging Targets for the Treatment of Osteoarthritis: New Investigational Methods to Identify Neo-Vessels as Possible Targets for Embolization. Diagnostics 2022, 12, 1403. [Google Scholar] [CrossRef]
  8. Ma, L.; Zheng, X.; Lin, R.; Sun, A.R.; Song, J.; Ye, Z.; Liang, D.; Zhang, M.; Tian, J.; Zhou, X.; et al. Knee Osteoarthritis Therapy: Recent Advances in Intra-Articular Drug Delivery Systems. Drug Des. Devel Ther. 2022, 16, 1311–1347. [Google Scholar] [CrossRef]
  9. Rai, V.; Khatoon, S.; Bisht, S.; Mehrotra, S. Effect of cadmium on growth, ultramorphology of leaf and secondary metabolites of Phyllanthus amarus Schum. and Thonn. Chemosphere 2005, 61, 1644–1650. [Google Scholar] [CrossRef] [PubMed]
  10. Zhao, Y.; Zhang, Y.; Kong, H.; Zhang, M.; Cheng, J.; Wu, J.; Qu, H.; Zhao, Y. Carbon Dots from Paeoniae Radix Alba Carbonisata: Hepatoprotective Effect. Int. J. Nanomed. 2020, 15, 9049–9059. [Google Scholar] [CrossRef] [PubMed]
  11. Tan, Y.-Q.; Chen, H.-W.; Li, J.; Wu, Q.-J. Efficacy, chemical constituents, and pharmacological actions of Radix Paeoniae Rubra and Radix Paeoniae Alba. Front. Pharmacol. 2020, 11, 1054. [Google Scholar] [CrossRef] [PubMed]
  12. Liu, L.; Yuan, Y.; Tao, J. Flavonoid-Rich Extract of Paeonia lactiflora Petals Alleviate d-Galactose-Induced Oxidative Stress and Restore Gut Microbiota in ICR Mice. Antioxidants 2021, 10, 1889. [Google Scholar] [CrossRef] [PubMed]
  13. Zheng, X.; Yin, M.; Chu, S.; Yang, M.; Yang, Z.; Zhu, Y.; Huang, L.; Peng, H. Comparative Elucidation of Age, Diameter, and “Pockmarks” in Roots of Paeonia lactiflora Pall. (Shaoyao) by Qualitative and Quantitative Methods. Front. Plant Sci. 2021, 12, 802196. [Google Scholar] [CrossRef] [PubMed]
  14. Zhou, C.; Zhang, Y.; Sheng, Y.; Zhao, D.; Lv, S.; Hu, Y.; Tao, J. Herbaceous peony (Paeonia lactiflora Pall.) as an alternative source of oleanolic and ursolic acids. Int. J. Mol. Sci. 2011, 12, 655–667. [Google Scholar] [CrossRef] [PubMed]
  15. Yan, B.F.; Chen, X.; Chen, Y.F.; Liu, S.J.; Xu, C.X.; Chen, L.; Wang, W.B.; Wen, T.T.; Zheng, X.; Liu, J. Aqueous extract of Paeoniae Radix Alba (Paeonia lactiflora Pall.) ameliorates DSS-induced colitis in mice by tunning the intestinal physical barrier, immune responses, and microbiota. J. Ethnopharmacol. 2022, 294, 115365. [Google Scholar] [CrossRef] [PubMed]
  16. Li, X.; Wang, W.; Su, Y.; Yue, Z.; Bao, J. Inhibitory effect of an aqueous extract of Radix Paeoniae Alba on calcium oxalate nephrolithiasis in a rat model. Ren. Fail. 2017, 39, 120–129. [Google Scholar] [CrossRef] [PubMed]
  17. Chen, P.; Ruan, A.; Zhou, J.; Zhang, X.; Ma, Y.; Peng, H.; Yang, T.; Wang, Q. Effect of total glucosides of paeony on inflammation and degeneration of chondrocytes in osteoarthritis. Chin. J. Tissue Eng. Res. 2021, 25, 4614. [Google Scholar] [CrossRef] [PubMed]
  18. Ye, H.Z.; Zheng, C.S.; Xu, X.J.; Wu, M.X.; Liu, X.X. Potential synergistic and multitarget effect of herbal pair Chuanxiong Rhizome-Paeonia Albifora Pall on osteoarthritis disease: A computational pharmacology approach. Chin. J. Integr. Med. 2011, 17, 698–703. [Google Scholar] [CrossRef] [PubMed]
  19. Lin, L.X.; Zheng, F.Y.; Huang, G.Z.; Li, S.X.; Liu, F.J.; Man, S.; Wu, J.R.; Li, Y.Y. Anti-arthritic-related metal bioavailability for optimal compatibility ratio of Aconiti Radix Cocta and Paeoniae Radix Alba. Zhongguo Zhong Yao Za Zhi 2020, 45, 5770–5776. [Google Scholar] [CrossRef]
  20. Zhang, L.; Han, L.; Wang, X.; Wei, Y.; Zheng, J.; Zhao, L.; Tong, X. Exploring the mechanisms underlying the therapeutic effect of Salvia miltiorrhiza in diabetic nephropathy using network pharmacology and molecular docking. Biosci. Rep. 2021, 41, BSR20203520. [Google Scholar] [CrossRef]
  21. Shaker, B.; Yu, M.S.; Lee, J.; Lee, Y.; Jung, C.; Na, D. User guide for the discovery of potential drugs via protein structure prediction and ligand docking simulation. J. Microbiol. 2020, 58, 235–244. [Google Scholar] [CrossRef] [PubMed]
  22. Hopkins, A.L. Network pharmacology: The next paradigm in drug discovery. Nat. Chem. Biol. 2008, 4, 682–690. [Google Scholar] [CrossRef] [PubMed]
  23. Zhang, R.; Zhu, X.; Bai, H.; Ning, K. Network Pharmacology Databases for Traditional Chinese Medicine: Review and Assessment. Front. Pharmacol. 2019, 10, 123. [Google Scholar] [CrossRef]
  24. Wang, X.; Wang, Z.Y.; Zheng, J.H.; Li, S. TCM network pharmacology: A new trend towards combining computational, experimental and clinical approaches. Chin. J. Nat. Med. 2021, 19, 1–11. [Google Scholar] [CrossRef] [PubMed]
  25. Luo, T.T.; Lu, Y.; Yan, S.K.; Xiao, X.; Rong, X.L.; Guo, J. Network Pharmacology in Research of Chinese Medicine Formula: Methodology, Application and Prospective. Chin. J. Integr. Med. 2020, 26, 72–80. [Google Scholar] [CrossRef] [PubMed]
  26. Zhang, R.Z.; Yu, S.J.; Bai, H.; Ning, K. TCM-Mesh: The database and analytical system for network pharmacology analysis for TCM preparations. Sci. Rep. 2017, 7, 2821. [Google Scholar] [CrossRef] [PubMed]
  27. Yan, B.; Shen, M.; Fang, J.; Wei, D.; Qin, L. Advancement in the chemical analysis of Paeoniae Radix (Shaoyao). J. Pharm. Biomed. Anal. 2018, 160, 276–288. [Google Scholar] [CrossRef]
  28. Deng, H.; Yan, C.; Xiao, T.; Yuan, D.; Xu, J. Total glucosides of Paeonia lactiflora Pall inhibit vascular endothelial growth factor-induced angiogenesis. J. Ethnopharmacol. 2010, 127, 781–785. [Google Scholar] [CrossRef]
  29. Wei, C.; Wan, L.; Wang, H.; Huang, G. Optimization of Extraction Process for Paeoniae radix Alba by Central Composite Design-Response Surface Methodology. Her. Med. 2016, 175–177. [Google Scholar]
  30. Shen, M.; Zhang, Q.; Qin, L.; Yan, B. Single Standard Substance for the Simultaneous Determination of Eleven Components in the Extract of Paeoniae Radix Alba (Root of Paeonia lactiflora Pall.). J. Anal. Methods Chem. 2021, 2021, 8860776. [Google Scholar] [CrossRef]
  31. Jeon, M.H.; Kwon, H.J.; Jeong, J.S.; Lee, Y.M.; Hong, S.P. Detection of albiflorin and paeoniflorin in Paeoniae radix by reversed-phase high-performance liquid chromatography with pulsed amperometric detection. J. Chromatogr. A 2009, 1216, 4568–4573. [Google Scholar] [CrossRef] [PubMed]
  32. Wang, R.; Peng, X.; Wang, L.; Tan, B.; Liu, J.; Feng, Y.; Yang, S. Preparative purification of peoniflorin and albiflorin from peony rhizome using macroporous resin and medium-pressure liquid chromatography. J. Sep. Sci. 2012, 35, 1985–1992. [Google Scholar] [CrossRef] [PubMed]
  33. Chu, C.; Zhang, S.; Tong, S.; Li, X.; Li, Q.; Yan, J. Elution-extrusion counter-current chromatography for the separation of two pairs of isomeric monoterpenes from Paeoniae Alba Radix. J. Sep. Sci. 2015, 38, 3110–3118. [Google Scholar] [CrossRef] [PubMed]
  34. Wu, Y.; Jiang, Y.; Zhang, L.; Zhou, J.; Yu, Y.; Zhang, S.; Zhou, Y. Green and efficient extraction of total glucosides from Paeonia lactiflora Pall.‘Zhongjiang’ by subcritical water extraction combined with macroporous resin enrichment. Ind. Crops Prod. 2019, 141, 111699. [Google Scholar] [CrossRef]
  35. Liu, E.H.; Qi, L.W.; Li, B.; Peng, Y.B.; Li, P.; Li, C.Y.; Cao, J. High-speed separation and characterization of major constituents in Radix Paeoniae Rubra by fast high-performance liquid chromatography coupled with diode-array detection and time-of-flight mass spectrometry. Rapid Commun. Mass. Spectrom. 2009, 23, 119–130. [Google Scholar] [CrossRef] [PubMed]
  36. Wang, Y.; Wang, P.; Xu, C.; Sun, S.; Zhou, Q.; Shi, Z.; Li, J.; Chen, T.; Li, Z.; Cui, W. Discrimination and chemical characterization of different Paeonia lactifloras (Radix Paeoniae Alba and Radix Paeoniae Rubra) by infrared macro-fingerprint analysis-through-separation. J. Mol. Struct. 2015, 1099, 68–76. [Google Scholar] [CrossRef]
  37. Medzhitov, R.; Horng, T. Transcriptional control of the inflammatory response. Nat. Rev. Immunol. 2009, 9, 692–703. [Google Scholar] [CrossRef] [PubMed]
  38. Xiao, L.; Lin, S.; Zhan, F. One of the active ingredients in Paeoniae Radix Alba functions as JAK1 inhibitor in rheumatoid arthritis. Front. Pharmacol. 2022, 13, 906763. [Google Scholar] [CrossRef] [PubMed]
  39. Sang, X.; Ying, J.; Wan, X.; Han, X.; Shan, Q.; Lyu, Q.; Yang, Q.; Wang, K.; Hao, M.; Liu, E.; et al. Screening of Bioactive Fraction of Radix Paeoniae Alba and Enhancing Anti-Allergic Asthma by Stir-Frying Through Regulating PI3K/AKT Signaling Pathway. Front. Pharmacol. 2022, 13, 863403. [Google Scholar] [CrossRef]
  40. Li, P.P.; Liu, D.D.; Liu, Y.J.; Song, S.S.; Wang, Q.T.; Chang, Y.; Wu, Y.J.; Chen, J.Y.; Zhao, W.D.; Zhang, L.L.; et al. BAFF/BAFF-R involved in antibodies production of rats with collagen-induced arthritis via PI3K-Akt-mTOR signaling and the regulation of paeoniflorin. J. Ethnopharmacol. 2012, 141, 290–300. [Google Scholar] [CrossRef]
  41. Wang, T.; Zhou, X.; Kuang, G.; Jiang, R.; Guo, X.; Wu, S.; Wan, J.; Yin, L. Paeoniflorin modulates oxidative stress, inflammation and hepatic stellate cells activation to alleviate CCl4-induced hepatic fibrosis by upregulation of heme oxygenase-1 in mice. J. Pharm. Pharmacol. 2021, 73, 338–346. [Google Scholar] [CrossRef] [PubMed]
  42. Qian, W.; Zhang, J.; Wang, W.; Wang, T.; Liu, M.; Yang, M.; Sun, Z.; Li, X.; Li, Y. Antimicrobial and antibiofilm activities of paeoniflorin against carbapenem-resistant Klebsiella pneumoniae. J. Appl. Microbiol. 2020, 128, 401–413. [Google Scholar] [CrossRef] [PubMed]
  43. Ho, J.Y.; Chang, H.W.; Lin, C.F.; Liu, C.J.; Hsieh, C.F.; Horng, J.T. Characterization of the anti-influenza activity of the Chinese herbal plant Paeonia lactiflora. Viruses 2014, 6, 1861–1875. [Google Scholar] [CrossRef] [PubMed]
  44. Wang, H.; Zhou, H.; Wang, C.X.; Li, Y.S.; Xie, H.Y.; Luo, J.D.; Zhou, Y. Paeoniflorin inhibits growth of human colorectal carcinoma HT 29 cells in vitro and in vivo. Food Chem. Toxicol. 2012, 50, 1560–1567. [Google Scholar] [CrossRef]
  45. Zheng, M.; Liu, C.; Fan, Y.; Shi, D.; Jian, W. Total glucosides of paeony (TGP) extracted from Radix Paeoniae Alba exerts neuroprotective effects in MPTP-induced experimental parkinsonism by regulating the cAMP/PKA/CREB signaling pathway. J. Ethnopharmacol. 2019, 245, 112182. [Google Scholar] [CrossRef] [PubMed]
  46. Liu, M.; Feng, J.; Du, Q.; Ai, J.; Lv, Z. Paeoniflorin Attenuates Myocardial Fibrosis in Isoprenaline-induced Chronic Heart Failure Rats via Inhibiting P38 MAPK Pathway. Curr. Med. Sci. 2020, 40, 307–312. [Google Scholar] [CrossRef] [PubMed]
  47. Jiang, D.; Chen, Y.; Hou, X.; Xu, J.; Mu, X.; Chen, W. Influence of Paeonia lactiflora roots extract on cAMP-phosphodiesterase activity and related anti-inflammatory action. J. Ethnopharmacol. 2011, 137, 914–920. [Google Scholar] [CrossRef] [PubMed]
  48. Liao, T.; Kang, J.; Ma, Z.; Jie, L.; Feng, M.; Liu, D.; Mao, J.; Wang, P.; Xing, R. Total glucosides of white paeony capsule alleviate articular cartilage degeneration and aberrant subchondral bone remodeling in knee osteoarthritis. Phytother. Res. 2024. [Google Scholar] [CrossRef] [PubMed]
  49. Hu, P.F.; Sun, F.F.; Jiang, L.F.; Bao, J.P.; Wu, L.D. Paeoniflorin inhibits IL-1β-induced MMP secretion via the NF-κB pathway in chondrocytes. Exp. Ther. Med. 2018, 16, 1513–1519. [Google Scholar] [CrossRef]
  50. Hu, P.F.; Chen, W.P.; Bao, J.P.; Wu, L.D. Paeoniflorin inhibits IL-1β-induced chondrocyte apoptosis by regulating the Bax/Bcl-2/caspase-3 signaling pathway. Mol. Med. Rep. 2018, 17, 6194–6200. [Google Scholar] [CrossRef]
  51. Wu, L.; Tang, R.; Xiong, W.; Song, S.; Guo, Q.; Zhang, Q. Paeoniflorin shows chondroprotective effects under IL-1β stress by regulating circ-PREX1/miR-140-3p/WNT5B axis. J. Orthop. Surg. Res. 2023, 18, 766. [Google Scholar] [CrossRef] [PubMed]
  52. Zhao, L.; Chang, Q.; Huang, T.; Huang, C. Paeoniflorin inhibits IL-1β-induced expression of inflammatory mediators in human osteoarthritic chondrocyte. Mol. Med. Rep. 2018, 17, 3306–3311. [Google Scholar] [CrossRef] [PubMed]
  53. Chao Wei, C.; Yi Shan, D.; Fei Yue, L.; Tian You, F. Total Glucosides of Peony Alleviate Joint Destruction in Rabbits with Osteoarthritis. Nat. Prod. Commun. 2023, 18, 1934578X231192211. [Google Scholar] [CrossRef]
  54. Ahmad, M.; Malik, K.; Tariq, A.; Zhang, G.; Yaseen, G.; Rashid, N.; Sultana, S.; Zafar, M.; Ullah, K.; Khan, M.P.Z. Botany, ethnomedicines, phytochemistry and pharmacology of Himalayan paeony (Paeonia emodi Royle.). J. Ethnopharmacol. 2018, 220, 197–219. [Google Scholar] [CrossRef] [PubMed]
  55. Luo, J.; Jin, D.E.; Yang, G.Y.; Zhang, Y.Z.; Wang, J.M.; Kong, W.P.; Tao, Q.W. Total glucosides of paeony for rheumatoid arthritis: A systematic review of randomized controlled trials. Complement. Ther. Med. 2017, 34, 46–56. [Google Scholar] [CrossRef] [PubMed]
  56. Martel-Pelletier, J.; Barr, A.J.; Cicuttini, F.M.; Conaghan, P.G.; Cooper, C.; Goldring, M.B.; Goldring, S.R.; Jones, G.; Teichtahl, A.J.; Pelletier, J.-P. Osteoarthritis. Nat. Rev. Dis. Primers 2016, 13, 16072. [Google Scholar] [CrossRef] [PubMed]
  57. Krustev, E.; Rioux, D.; McDougall, J.J. Mechanisms and mediators that drive arthritis pain. Curr. Osteoporos. Rep. 2015, 13, 216–224. [Google Scholar] [CrossRef] [PubMed]
  58. Huang, H.; Lou, Z.; Zheng, S.; Wu, J.; Yao, Q.; Chen, R.; Kou, L.; Chen, D. Intra-articular drug delivery systems for osteoarthritis therapy: Shifting from sustained release to enhancing penetration into cartilage. Drug Deliv. 2022, 29, 767–791. [Google Scholar] [CrossRef] [PubMed]
  59. Zhong, W.; Chen, J.; Li, Y.; Liu, M.; Yang, S. Efficacy and safety of traditional Chinese medicine rehabilitation program in the treatment of knee osteoarthritis: A randomized controlled trial protocol. Ann. Palliat. Med. 2021, 10, 6909–6918. [Google Scholar] [CrossRef]
  60. Molnar, V.; Matišić, V.; Kodvanj, I.; Bjelica, R.; Jeleč, Ž.; Hudetz, D.; Rod, E.; Čukelj, F.; Vrdoljak, T.; Vidović, D. Cytokines and chemokines involved in osteoarthritis pathogenesis. Int. J. Mol. Sci. 2021, 22, 9208. [Google Scholar] [CrossRef]
  61. Robinson, W.H.; Lepus, C.M.; Wang, Q.; Raghu, H.; Mao, R.; Lindstrom, T.M.; Sokolove, J. Low-grade inflammation as a key mediator of the pathogenesis of osteoarthritis. Nat. Rev. Rheumatol. 2016, 12, 580–592. [Google Scholar] [CrossRef]
  62. Zhang, W.; Li, P.; Song, D.; Niu, H.; Shi, S.; Wang, S.; Duan, J. Structural characterization and biological activities of two α-glucans from Radix Paeoniae Alba. Glycoconj. J. 2016, 33, 147–157. [Google Scholar] [CrossRef] [PubMed]
  63. Chen, C.; Yin, Q.; Tian, J.; Gao, X.; Qin, X.; Du, G.; Zhou, Y. Studies on the Changes of Pharmacokinetics Behaviors of Phytochemicals and the Influence on Endogenous Metabolites After the Combination of Radix Bupleuri and Radix Paeoniae Alba Based on Multi-Component Pharmacokinetics and Metabolomics. Front. Pharmacol. 2021, 12, 630970. [Google Scholar] [CrossRef] [PubMed]
  64. Zhao, S.; Li, S. Network-based relating pharmacological and genomic spaces for drug target identification. PLoS ONE 2010, 5, e11764. [Google Scholar] [CrossRef]
  65. Lee, W.S.; Lee, E.G.; Sung, M.S.; Yoo, W.H. Kaempferol inhibits IL-1β-stimulated, RANKL-mediated osteoclastogenesis via downregulation of MAPKs, c-Fos, and NFATc1. Inflammation 2014, 37, 1221–1230. [Google Scholar] [CrossRef]
  66. Zhuang, Z.; Ye, G.; Huang, B. Kaempferol alleviates the interleukin-1β-induced inflammation in rat osteoarthritis chondrocytes via suppression of NF-κB. Med. Sci. Monit. Int. Med. J. Exp. Clin. Res. 2017, 23, 3925. [Google Scholar] [CrossRef]
  67. Jiang, R.; Hao, P.; Yu, G.; Liu, C.; Yu, C.; Huang, Y.; Wang, Y. Kaempferol protects chondrogenic ATDC5 cells against inflammatory injury triggered by lipopolysaccharide through down-regulating miR-146a. Int. Immunopharmacol. 2019, 69, 373–381. [Google Scholar] [CrossRef]
  68. Yimam, M.; Lee, Y.-C.; Jiao, P.; Hong, M.; Nam, J.-B.; Brownell, L.; Hyun, E.; Jia, Q. UP1306, a botanical composition with analgesic and anti-inflammatory effect. Pharmacogn. Res. 2016, 8, 186. [Google Scholar] [CrossRef]
  69. Du, W.; Liang, X.; Wang, S.; Lee, P.; Zhang, Y. The underlying mechanism of Paeonia lactiflora Pall. in Parkinson’s disease based on a network pharmacology approach. Front. Pharmacol. 2020, 11, 581984. [Google Scholar] [CrossRef]
  70. Fang, H.; Beier, F. Mouse models of osteoarthritis: Modelling risk factors and assessing outcomes. Nat. Rev. Rheumatol. 2014, 10, 413–421. [Google Scholar] [CrossRef]
  71. Gu, R.; Liu, N.; Luo, S.; Huang, W.; Zha, Z.; Yang, J. MicroRNA-9 regulates the development of knee osteoarthritis through the NF-kappaB1 pathway in chondrocytes. Medicine 2016, 95, e4315. [Google Scholar] [CrossRef] [PubMed]
  72. Lu, J.; Zhang, H.; Pan, J.; Hu, Z.; Liu, L.; Liu, Y.; Yu, X.; Bai, X.; Cai, D.; Zhang, H. Fargesin ameliorates osteoarthritis via macrophage reprogramming by downregulating MAPK and NF-κB pathways. Arthritis Res. Ther. 2021, 23, 142. [Google Scholar] [CrossRef] [PubMed]
  73. Jha, B.; Jza, B.; Jwa, B.; Quan, C.; Wd, D.; Ffa, B.; Hya, B.; Sya, B.; Hjab, C.; Ptab, E. Loganin ameliorates cartilage degeneration and osteoarthritis development in an osteoarthritis mouse model through inhibition of NF-κB activity and pyroptosis in chondrocytes. J. Ethnopharmacol. 2020, 247, 112261. [Google Scholar] [CrossRef] [PubMed]
  74. Gao, H.; Peng, L.; Li, C.; Ji, Q.; Li, P. Salidroside Alleviates Cartilage Degeneration Through NF-κB Pathway in Osteoarthritis Rats. Drug Des. Dev. Ther. 2020, 14, 1445–1454. [Google Scholar] [CrossRef] [PubMed]
  75. McGettrick, A.F.; O’Neill, L.A.J. The Role of HIF in Immunity and Inflammation. Cell Metab. 2020, 32, 524–536. [Google Scholar] [CrossRef] [PubMed]
  76. Yudoh, K.; Nakamura, H.; Masukohongo, K.; Kato, T.; Nishioka, K. Catabolic stress induces expression of hypoxiainducible factor (HIF)-1α in articular chondrocytes: Involvement of HIF-1α in the pathogenesis of osteoarthritis. Arthritis Res. Ther. 2005, 7, 225. [Google Scholar] [CrossRef]
  77. Pf, D.; Swoboda, B.; Cramer, T. Commentary The role of HIF-1α in maintaining cartilage homeostasis and during the pathogenesis of osteoarthritis. Arthritis Res. Ther. 2006, 8, 104. [Google Scholar] [CrossRef]
  78. Chang, J.; Jackson, S.G.; Wardale, J.; Jones, S.W. Hypoxia Modulates the Phenotype of Osteoblasts Isolated From Knee Osteoarthritis Patients, Leading to Undermineralized Bone Nodule Formation. Arthritis Rheum. 2014, 66, 1789–1799. [Google Scholar] [CrossRef]
  79. Hu, S.; Zhang, C.; Ni, L.; Huang, C.; Chen, D.; Shi, K.; Jin, H.; Zhang, K.; Li, Y.; Xie, L. Stabilization of HIF-1α alleviates osteoarthritis via enhancing mitophagy. Cell Death Dis. 2020, 11, 481. [Google Scholar] [CrossRef]
  80. Hamilton, J.L.; Nagao, M.; Levine, B.R.; Chen, D.; Olsen, B.R.; Im, H.J. Targeting VEGF and Its Receptors for the Treatment of Osteoarthritis and Associated Pain. J. Bone Min. Res. 2016, 31, 911–924. [Google Scholar] [CrossRef]
  81. Qian, J.J.; Xu, Q.; Xu, W.M.; Cai, R.; Huang, G.C. Expression of VEGF-A Signaling Pathway in Cartilage of ACLT-induced Osteoarthritis Mouse Model. J. Orthop. Surg. Res. 2021, 16, 379. [Google Scholar] [CrossRef] [PubMed]
  82. Wang, Y.H.; Kuo, S.J.; Liu, S.C.; Wang, S.W.; Tsai, C.H.; Fong, Y.C.; Tang, C.H. Apelin Affects the Progression of Osteoarthritis by Regulating VEGF-Dependent Angiogenesis and miR-150-5p Expression in Human Synovial Fibroblasts. Cells 2020, 9, 594. [Google Scholar] [CrossRef] [PubMed]
  83. Su, L.; Wang, Y.; Bao, Y.; Liu, X.; Xu, H. LncRNA MEG3 promotes recovery of knee osteoarthritis in rats through regulating VEGF expression. Panminerva Med. 2020. [Google Scholar] [CrossRef] [PubMed]
  84. Ru, J.; Li, P.; Wang, J.; Zhou, W.; Li, B.; Huang, C.; Li, P.; Guo, Z.; Tao, W.; Yang, Y.; et al. TCMSP: A database of systems pharmacology for drug discovery from herbal medicines. J. Cheminform 2014, 6, 13. [Google Scholar] [CrossRef] [PubMed]
  85. Wishart, D.S.; Knox, C.; Guo, A.C.; Shrivastava, S.; Hassanali, M.; Stothard, P.; Chang, Z.; Woolsey, J. DrugBank: A comprehensive resource for in silico drug discovery and exploration. Nucleic Acids Res. 2006, 34, D668–D672. [Google Scholar] [CrossRef] [PubMed]
  86. Wang, Y.; Xiao, J.; Suzek, T.O.; Zhang, J.; Wang, J.; Bryant, S.H. PubChem: A public information system for analyzing bioactivities of small molecules. Nucleic Acids Res. 2009, 37, W623–W633. [Google Scholar] [CrossRef] [PubMed]
  87. Szklarczyk, D.; Franceschini, A.; Wyder, S.; Forslund, K.; Heller, D.; Huerta-Cepas, J.; Simonovic, M.; Roth, A.; Santos, A.; Tsafou, K.P.; et al. STRING v10: Protein-protein interaction networks, integrated over the tree of life. Nucleic Acids Res. 2015, 43, D447–D452. [Google Scholar] [CrossRef] [PubMed]
  88. Hamosh, A.; Scott, A.F.; Amberger, J.S.; Bocchini, C.A.; McKusick, V.A. Online Mendelian Inheritance in Man (OMIM), a knowledgebase of human genes and genetic disorders. Nucleic Acids Res. 2005, 33, D514–D517. [Google Scholar] [CrossRef]
  89. Qiu, J. Traditional medicine: A culture in the balance. Nature 2007, 448, 126–128. [Google Scholar] [CrossRef]
Figure 1. The results of potential target genes of PRA therapy for OA. (A) Ven diagram of OA and PRA intersection targets; (B) 232 targets related to OA were utilized to construct the PPI network; (C) The top 10 targets ranked by degree value were identified as the kernel targets.
Figure 1. The results of potential target genes of PRA therapy for OA. (A) Ven diagram of OA and PRA intersection targets; (B) 232 targets related to OA were utilized to construct the PPI network; (C) The top 10 targets ranked by degree value were identified as the kernel targets.
Separations 11 00184 g001
Figure 2. The PRA–OA–potential target gene network. The gray connecting lines indicate that each node is interconnected.
Figure 2. The PRA–OA–potential target gene network. The gray connecting lines indicate that each node is interconnected.
Separations 11 00184 g002
Figure 3. Functional enrichment analysis of the top 20 key gene ontology (GO) of PRA in the treatment of OA. (A) GO function analysis, including biological process (BP), cellular component (CC), and molecular function (MF). In the histogram, the common axis represents the degree of enrichment; (B) Bubble plot of Biological Processes (BP); (C) Bubble plot of Molecular Functions (MF); (D) Bubble plot of Cellular Components (CC). The top 20 key enriched GO terms are illustrated as bubble plots. GO terms with amended p-value < 0.01 were believed to be importantly enriched. X-axis, gene ratio (the ratio of targets in the background terms. bubble size, number of genes enriched. bubble color, p-value).
Figure 3. Functional enrichment analysis of the top 20 key gene ontology (GO) of PRA in the treatment of OA. (A) GO function analysis, including biological process (BP), cellular component (CC), and molecular function (MF). In the histogram, the common axis represents the degree of enrichment; (B) Bubble plot of Biological Processes (BP); (C) Bubble plot of Molecular Functions (MF); (D) Bubble plot of Cellular Components (CC). The top 20 key enriched GO terms are illustrated as bubble plots. GO terms with amended p-value < 0.01 were believed to be importantly enriched. X-axis, gene ratio (the ratio of targets in the background terms. bubble size, number of genes enriched. bubble color, p-value).
Separations 11 00184 g003aSeparations 11 00184 g003b
Figure 4. Kyoto Encyclopedia of Genes and Genomes (KEGG) pathway enrichment analyses of PRA for OA. (A) The first 20 significantly enriched KEGG pathways are shown as bubble points. KEGG pathways with regulated p-value < 0.01 were greatly enriched. The x-axis, count. Bubble size, massive genes enriched. Bubble color, p-value; (B) Enrichment analysis of KEGG pathway related to the therapy of OA with PRA.
Figure 4. Kyoto Encyclopedia of Genes and Genomes (KEGG) pathway enrichment analyses of PRA for OA. (A) The first 20 significantly enriched KEGG pathways are shown as bubble points. KEGG pathways with regulated p-value < 0.01 were greatly enriched. The x-axis, count. Bubble size, massive genes enriched. Bubble color, p-value; (B) Enrichment analysis of KEGG pathway related to the therapy of OA with PRA.
Separations 11 00184 g004
Figure 5. The NF-kappa B signaling pathway of potential target genes relationship between PRA and OA. The purple arrows mainly indicate the upstream–downstream relationship between genes. The red is a PRA target gene in the network.
Figure 5. The NF-kappa B signaling pathway of potential target genes relationship between PRA and OA. The purple arrows mainly indicate the upstream–downstream relationship between genes. The red is a PRA target gene in the network.
Separations 11 00184 g005
Figure 6. Heat map of binding energies 8 active components and 10 core genes in PRA (A,B).
Figure 6. Heat map of binding energies 8 active components and 10 core genes in PRA (A,B).
Separations 11 00184 g006
Figure 7. The 6 best docking results of 3 compounds and 2 core targets. The docking between kaempferol and PTGS2 (A), (+)-catechin and PTGS2 (B), DPHCD and PTGS2 (C), kaempferol and MMP9 (D), (+)-catechin and MMP9 (E), DPHCD and MMP9 (F). The yellow structural formula represents the active ingredient. The protein targets bound to the active ingredient are shown as purple rods, and the binding sites are linked by yellow hydrogen bonds.
Figure 7. The 6 best docking results of 3 compounds and 2 core targets. The docking between kaempferol and PTGS2 (A), (+)-catechin and PTGS2 (B), DPHCD and PTGS2 (C), kaempferol and MMP9 (D), (+)-catechin and MMP9 (E), DPHCD and MMP9 (F). The yellow structural formula represents the active ingredient. The protein targets bound to the active ingredient are shown as purple rods, and the binding sites are linked by yellow hydrogen bonds.
Separations 11 00184 g007
Table 1. Characteristics of active ingredients in PRA.
Table 1. Characteristics of active ingredients in PRA.
NoMol IDCompoundMolecular
Structure
Molecular
Weight
AlogPHdonHaccOB (%)DLHL
1MOL00191011α,12α-epoxy-3ꞵ-23-dihydroxy-30-norolean-20-en-28,12ꞵ-olideSeparations 11 00184 i001470.713.912564.770.42.62
2MOL001918PaeoniflorgenoneSeparations 11 00184 i002318.350.791687.590.47.45
3MOL001919DPHCDSeparations 11 00184 i003358.522.692443.560.54.34
4MOL001921LactiflorinSeparations 11 00184 i004462.49−0.5731049.120.87.26
5MOL001924PaeoniflorinSeparations 11 00184 i005480.51−1.2851153.870.813.9
6MOL001925paeoniflorin_qtSeparations 11 00184 i006318.350.462668.180.48.81
7MOL001928albiflorin_qtSeparations 11 00184 i007318.350.422666.640.36.54
8MOL001930benzoyl paeoniflorinSeparations 11 00184 i008584.620.8141231.270.8−1.9
9MOL000211MairinSeparations 11 00184 i009456.786.522355.380.88.87
10MOL000358beta-sitosterolSeparations 11 00184 i010414.798.081136.910.85.36
11MOL000359SitosterolSeparations 11 00184 i011414.798.081136.910.85.37
12MOL000422KaempferolSeparations 11 00184 i012286.251.774641.880.214.7
13MOL000492(+)-catechinSeparations 11 00184 i013290.291.925654.830.20.61
Table 2. 232 potential target genes of PRA therapy for OA.
Table 2. 232 potential target genes of PRA therapy for OA.
NOTargetSymbolNOTargetSymbol
1Gamma-aminobutyric acid receptor subunit alpha-1GABRA1117Insulin-like growth factor I receptorIGF1R
2Progesterone receptorPGR118Tyrosine-protein kinase SRCSRC
3Androgen ReceptorAR119Vascular endothelial growth factor receptor 2KDR
4Cytochrome P450 19A1CYP19A1120MAP kinase p38 alphaMAPK14
5Testis-specific androgen-binding proteinSHBG121Peroxisome proliferator-activated receptor gammaPPARG
6Glucocorticoid receptorNR3C1122ThrombinF2
7Estrogen receptor alphaESR1123Nitric-oxide synthase, endothelialNOS3
8Tyrosine-protein kinase JAK1JAK1124Coagulation factor VIIF7
9Tyrosine-protein kinase JAK2JAK2125Transcription factor p65RELA
10Vanilloid receptorTRPV1126Inhibitor of nuclear factor kappa-B kinase subunit betaIKBKB
11Cytochrome P450 2D6CYP2D6127RAC-alpha serine/threonine-protein kinaseAKT1
12Cytochrome P450 2C9CYP2C9128Tumor necrosis factorTNF
13Tyrosine-protein kinase JAK3JAK3129Xanthine dehydrogenase/oxidaseXDH
14Tyrosine-protein kinase TYK2TYK2130Interstitial collagenaseMMP1
15Delta opioid receptor (by homology)OPRD1131Signal transducer and activator of transcription 1-alpha/betaSTAT1
16Kappa Opioid receptor (by homology)OPRK1132Heme oxygenase 1HMOX1
17Serotonin 1a (5-HT1a) receptorHTR1A133Cytochrome P450 3A4CYP3A4
18Steryl-sulfataseSTS134CYP1A2CYP1A2
19Poly [ADP-ribose] polymerase-1PARP1135Cytochrome P450 1A1CYP1A1
20c-Jun N-terminal kinase 1MAPK8136Intercellular adhesion molecule 1ICAM1
21Glycogen synthase kinase-3 betaGSK3B137E-selectinSELE
22c-Jun N-terminal kinase 3MAPK10138Vascular cell adhesion protein 1VCAM1
23Dual specificity mitogen-activated protein kinase kinase 4MAP2K4139Cytochrome P450 1B1CYP1B1
24MAP kinase p38 deltaMAPK13140Arachidonate 5-lipoxygenaseALOX5
25Cyclin-dependent kinase 2/cyclin ACDK2141Hyaluronan synthase 2HAS2
26Dual specificity mitogen-activated protein kinase kinase 7MAP2K7142Glutathione S-transferase PGSTP1
27c-Jun N-terminal kinase 2MAPK9143Aryl hydrocarbon receptorAHR
28CDC7CDC7144Solute carrier family 2, facilitated glucose transporter member 4SLC2A4
29Interleukin-6 receptor subunit betaIL6ST145Type I iodothyronine deiodinaseDIO1
30Estrogen receptor betaESR2146Glutathione S-transferase Mu 1GSTM1
31Catechol O-methyltransferaseCOMT147AntileukoproteinaseSLPI
32C-C chemokine receptor type 1CCR1148NADPH oxidase 4NOX4
33Estrogen-related receptor alphaESRRA149Aldose reductase (by homology)AKR1B1
34Estrogen-related receptor betaESRRB150TyrosinaseTYR
35Serine/threonine-protein kinase RIPK2RIPK2151Multidrug resistance-associated protein 1ABCC1
36Muscarinic acetylcholine receptor M2CHRM2152P-glycoprotein 1ABCB1
37Thyroid hormone receptor alphaTHRA153ATP-binding cassette sub-family G member 2ABCG2
38Dipeptidyl peptidase IVDPP4154Monoamine oxidase AMAOA
39Endothelin receptor ET-A (by homology)EDNRA155Tyrosine-protein kinase SYKSYK
40Thrombin and coagulation factor XF10156Matrix metalloproteinase 9MMP9
41Cyclin-dependent kinase 1CDK1157Matrix metalloproteinase 2MMP2
42Carbonic anhydrase IICA2158Arachidonate 15-lipoxygenaseALOX15
43Carbonic anhydrase ICA1159Adenosine A2a receptor (by homology)ADORA2A
44P2X purinoceptor 3P2RX3160TransthyretinTTR
45GABA-A receptor; alpha-3/beta-3/gamma-2GABRB3161Tankyrase-2TNKS
46GABA-A receptor; alpha-2/beta-3/gamma-2GABRA2162Casein kinase II alphaCSNK2A1
47Mitogen-activated protein kinase kinase kinase 5MAP3K5163Epidermal growth factor receptor erbB1EGFR
48Cannabinoid receptor 2CNR2164MyeloperoxidaseMPO
49TYRO3TYRO3165PI3-kinase p85-alpha subunitPIK3R1
50Type-1 angiotensin II receptor (by homology)AGTR1166Focal adhesion kinase 1PTK2
51DNA-dependent protein kinasePRKDC167Matrix metalloproteinase 13MMP13
52Cannabinoid receptor 1CNR1168Matrix metalloproteinase 3MMP3
53Rho-associated protein kinase 1ROCK1169Carbonic anhydrase IIICA3
54Sodium channel protein type I× alpha subunitSCN9A170Serine/threonine-protein kinase PLK1PLK1
55Lysine-specific histone demethylase 1KDM1A171Hepatocyte growth factor receptorMET
56Telomerase reverse transcriptaseTERT172Interleukin-8 receptor ACXCR1
57Protein-tyrosine phosphatase 2CPTPN11173CaM kinase II betaCAMK2B
58Toll-like receptor (TLR7/TLR9)TLR9174Tyrosine-protein kinase receptor UFOAXL
59Heat shock factor protein 1HSF1175Aldo-keto reductase family 1 member C2 (by homology)AKR1C2
60Nerve growth factor receptor Trk-ANTRK1176Aldo-keto reductase family 1 member C1 (by homology)AKR1C1
61Protein kinase C deltaPRKCD177Aldo-keto reductase family 1 member C4 (by homology)AKR1C4
62Prostaglandin G/H synthase 1PTGS1178Beta-amyloid A4 proteinAPP
63Prostaglandin G/H synthase 2PTGS2179Matrix metalloproteinase 12MMP12
64Phosphatidylinositol-4,5-bisphosphate 3-kinase catalytic subunit, gamma isoformPIK3CG180CD38CD38
655-hydroxytryptamine 2A receptorHTR2A181Cystic fibrosis transmembrane conductance regulatorCFTR
66Gamma-aminobutyric-acid receptor alpha-5 subunitGABRA51826-phosphofructo-2-kinase/fructose-2,6-bisphosphatase 3PFKFB3
67Gamma-aminobutyric-acid receptor alpha-3 subunitGABRA3183G protein-coupled receptor kinase 6GRK6
68Beta-2 adrenergic receptorADRB2184Microtubule-associated protein tauMAPT
69Sodium-dependent serotonin transporterSLC6A4185CATCAT
70Mu-type opioid receptorOPRM1186Bile acid receptor FXRNR1H4
71Neuronal acetylcholine receptor protein, alpha-7 chainCHRNA7187G-protein coupled bile acid receptor 1GPBAR1
72Apoptosis regulator Bcl-2BCL21885-lipoxygenase activating proteinALOX5AP
73BAXBAX189Fatty acid binding protein intestinalFABP2
74Caspase-9CASP9190Anandamide amidohydrolaseFAAH
75Caspase-3CASP3191Solute carrier family 22 member 6 (by homology)SLC22A6
76Caspase-8CASP8192Angiotensin-converting enzymeACE
77Transforming growth factor beta-1TGFB1193Retinoic acid receptor betaRARB
78Serum paraoxonase/arylesterase 1PON1194Retinoid X receptor alphaRXRA
79HMG-CoA reductaseHMGCR195Cytosolic phospholipase A2PLA2G4A
80LXR-alphaNR1H3196Integrin alpha-4/beta-1ITGB1
81Sterol regulatory element-binding protein 2SREBF2197G protein-coupled receptor 44PTGDR2
82Cytochrome P450 2C19CYP2C19198Integrin alpha-V/beta-3ITGAV
83ButyrylcholinesteraseBCHE199Leukocyte common antigenPTPRC
84Nuclear receptor ROR-alphaRORA200Endothelin-converting enzyme 1ECE1
85Protein-tyrosine phosphatase 1BPTPN1201AMP deaminase 2AMPD2
86Corticosteroid binding globulinSERPINA6202Interleukin-6IL6
87Glucose-6-phosphate 1-dehydrogenaseG6PD203Monocyte differentiation antigen CD14CD14
88Nuclear receptor subfamily 1 group I member 3 (by homology)NR1I3204Lipopolysaccharide-binding proteinLBP
89AcetylcholinesteraseACHE205Galectin-3LGALS3
90Vitamin D receptorVDR206Galectin-9LGALS9
91LXR-betaNR1H2207Heat shock protein HSP 90-alphaHSP90AA1
92Prostanoid EP1 receptor (by homology)PTGER1208Somatostatin receptor 2SSTR2
93Prostanoid EP2 receptor (by homology)PTGER2209Somatostatin receptor 1SSTR1
9411-beta-hydroxysteroid dehydrogenase 1HSD11B1210Vascular endothelial growth factor AVEGFA
95Glycine receptor subunit alpha-1GLRA1211Acidic fibroblast growth factorFGF1
96Peroxisome proliferator-activated receptor deltaPPARD212Basic fibroblast growth factorFGF2
97Nitric oxide synthase, inducible (by homology)NOS2213HeparanaseHPSE
98UDP-glucuronosyltransferase 2B7UGT2B7214Plasminogen activator inhibitor-1SERPINE1
9911-beta-hydroxysteroid dehydrogenase 2HSD11B2215Matrix metalloproteinase 7MMP7
100Prolyl endopeptidasePREP216Matrix metalloproteinase 8MMP8
101Prostanoid EP4 receptor (by homology)PTGER4217Beta-glucocerebrosidaseGBA
102Indoleamine 2,3-dioxygenaseIDO1218Hexokinase type IIHK2
103Protein Mdm4MDM4219Histone deacetylase 1HDAC1
104p53-binding protein Mdm-2MDM2220ADAM17ADAM17
105Prostanoid IP receptorPTGIR221Somatostatin receptor 5SSTR5
106Fatty acid binding protein adipocyteFABP4222Fibroblast growth factor receptor 1FGFR1
107Prostaglandin E synthasePTGES223Glyceraldehyde-3-phosphate dehydrogenase liverGAPDH
108T-cell protein-tyrosine phosphatasePTPN2224Lysosomal alpha-glucosidaseGAA
109MAP kinase ERK1MAPK3225Heat shock cognate 71 kDa proteinHSPA8
110Aldo-keto reductase family 1 member B10AKR1B10226DNA (cytosine-5)-methyltransferase 1DNMT1
111Phospholipase A2 group 1BPLA2G1B227Adenosine deaminaseADA
112Protein kinase C etaPRKCH228Mitogen-activated protein kinase kinase kinase 7MAP3K7
113Neurokinin 1 receptor (by homology)TACR1229Thymidine kinase, cytosolicTK1
114Peroxisome proliferator-activated receptor alphaPPARA230S-adenosylmethionine decarboxylase 1AMD1
115Proteinase-activated receptor 1F2R231Thymidine phosphorylaseTYMP
116ALK tyrosine kinase receptorALK232AICAR transformylaseATIC
Table 3. Top 19 high-degree genes in the network.
Table 3. Top 19 high-degree genes in the network.
NOTargetSymbolBetweenness CentralityCloseness CentralityDegree
1Progesterone receptorPGR0.0006498640.5031315246
2Androgen ReceptorAR0.0006498640.5031315246
3Cytochrome P450 19A1CYP19A10.0006498640.5031315246
4Estrogen receptor alphaESR10.0021906250.5031315246
5Estrogen receptor betaESR20.0006498640.5031315246
6Type-1 angiotensin II receptor (by homology)AGTR10.0007661020.5031315246
7Prostaglandin G/H synthase 2PTGS20.0017909720.5010395016
8Testis-specific androgen-binding proteinSHBG0.0004342230.5010395015
9Glucocorticoid receptorNR3C10.0004342230.5010395015
10Muscarinic acetylcholine receptor M2CHRM20.000465550.5010395015
11Gamma-aminobutyric-acid receptor alpha-2 subunitGABRA20.0004618640.5010395015
12Protein-tyrosine phosphatase 2CPTPN110.0004342230.5010395015
13Prostaglandin G/H synthase 1PTGS10.0017909720.5010395015
14Protein-tyrosine phosphatase 1BPTPN10.0005199940.5010395015
15Nitric oxide synthase, inducible (by homology)NOS20.0004146110.5010395015
16Aldo-keto reductase family 1 member B10AKR1B100.0004146110.5010395015
17Phospholipase A2 group 1BPLA2G1B0.0004146110.5010395015
18Tyrosine-protein kinase SRCSRC0.0005424480.5010395015
19Vascular endothelial growth factor receptor 2KDR0.0005424480.5010395015
Table 4. KEGG pathway analysis results.
Table 4. KEGG pathway analysis results.
Pathway NameSymbol
Pathways in cancerAGTR1, AKT1, ALK, AR, BAX, BCL2, CAMK2B, CASP3, CASP8, CASP9, CDK2, EDNRA, EGFR, ESR1, ESR2, F2, F2R, FGF1, FGF2, FGFR1, GSK3B, GSTM1, GSTP1, HDAC1, HMOX1, HSP90AA1, IGF1R, IKBKB, IL6, IL6ST, ITGAV, ITGB1, JAK1, JAK2, JAK3, MDM2, MET, MMP1, MMP2, MMP9, NOS2, NTRK1, PIK3R1, PPARD, PPARG, MAPK3, MAPK8, MAPK9, MAPK10, PTGER1, PTGER2, PTGER4, PTGS2, PTK2, RARB, RELA, ROCK1, RXRA, STAT1, TERT, TGFB1, VEGFA
Lipid and atherosclerosisAKT1, BAX, BCL2, CAMK2B, CASP3, CASP8, CASP9, CD14, MAPK14, CYP1A1, CYP2C9, GSK3B, HSPA8, HSP90AA1, ICAM1, IKBKB, IL6, JAK2, LBP, MAP3K5, MMP1, MMP3, MMP9, NOS3, PIK3R1, PPARG, MAPK3, MAPK8, MAPK9, MAPK10, MAPK13, MAP2K7, PTK2, RELA, RXRA, SELE, MAP2K4, SRC, MAP3K7, TNF, VCAM1, GSTM1, GSTP1, HMOX1, ITGAV, KDR, MMP2, VEGFA, AGTR1, SERPINE1, PRKCD, STAT1, TGFB1, NOX4, CCR1, FGF2, IL6ST, JAK1, PIK3CG, PTGS2, SYK, TYK2, CDK2, JAK3, AHR, CAT, AKR1C4, CYP1A2, CYP1B1, AKR1C1, AKR1C2, EGFR, MET, PTPN1, PTPN11, ALOX5, ITGB1, NOS2, VDR, RIPK2, TLR9, HDAC1, MDM2, CSNK2A1, MMP13, CYP2D6, ESR1, ESR2, IGF1R, RORA, NTRK1, ACE, PPARA, HK2, ROCK1, GAPDH, PTPRC, NR1H3, F2, ADAM17, F2R, SLC2A4, NR1H2, PARP1, G6PD, ALOX5AP, PLA2G4A, CXCR1, HTR2A, PRKCH, PTGER2, PTGER4, TRPV1, ABCC1, OPRD1, ALK, CDK1, APP, CHRNA7, MAPT, SSTR1, SSTR2, SSTR5, MMP7, TERT, COMT, MAOA
Chemical carcinogenesis—receptor activationADRB2, AHR, AKT1, AR, BCL2, CHRNA7, CYP1A1, CYP1A2, CYP1B1, CYP3A4, EGFR, ESR1, ESR2, FGF2, GSTM1, HSP90AA1, JAK2, PGR, PIK3R1, PPARA, MAPK3, RELA, RXRA, SRC, UGT2B7, VDR, VEGFA, NR1I3, HSPA8, MMP2, MMP9, NOS3, OPRM1, PRKCD
Neuroactive ligand receptor interactionADORA2A, ADRB2, AGTR1, CHRM2, CHRNA7, CNR1, CNR2, EDNRA, F2, F2R, GABRA1, GABRA2, GABRA3, GABRA5, GABRB3, GLRA1, NR3C1, HTR1A, HTR2A, OPRD1, OPRK1, OPRM1, P2RX3, PTGER1, PTGER2, PTGER4, PTGIR, SSTR1, SSTR2, SSTR5, TACR1, THRA, TRPV1, CAMK2B, CD38, EGFR, FGF1, FGF2, FGFR1, KDR, MET, NOS2, NOS3, NTRK1, VEGFA
Proteoglycans in cancerAKT1, CAMK2B, CASP3, MAPK14, EGFR, ESR1, FGF2, FGFR1, IGF1R, ITGAV, ITGB1, KDR, MDM2, MET, MMP2, MMP9, PIK3R1, MAPK3, MAPK13, PTK2, PTPN11, ROCK1, SRC, TGFB1, TNF, VEGFA, HPSE, CD14, FGF1, HSPA8, IKBKB, MAPT, MAP3K5, NTRK1, PLA2G4A, MAPK8, MAPK9, MAPK10, MAP2K7, RELA, MAP2K4, MAP3K7, AXL, BAX, BCL2, GSK3B, IL6, JAK1, JAK2, PLA2G1B, ADORA2A, CNR1, F2R
HIF-1 signaling pathwayAKT1, BCL2, CAMK2B, EGFR, GAPDH, HK2, HMOX1, IGF1R, IL6, NOS2, NOS3, SERPINE1, PFKFB3, PIK3R1, MAPK3, RELA, VEGFA
FoxO signaling pathwayAKT1, CAT, CDK2, MAPK14, EGFR, IGF1R, IKBKB, IL6, MDM2, PIK3R1, PLK1, MAPK3, MAPK8, MAPK9, MAPK10, MAPK13, SLC2A4, TGFB1, CDK1, HSP90AA1, PGR, SERPINE1, RELA, AR, CAMK2B
cAMP signaling pathwayADORA2A, ADRB2, AKT1, CAMK2B, CFTR, CHRM2, EDNRA, F2R, HTR1A, PIK3R1, PPARA, MAPK3, MAPK8, MAPK9, MAPK10, PTGER2, RELA, ROCK1, SSTR1, SSTR2, SSTR5
Serotonergic synapseALOX5, ALOX15, APP, CASP3, CYP2C19, CYP2C9, CYP2D6, GABRB3, HTR1A, HTR2A, MAOA, PLA2G4A, MAPK3, PTGS1, PTGS2, SLC6A4
VEGF signaling pathwayAKT1, CASP9, MAPK14, KDR, NOS3, PIK3R1, PLA2G4A, MAPK3, MAPK13, PTGS2, PTK2, SRC, VEGFA, F2, F2R, ITGB1, PIK3CG, PTGIR, PTGS1, ROCK1, SYK, AGTR1, EGFR, CXCR1, PTPN11, ACHE, BCL2, CAMK2B, CHRM2, CHRNA7, JAK2, HDAC1, MPO, RELA, MAP3K7, CD38, ADRB2, EDNRA, OPRD1, NOS2, SERPINE1
Steroid hormone biosynthesisSTS, AKR1C4, COMT, CYP1A1, CYP1A2, CYP1B1, CYP3A4, CYP19A1, AKR1C1, AKR1C2, HSD11B1, HSD11B2, UGT2B7, CYP2C19, CYP2C9, GSTM1, GSTP1, PTGS2, CYP2D6, MAOA, CAT, IDO1
NecroptosisPARP1, ALOX15, BAX, BCL2, CAMK2B, CASP8, HSP90AA1, JAK1, JAK2, JAK3, PLA2G4A, MAPK8, MAPK9, MAPK10, STAT1, TNF, TYK2, CASP3, CASP9, CSNK2A1, GSK3B, MMP7, PPARD, MAP3K7, CFTR, ITGB1, MAP3K5, MAP2K7, ROCK1, SRC, ADORA2A, MAOA, MAPT, HSPA8, RELA, HDAC1, PPARG
Gastric cancerAKT1, BAX, BCL2, CDK2, EGFR, FGF1, FGF2, GSK3B, MET, ABCB1, PIK3R1, MAPK3, RARB, RXRA, TERT, TGFB1, FGFR1, IGF1R, MDM2, ALK, CASP9, JAK3, ESR1, ESR2, PGR, HDAC1, IKBKB, PTPN11, RELA, GSTM1, GSTP1, HMOX1, G6PD, HK2, NTRK1, CD14, MPO, PPARD, CAMK2B, VEGFA, SYK, PLA2G4A, PRKCD, PTPRC, TNF, HSD11B2
Chemokine signaling pathwayAKT1, CCR1, GRK6, GSK3B, IKBKB, CXCR1, JAK2, JAK3, PIK3CG, PIK3R1, PRKCD, MAPK3, PTK2, RELA, ROCK1, SRC, STAT1, CASP9, DIO1, ESR1, HDAC1, ITGAV, MDM2, RXRA, THRA
Platinum drug resistanceAKT1, BAX, BCL2, CASP3, CASP8, CASP9, GSTM1, GSTP1, MDM2, MAP3K5, PIK3R1, MAPK3, CDK1, CDK2, HDAC1, IL6ST, JAK1, JAK3, RELA, SRC, SYK, MDM4, SERPINE1
NF-kappa B signaling pathwayPARP1, BCL2, CD14, CSNK2A1, ICAM1, IKBKB, LBP, PTGS2, RELA, SYK, MAP3K7, TNF, VCAM1
MicroRNAs in cancerBCL2, CASP3, CYP1B1, DNMT1, EGFR, HDAC1, HMOX1, IKBKB, MDM2, MDM4, MET, MMP9, ABCC1, ABCB1, PIK3R1, MAPK3, PTGS2, ROCK1, VEGFA
Signaling pathways regulating pluripotency of stem cellsAKT1, MAPK14, ESRRB, FGF2, FGFR1, GSK3B, IGF1R, IL6ST, JAK1, JAK2, JAK3, PIK3R1, MAPK3, MAPK13, BCL2, EGFR, IL6, PTPN2, PTPN11, STAT1, TYK2
Retrograde endocannabinoid signalingCNR1, MAPK14, FAAH, GABRA1, GABRA2, GABRA3, GABRA5, GABRB3, MAPK3, MAPK8, MAPK9, MAPK10, MAPK13, PTGS2, CHRNA7, HTR1A, P2RX3, SCN9A, GRK6, OPRM1, SRC
Bladder cancerTYMP, EGFR, MDM2, MMP1, MMP2, MMP9, MAPK3, SRC, VEGFA, CSNK2A1, FGFR1, IGF1R, MET, PTPN1, MAP3K7, CDK1, HTR2A, GRK6, HSPA8, CXCR1
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

Wang, B.; Bai, C.; Zhang, Y. Paeoniae Radix Alba and Network Pharmacology Approach for Osteoarthritis: A Review. Separations 2024, 11, 184. https://doi.org/10.3390/separations11060184

AMA Style

Wang B, Bai C, Zhang Y. Paeoniae Radix Alba and Network Pharmacology Approach for Osteoarthritis: A Review. Separations. 2024; 11(6):184. https://doi.org/10.3390/separations11060184

Chicago/Turabian Style

Wang, Bo, Changcai Bai, and Yuanyuan Zhang. 2024. "Paeoniae Radix Alba and Network Pharmacology Approach for Osteoarthritis: A Review" Separations 11, no. 6: 184. https://doi.org/10.3390/separations11060184

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