Next Article in Journal
Novel HSA-PMEMA Nanomicelles Prepared via Site-Specific In Situ Polymerization-Induced Self-Assembly for Improved Intracellular Delivery of Paclitaxel
Previous Article in Journal
Assessing Drug–Drug Interaction and Food Effect for BCS Class 2 Compound BI 730357 (Retinoic Acid-Related Orphan Receptor Gamma Antagonist, Bevurogant) Using a Physiology-Based Pharmacokinetics Modeling (PBPK) Approach with Semi-Mechanistic Absorption
Previous Article in Special Issue
Discovery of Non-Peptide GLP-1 Positive Allosteric Modulators from Natural Products: Virtual Screening, Molecular Dynamics, ADMET Profiling, Repurposing, and Chemical Scaffolds Identification
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Computer-Aided Drug Design in Research on Chinese Materia Medica: Methods, Applications, Advantages, and Challenges

1
Key Laboratory of Fermentation Engineering (Ministry of Education), Cooperative Innovation Centre of Industrial Fermentation (Ministry of Education & Hubei Province), Hubei University of Technology, Wuhan 430068, China
2
School of Chinese Herbal Medicine, Guangzhou University of Chinese Medicine, Guangzhou 510006, China
*
Author to whom correspondence should be addressed.
Pharmaceutics 2025, 17(3), 315; https://doi.org/10.3390/pharmaceutics17030315
Submission received: 12 February 2025 / Revised: 27 February 2025 / Accepted: 28 February 2025 / Published: 1 March 2025
(This article belongs to the Special Issue Computer-Aided Development: Recent Advances and Expectations)

Abstract

:
Chinese materia medica (CMM) refers to the medicinal substances used in traditional Chinese medicine. In recent years, CMM has become globally prevalent, and scientific research on CMM has increasingly garnered attention. Computer-aided drug design (CADD) has been employed in Western medicine research for many years, contributing significantly to its progress. However, the role of CADD in CMM research has not been systematically reviewed. This review briefly introduces CADD methods in CMM research from the perspectives of computational chemistry (including quantum chemistry, molecular mechanics, and quantum mechanics/molecular mechanics) and informatics (including cheminformatics, bioinformatics, and data mining). Then, it provides an exhaustive discussion of the applications of these CADD methods in CMM research through rich cases. Finally, the review outlines the advantages and challenges of CADD in CMM research. In conclusion, despite the current challenges, CADD still offers unique advantages over traditional experiments. With the development of the CMM industry and computer science, especially driven by artificial intelligence, CADD is poised to play an increasingly pivotal role in advancing CMM research.

1. Introduction

Chinese materia medica (CMM) refers to the medicinal substances used in traditional Chinese medicine (TCM) clinical practice and is primarily derived from plants, animals, and minerals. CMM has been used for disease prevention, treatment, and rehabilitation for thousands of years in China. In September 2022, the National Administration of Traditional Chinese Medicine announced that CMM had spread to 196 countries and regions, indicating its growing worldwide acceptance [1]. Nowadays, CMM is recognized by the World Health Organization as a complementary alternative therapy approach [2].
To further facilitate the application and dissemination of CMM, the Chinese government has adopted a supportive stance toward academic research on CMM. Data from the National Natural Science Foundation of China (NSFC) show that the Chinese government has gradually increased its funding for CMM research in recent years. In 2023, the NSFC’s funding for TCM research surpassed CNY 770 million [3]. Moreover, since 2016, the Chinese government has released a series of seminal documents to support the development of TCM [4]. Owing to this official support, academic research on CMM has become a popular scientific field. A typical CMM study may employ analytical techniques to identify specific ingredients from CMM extracts, biochemical experiments to assess the potential activity of these ingredients, and pharmacological experiments to confirm their mechanisms of action (Figure 1) [5]. This approach allows researchers to screen CMM active ingredients (CAIs) and elucidate the properties and actions of a CMM herb.
Nevertheless, CMM formulae are the main means of health maintenance and disease treatment in TCM clinical practice, and they treat diseases through multitarget interventions involving multiple ingredients, differing from Western medical practices. The combination of various CMM herbs in a formula can produce synergistic or antagonistic effects, significantly complicating CMM research [6]. Consequently, typical research methods face significant challenges because of their expense and time-consuming nature, resulting in a relatively weak foundation for CMM research.
To strengthen CMM research, computer-aided drug design (CADD) has been applied in recent years. The concept of CADD was developed as early as the 1970s and it has already been used in Western medicine (WM) research [7]. Initially, CADD referred specifically to molecular simulation techniques based on computational chemistry, which can be subdivided into structure-based drug design (SBDD) and ligand-based drug design (LBDD) [8]. After decades of development, the concept of CADD has expanded significantly, and many studies refer to various computer-aided techniques used in pharmacy research as CADD, including artificial intelligence drug design (AIDD) based on machine learning (ML) and deep learning (DL), which has gained much attention in recent years [9]. Recently, several drugs developed through CADD have been approved by the US Food and Drug Administration (FDA), thereby contributing to the growing prominence of CADD in CMM research [10]. However, due to the relatively late application of CADD in CMM research, there are few comprehensive reviews specifically addressing this topic.
Therefore, this review omits the detailed definition of CADD and instead briefly describes the CMM research-related CADD methods from the two main perspectives of computational chemistry and informatics. For these methods, this review exhaustively elaborates on the application cases that are closely related to CMM research. Finally, the review discusses the advantages and challenges of applying CADD to CMM research. This review will bridge the gap between traditional CMM theories and modern computational approaches, providing a roadmap for future interdisciplinary research.

2. Computational Chemistry in CMM Research

Computational chemistry is a branch of theoretical chemistry and a core aspect of CADD. It uses mathematical approximations and computer programs to address specific chemical problems. The computational chemistry methods related to CMM research mainly include quantum chemistry (QC), molecular mechanics (MM), and quantum mechanics/molecular mechanics (QM/MM).

2.1. QC

2.1.1. Brief Introduction to QC

QC employs the fundamental principles and methods of quantum mechanics (QM) to study and address chemical problems. QM posits that the motion of microscopic systems can be described by the Schrödinger equation. The solution to this equation, known as the wave function, theoretically describes the motion state of a microscopic system and the various properties determined by this state [11]. Therefore, the development of solving methods for the Schrödinger equation has always been a pivotal aspect in quantum chemical calculation (QCC). Typical QCC methods are primarily divided into semi-empirical methods, ab initio calculations, and density functional theory (DFT). Among these three methods, DFT generally offers a good balance between computational costs and accuracy, establishing itself as the prevailing method in QCC [11].
The workflow of a QCC study is illustrated in Figure 2. The initial step involves providing the molecular structure, followed by selecting an appropriate computational method. For DFT calculations, the choice of an appropriate functional is critical. After selecting the method, it is usually necessary to set up the base set for the calculation. Some semi-empirical methods, which already have built-in basis sets, do not require additional basis sets. Based on the molecular structure and the chosen method, the QCC program can approximate the wave function, which contains a wealth of chemically significant information, such as structural, thermodynamic, and kinetic parameters. Currently, common QCC programs include Gaussian 09, ORCA 5.0, and xTB 6.7.1 [12,13,14].

2.1.2. Applications of QC in CMM Research

  • Conformation and spectra analysis of CAIs
Investigating the molecular structures of CAIs is crucial in understanding the material basis of CMM. However, the experimental characterization of flexible and complex structures is often challenging. In such cases, conformation searching using QCC offers a robust alternative, facilitating the exploration of potential energy surfaces and the identification of low-energy conformations (Figure 2). Typically, researchers employ less computationally intensive methods, such as MM or semi-empirical methods, for initial structure optimization and energy comparison across numerous conformations. Subsequently, they apply DFT calculations, which are more accurate but computationally demanding, to confirm the dominant conformations [15,16]. Conformational analysis provides precise spatial structural parameters of CAIs, including bond lengths, bond angles, and dihedral angles, often in strong agreement with experimental data. By integrating QCC with experimental data, researchers can achieve enhanced accuracy in characterizing CAIs or complex systems containing them. This significantly reduces the complexity of structural characterization [17]. However, QCC methods based on electronic structure calculations are computationally expensive and unsuitable for large-scale dataset screening. To address these challenges, researchers have developed artificial intelligence (AI) techniques to accelerate the QCC process. Lu et al. introduced Uni-Mol+ (https://github.com/deepmodeling/Uni-Mol/, accessed on 27 February 2025), a DL approach that generates an initial 3D conformation using RDKit (https://www.rdkit.org/, accessed on 27 February 2025) and iteratively refines it toward the DFT-equilibrium conformation via neural networks [18].
Based on accurately characterized conformations, QCC can calculate various spectra of CAIs, including circular dichroism (CD), nuclear magnetic resonance (NMR), Raman, X-ray, ultraviolet–visible (UV–Vis), and infrared (IR) spectra [19,20,21,22,23]. Beyond direct spectral calculations, QCC can also indirectly predict spectra for CAIs. Wang et al. demonstrated a correlation between DFT-calculated energy data and the mass spectrometry (MS)-based fragment abundances of isomeric CAIs, providing a robust theoretical framework for the understanding of the mechanisms underlying fragment ion differences among isomers [24]. Li et al. calculated the dipole moments of isomeric CAIs using DFT, successfully explaining the chromatographic retention times of these isomers [25].
Recent advancements in QCC-based spectral analysis highlight the potential of integrating AI to improve the calculation accuracy while significantly reducing the computational costs [26]. A prominent example is the structure elucidation of natural products (NPs) using ML-assisted quantum chemical NMR calculations, which has emerged as a key area of interest. Tsai et al. applied kernel ridge regression (KRR) to refine NRM shielding constant calculations from DFT, enhancing the precision of the predicted chemical shifts. This approach enables the rapid determination of NP configurations within minutes [27]. Since many NPs also contain CAIs, the incorporation of AI into QCC-based spectral analysis holds significant promise in advancing CMM research.
2.
Physicochemical analysis of CAIs
The conformation and spectral analysis of CAIs provide a robust foundation for an understanding of their bioactivity. Nevertheless, structural information alone is insufficient to fully elucidate bioactivity; it is equally important to investigate the physicochemical properties of CAIs.
Over the past decade, researchers have utilized QCC to study the key physicochemical properties of CAIs (Figure 2). Zeng et al. calculated the pKa values of 96 carboxylic acids in an aqueous solution, establishing a reliable procedure that accurately reproduces theoretical pKa values in close agreement with experimental data [28]. Li et al. employed DFT to calculate the polarity of CAIs, successfully predicting their chromatographic retention times [25]. Guan et al. developed a QC model to predict the octanol–water partition coefficients (LogP) of small organic molecules, including CAIs, as part of the SAMPL6 LogP Prediction Challenge [29].
Physicochemical properties such as pKa, polarity, and LogP are relatively straightforward to calculate using QCC. Many other properties require QCC to model specialized systems or parameters for indirect prediction. Yang et al. investigated solute–solvent and solvent–solvent interactions by calculating the solvation free energy and analyzing radial distribution functions to elucidate the solubility of benzoin in three solvent mixtures [30]. Many studies have assessed the stability of CAIs by evaluating the bond energies of key chemical bonds, as well as the energy difference between the highest occupied molecular orbital (HOMO) and the lowest unoccupied molecular orbital (LUMO), which is known as the HOMO–LUMO gap, an approximation of the fundamental gap and optical gap [31,32,33].
The physicochemical properties are closely related to drug-like properties, such as absorption, distribution, metabolism, excretion, and toxicity (ADMET), and are mainly derived from QCC-based structural and thermodynamic parameters [34]. Additionally, QCC-based kinetic parameters can provide valuable insights into the physicochemical properties of CAIs. For instance, calculations of the transition states, intrinsic reaction coordinates (IRC), and reaction rates provide a deeper understanding of the reactivity of CAIs. Such insights elucidate their biosynthetic pathways in vivo and inform their synthesis and structural modification in vitro [32,35,36].
In fact, based on structural, thermodynamic, and kinetic parameters, QCC can predict an even broader range of physicochemical properties, such as the redox potential, density, enthalpy of evaporation, boiling point, polarizability, and magnetic moment. These predictions are widely applied in WM research and related disciplines [37,38,39,40,41,42,43]. Since both CAIs and WMs are fundamentally chemical molecules, QCC is poised to play an increasingly significant role in the physicochemical analysis of CAIs.
3.
Bioactivity analysis of CAIs
Traditional physicochemical properties are typically measurable through experimental methods. However, the wave functions generated by QCC can be analyzed to derive numerous QC parameters that, while purely theoretical, are essential in evaluating the bioactivity of CAIs (Figure 2). Wave function analysis is commonly performed using software tools such as GaussView 6 and Multiwfn 3.8 [44,45].
Chen et al. confirmed the antioxidant activity, sites, mechanisms, and products of several CAIs via experiments and various QC parameters, such as molecular surface electrostatic potentials (ESP), electron density differences, and spin densities [32]. Liu et al. investigated the natural population analysis (NPA) charges, frontier molecular orbitals, molecular ESP, and chemical reactivity descriptors for evodiamine and rutaecarpine, based on which the potential anti-cancer activity and mechanisms of the CAIs were predicted [46]. Zhang et al. predicted the anti-inflammatory activity of several CAIs by elucidating their interactions with related targets based on the ESP, HOMO energy, and LUMO energy [47]. QC parameters also facilitate the evaluation of a broad spectrum of bioactivity, including anti-viral, anti-obesity, and anti-diabetic effects. These evaluations are primarily conducted by analyzing the physicochemical properties of CAIs and their interactions with biological targets [48,49,50].
4.
Quantitative structure–property relationship (QSPR) and quantitative structure–activity relationship (QSAR) analysis of CAIs
Since the physicochemical properties and bioactivity of CAIs can be analyzed using QCC, many studies have tried to predict the physicochemical properties and bioactivity of molecules using QCC-aided mathematical models. These models are referred to as quantitative structure–property relationship (QSPR) and quantitative structure–activity relationship (QSAR) models, respectively (Figure 2).
In QCC-aided QSPR/QSAR studies, the structural, thermodynamic, kinetic, and QC parameters are commonly used as descriptors. These descriptors serve as input variables in statistical models to establish correlations with specific properties or types of activity, enabling the creation of predictive models that inform drug design and development. Currently, QCC-based QSPR/QSAR studies focused on CAIs remain limited, with only a handful published sporadically over the past two decades. This scarcity may be attributed to the time-intensive nature of QCC and experimental research on large datasets of CAIs [51,52,53,54].
Conversely, QCC-aided QSPR/QSAR studies are significantly more prevalent in WM and other disciplines, such as analytical, materials, and environmental chemistry. These fields benefit from the compatibility, clear physicochemical relevance, and high accuracy of QCC descriptors [53]. Notable advancements have been made, particularly in the development of databases and online tools that incorporate QC descriptors into predictive modeling processes [55,56,57]. The integration of ML and DL algorithms into QSPR/QSAR analysis has further enhanced the predictive accuracy and applicability by capturing complex nonlinear relationships and managing large-scale, high-dimensional data [58]. With the availability of such databases and tools, researchers often favor direct property or activity predictions over de novo QSPR/QSAR modeling for CAIs [31,59].

2.2. MM

2.2.1. Brief Introduction to MM

QCC is inherently complex and primarily applied to small molecules. MM relies on classical Newtonian mechanics, using simplified functional forms to describe potential interactions between atoms [60]. Unlike QCC, MM does not account for electronic motion, treating entire atoms—or even groups of atoms—as single particles. The simplicity of MM enables extremely rapid calculations, making it particularly suitable for studying large molecular systems.
The core of MM lies in the development of a force field, which consists of energy terms and their specific functional forms. These energy terms are typically categorized into bonding and non-bonding terms. Bonding terms represent interactions between adjacent atoms within a molecule, while non-bonding terms capture interactions between molecules and between non-adjacent atoms within a molecule. Several molecular force fields commonly used in CMM-related molecular systems are summarized in Table 1. When MM calculations incorporate the temporal evolution of molecular states, the method is referred to as molecular dynamics (MD). MD solves Newton’s equations of motion to determine the positions and velocities of atoms over time, providing dynamic insights into molecular behavior [61]. Research methodologies derived from MM and MD, such as molecular docking and molecular dynamics simulation (MDS), are collectively known as molecular simulation techniques.
Molecular docking is a technique used to predict interactions between two or more molecules. Widely used molecular docking programs include AutoDock 4.2.6, AutoDock Vina 1.1.2, Schrödinger Glide 2024-4, MOE 2019, and DOCK 4.0 [91,92,93,94,95]. The typical process for the use of these programs involves the following steps: (1) the preparation and preprocessing of the ligand structure (e.g., a medicinal small molecule) and the receptor structure (e.g., a biological macromolecule); (2) setting the docking parameters, such as searching algorithms, the number of runs, and docking areas; (3) searching for the optimal binding modes and calculating the docking score, a numerical value that quantifies the binding affinity; (4) analyzing and visualizing the results, including the system structure, binding sites, and binding mechanisms (e.g., hydrogen bonding, hydrophobic interactions, and electrostatic interactions).
In subsequent studies involving molecular docking, MDS is commonly employed. This is because molecular docking generates a static model of the ligand–receptor complex, which cannot capture the dynamic nature of the binding process in a specific environment. MDS addresses the limitations of molecular docking and traditional experiments by offering insights into the dynamic behavior of molecular interactions over time, under specified conditions regarding the solvent, temperature, and pressure. MDS is typically performed using specialized programs like Amber 24, Gromacs 2020.3, and GROMOS 1.6.0, with the basic workflow depicted in Figure 3 [96,97,98].

2.2.2. Applications of MM in CMM Research

  • Virtual screening of CAIs
Over the past two decades, more than 100,000 publications on molecular docking have been recorded across platforms such as Google Scholar, Web of Science, PubMed, and the China National Knowledge Infrastructure (CNKI). Notably, approximately half of the 20,000 CNKI publications originate from the field of traditional Chinese pharmacy, highlighting the extensive application of molecular docking in CMM research.
Since its introduction in the 1970s, molecular docking has become a cornerstone technique in CADD. In SBDD, virtual screening plays a pivotal role, leveraging molecular docking and other computational tools to identify potential drug candidates. When target information is available, molecular docking facilitates the identification process. Conversely, when target information is unavailable but drug activity is known, MM-based molecular descriptors can be utilized for QSPR/QSAR analysis, which constitutes a primary approach in LBDD. Recently, the widespread adoption of MM-derived methods has enabled the integration of both SBDD and LBDD into CMM research [99,100]. While the key steps of QSPR/QSAR analysis in LBDD were discussed in the preceding section, this section focuses on virtual screening within the framework of SBDD.
Literature research indicates that molecular docking has been employed for the virtual screening of CAIs since the early 2000s. In 2004, Zhang et al. elucidated the pharmacological mechanism of a natural potassium channel blocker derived from the venom of Buthus martensii Karsch [101]. Two years later, in 2006, Gao et al. used human immunodeficiency virus (HIV)-1 protease as a receptor to screen potential anti-HIV drugs from a TCM database [102]. Despite the time that has elapsed since these studies, they continue to exemplify two primary applications of molecular docking in CMM research: (1) the virtual screening of CAIs based on binding mechanisms and (2) the virtual screening of CAIs based on binding affinities. The former approach relies on the structural characteristics of docked complexes, while the latter utilizes quantitative metrics, such as scoring functions, to evaluate the binding ability.
In recent years, molecular docking has undergone significant advancements, primarily in four key areas: enhanced computational efficiency, improved scoring functions, expanded docking systems (e.g., flexible macromolecules, metal-containing molecules, and covalent-linked complexes), and increased interdisciplinary integration (e.g., combining molecular docking with QC, AI, and big data) [103,104]. These developments have also contributed to the progress of virtual screening for CAIs [105,106].
As previously noted, MDS is frequently employed as a follow-up to molecular docking studies. Using trajectories generated from MDS, the binding free energy of a ligand–receptor complex can be calculated and decomposed through methods such as thermodynamic integration, free energy perturbation, linear interaction energy, and molecular mechanics Poisson–Boltzmann/generalized Born surface area (MM/PB(GB)SA, Figure 3) [107,108,109,110]. From a theoretical perspective, MDS and binding free energy calculations offer greater accuracy compared to molecular docking but are significantly more complex and computationally intensive. As a result, researchers commonly adopt a tiered approach: molecular docking is used for initial high-throughput screening, followed by MDS and experimental validation to refine the screening process. This integrated strategy has been extensively applied in the virtual screening of CAIs [111,112].
2.
MDS for various systems related to CMM research
The virtual screening of CAIs typically requires known targets, such as proteins, enzymes, and DNA. However, the interactions between CAIs and their targets are not always limited to virtual screening purposes. Zhang et al. utilized MDS to analyze the intermolecular interactions between various CAIs and bovine serum albumin, providing insights into CAIs’ penetration during the cross-flow ultrafiltration of CMM solutions [113]. In fact, when studying the penetration effects of CAIs, more studies select bio-membranes rather than proteins as models (Figure 3). These studies have not only enhanced the bioavailability of CAIs but also facilitated the development of penetration enhancers derived from CAIs [114,115,116]. In addition to penetration enhancers, other CMM-related pharmaceutical excipients have been explored using MDS. Zhou et al. employed MDS to simulate the interaction mechanism between binders and powders, offering guidance for the industrial production of CMM granules [117]. Shen et al. used MDS to screen the excipients and investigate the blend system of ginsenoside K nanostructured lipid carriers [118]. Dai et al. used MDS to study the self-assembly process of ginsenoside Ro, providing a theoretical foundation for vesicle formation in ginsenoside Ro and other CMM saponins, which show potential as surfactants and solubilizers [119]. These studies highlight the applicability of MDS for CAIs in complex solution systems such as solvent mixtures, CAIs interacting with pharmaceutical excipients, and the self-assembly of CAIs (Figure 3).
The self-assembly of CAIs occurs frequently, both in vitro and in vivo. Lei et al. performed MDS to simulate the self-assembly of berberine and glycyrrhizic acid, revealing that the process is driven by hydrogen bonding, π–π stacking, and electrostatic interactions to form a hydrogel. This hydrogel demonstrated injectability, safety, favorable release properties, and a significant anti-inflammatory effect in treating ulcerative colitis [120]. Zhang et al. provided evidence that supramolecular pigments could form in vivo during CMM metabolism via simultaneous covalent and non-covalent assembly, potentially playing crucial roles in pharmacological activity. Their study employed coarse-grained and all-atom MDS, both of which aligned well with the experimental findings [121]. CAIs can self-assemble not only with other CAIs but also with molecules such as organic acids and metal ions. These interactions, elucidated through MDS, result in advanced bioactive materials like nanoparticles, micelles, and gels [122]. Derived from renewable resources, these materials are easy to prepare and multifunctional, offering properties like controlled release, smart-responsive release, and potent biological effects for the treatment of various diseases [122]. Thus, such materials retain the benefits of CAIs while overcoming their limitations.
Despite its growing application in CMM research, MDS faces several limitations, including insufficient sampling, inaccuracies in atomistic models, and challenges in analyzing and interpreting trajectories [123]. Recent advances in AI, particularly in ML and DL, present opportunities to address these challenges. Developments in AI have led to the creation of AI-based force fields, improved techniques for conformational space sampling, and innovative methods for trajectory analysis [123].

2.3. QM/MM

2.3.1. Brief Introduction to QM/MM

QM is well-suited for the calculation of properties related to electron motion, such as electronic structures, spectroscopic properties, molecular orbitals, and chemical bond breaking. In contrast, MM is more appropriate for the analysis of the conformations and motions of large molecules. Due to their differing computational principles, QC offers higher accuracy but at the expense of slower computation, whereas MM calculations are faster but less precise.
To combine the strengths of QM and MM, Warshel and Levitt introduced the QM/MM method in 1976 [124]. This hybrid approach partitions the system into three regions: the QM region, the MM region, and the boundary region (Figure 4). The total energy of the system is defined as the sum of the energies of these three components. QM methods are applied to regions where chemical changes occur, such as enzyme binding sites or substrate interactions (i.e., the QM region). MM methods handle the remainder of the system, including receptors and solvent molecules (i.e., the MM region). To address the interface between QM and MM regions (i.e., the boundary region), techniques such as atom-linking methods, boundary atom methods, or local orbital methods are employed to ensure accurate and seamless integration [124]. Currently, there are many QCC and MDS programs that can perform QM/MM calculations independently, such as Gaussian 09, ORCA 5.0, CP2K 2025.1, and Amber 24 [13,96,125]. To combine the strengths of different programs, researchers prefer to use QCC programs in conjunction with MDS programs to perform QM/MM calculations, such as by coupling Gaussian 09 and Gromacs 2020.3 [126]. To facilitate this combination, researchers have also worked on the development of specialized packages to call QCC and MDS programs and integrate commonly used options in QM/MM calculations [127].

2.3.2. Applications of QM/MM in CMM Research

In CMM research, QM/MM is frequently used to simulate the interactions between CAIs and proteins. Due to the application of QCC for key regions, QM/MM offers higher accuracy than MDS alone, providing a deeper understanding of the binding modes between CAIs and the target proteins [128]. A novel molecular docking method based on the QM/MM theory, known as QM-polarized ligand docking, is more accurate than traditional molecular docking. This method is useful for the virtual screening and QSPR/QSAR analysis of CAIs [129]. Moreover, QM/MM is particularly well suited for studying interactions between CAIs and enzymes, offering a powerful tool for the investigation of enzymatic reactions in CMM systems [130,131]. This capability is especially valuable in synthetic biology, where enzymatic reactions in CMM systems guide the biosynthesis pathways of target CAIs [126,132].

3. Informatics

Computational chemistry and traditional experimental approaches have generated vast amounts of data in CMM research. Efficient data processing methods have, in turn, significantly advanced the field of computational chemistry. Cheminformatics and bioinformatics now provide robust theoretical frameworks and practical tools for data processing, while data mining has further strengthened the connection between raw data and actionable information in practical applications. Collectively, these disciplines form the informatics foundation of CADD.

3.1. Cheminformatics and Bioinformatics

3.1.1. Brief Introduction to Cheminformatics and Bioinformatics

Cheminformatics leverages computational tools to represent, manage, analyze, simulate, and disseminate chemical information. One of the primary challenges in this field is efficiently parsing large datasets to extract meaningful molecular information, making the accurate representation of molecular structures a critical focus. This representation is also a prerequisite for computational chemistry studies of CAIs. As demonstrated in the aforementioned cases, cheminformatics was integral to the studies conducted by Zhang and Gao, who utilized 3D conformations and molecular fingerprints, respectively, to represent CAIs. Their computational and experimental studies, grounded in these representations, ultimately achieved their intended objectives [101,102]. Such approaches have become commonplace in recent decades, converting molecular structures into various machine-readable formats, including the simplified molecular input line entry system (SMILES), International Chemical Identifier (InChI), molecular fingerprints, and graphical representations. These formats facilitate the analysis and comparison of molecular structures, as well as virtual screening and QSPR/QSAR analyses for CAIs [133,134].
Initially developed to streamline the drug discovery and development process, cheminformatics has since grown to play an increasingly vital role in biology and biochemistry [135]. From this point of view, cheminformatics and bioinformatics are almost inseparable as the cheminformatics approach is a necessary pathway for bioinformatics, which involves the use of computational methods to analyze biological data. Given this inherent overlap, in addition to the representation of molecular structures by means of cheminformatics, the subsequent application cases will no longer strictly distinguish between cheminformatics and bioinformatics (Figure 5).

3.1.2. Application of Cheminformatics and Bioinformatics in CMM Research

  • TCM database establishment
In CMM research, the integration of cheminformatics and bioinformatics has revolutionized how chemical and biological data are explored and utilized, particularly within the realm of TCM databases [136]. Today, most TCM databases not only include chemical information about CAIs but also encompass TCM-related biological data. Furthermore, some TCM databases provide additional information on botanical sources, formulae, clinical cases, and TCM theories. These comprehensive databases have been made possible by the advanced data processing capabilities of cheminformatics and bioinformatics.
For a long time, TCM databases have significantly contributed to CMM research by offering a valuable source of TCM-related information. Unfortunately, many of these databases have either ceased to be maintained or are no longer accessible. We have compiled a list of free TCM databases and tested their accessibility (Table 2). It is important to note that all the databases in Table 2 explicitly focus on TCM-related information, rather than covering a broader range of NPs. Additionally, we have only included comprehensive TCM databases and excluded others focused on specific diseases, as the number of such databases is too large to cover in this paper. Other well-known databases based on cheminformatics and bioinformatics, such as RCSB PDB (https://www.rcsb.org/, accessed on 27 February 2025), NCBI (https://www.ncbi.nlm.nih.gov/, accessed on 27 February 2025), and GeneCards (https://www.genecards.org/, accessed on 27 February 2025), are not included in our list because they are not specialized TCM databases [136].
While TCM databases provide invaluable resources for CMM research, the discontinuation or inaccessibility of many highlights the urgent need for the sustainable management and continuous updating of these repositories. Moreover, recent trends show growing homogeneity in the development and application of these databases. Several studies have indicated significant data overlap between databases, with some even containing notable errors when compared to experimental results [137]. There is a pressing need to ensure clear and reliable data sources when incorporating them into databases.
Table 2. Specialized databases for traditional Chinese medicine research.
Table 2. Specialized databases for traditional Chinese medicine research.
DatabaseNumber of FormulaeNumber of HerbsNumber of IngredientsNumber of TargetsNumber of CasesWebsiteAccessibility
BATMAN-TCM [138]54,832840439,1712,319,272-http://bionet.ncpsb.org.cn/batman-tcm/#/homeYes
CancerHSP [139]-24393575--https://old.tcmsp-e.com/CancerHSP.phpYes
CEMTDD [140]-62140602163-http://www.cemtdd.com/index.htmlNo
CMAUP [141]-786560,222758-https://www.bidd.group/CMAUP/Yes
CMCR----111,653https://cmcr.yiigle.com/Yes
CPMCP [142]656156027,92820,965-http://cpmcp.topNo
ETCM [143]395940272842266-http://www.tcmip.cn/ETCM/Yes
ETM-DB [144]57310544285--http://biosoft.kaist.ac.kr/etm/home.phpNo
Herb [145]-726349,25812,933-http://herb.ac.cn/Yes
HIT [146]-125012372208-http://www.badd-cao.net:2345/Yes
IGTCM [147]-831033--http://yeyn.group:96/Yes
iTCM [148]25,875845443,43018,851-http://itcm.biotcm.net/Yes
LTM-TCM [149]48,126912234,96713,109-http://cloud.tasly.com/#/tcm/homeNo
SuperTCM [150]-651655,772543-http://tcm.charite.de/supertcmYes
SymMap [151]-69825,97520,965-https://www.symmap.orgNo
TCM@taiwan [152]-35237,170--http://tcm.cmu.edu.tw/No
TCMBank [153]-919361,96532,529-https://tcmbank.cn/No
TCM-ID [154]744327517375768-https://www.bidd.group/TCMID/Yes
TCMID [155]99,58210,84643,413--https://www.megabionet.org/tcmid/No
TCMIO [156]149361816,437400-http://tcmio.xielab.netYes
TCMIP V2.0395940272842266-http://www.tcmip.cn/Yes
TCMM [157]48,043893269,81676,449 www.tcmm.net.cn/zh-hans/Yes
TCM-Mesh [158]-6235383,840--http://mesh.tcm.microbioinformatics.org/No
TCMSID [159]-49920,0153270-https://tcm.scbdd.com/No
TCMSP [160]-49929,3843311-https://old.tcmsp-e.com/tcmsp.phpYes
TCMSSD [161]133,518825943,41317,602-http://tcmssd.ratcm.cn/Yes
TCM-suite [162]66927322704,321--http://tcm-suite.aimicrobiome.cn/Yes
TM-MC [163]507563534,65613,971-https://tm-mc.kr/material.doYes
YaTCM [164]1813622047,69618,697-http://cadd.pharmacy.nankai.edu.cn/yatcm/homeNo
Imedbooks95,2608980---https://www.imedbooks.com/Yes
TCMkb----465,784http://www.tcmkb.cn/consiliaweb/Yes
Shoudao Zhongyi----400,000https://www.shoudaozhongyi.com/Yes
Yian586895--30http://www.zhongyoo.com/yian/Yes
Yideng Xuyan400,000---102,000http://db.yidxy.com/prescriptionsYes
TCMdoc80,00011,239--60,000http://www.tcmdoc.cn/YiAn/index.aspxYes
Note: Many databases specializing in Chinese medicine are not listed in the table because they are accessible for a fee or the statistics are unavailable. The accessibility of all websites was tested on 27 February 2025. The “-” means no data.
2.
Screening of drug targets for CMM
Due to the vastness of TCM databases and beyond, researchers have developed a range of bioinformatics-based methods to screen drug targets for CMM. These methods often integrate computational chemistry with omics techniques, particularly genomics, transcriptomics, and proteomics, to systematically analyze the molecular characteristics and network relationships within the genome, transcriptome, and proteome.
Genomics provides a comprehensive approach to annotating and analyzing genomic data, which facilitates the identification of potential targets for CAIs. Gene ontology (GO) and pathway enrichment analysis are commonly used to link genes with specific biological functions and pathways relevant to CMM [165]. Moreover, genome-wide association studies (GWAS) and network-based methods help to uncover gene–disease associations, establishing a systematic framework for the identification of CMM-targeted genes [166]. Transcriptomics focuses on the study of RNA transcripts, enabling the investigation of gene expression changes induced by CMM treatments [167]. High-throughput RNA sequencing has become an essential tool in transcriptomics, allowing for the identification of differentially expressed genes (DEGs) associated with specific CAIs [168]. Network-based methods like weighted gene co-expression network analysis (WGCNA) provide insights into regulatory modules and their potential drug–target relationships [169]. Proteomics examines the complete set of proteins expressed in a biological system and their dynamic interactions. As CAIs often interact directly with proteins, proteomics plays a crucial role in target prediction. Computational chemistry supports the virtual screening of CAIs against potential protein targets, helping to identify high-affinity interactions [170]. Protein–protein interaction (PPI) networks, constructed using bioinformatics platforms like Cytoscape 3.10.3, allow for the visualization and analysis of CMM-induced changes in protein networks [171,172].
Researchers increasingly prefer to use multi-omics techniques to screen targets for CAIs, rather than relying on specific omics methods. For instance, Guo et al. used a DL model based on graph neural networks to predict and discover novel berberine derivatives that could target Helicobacter pylori. The efficacy and mechanisms of the predicted CAIs were verified using pharmacokinetic and multi-omics approaches [173]. In addition to screening potential targets for individual CAIs, multi-omics techniques are often combined with AI to identify targets for CMM herbs, CMM formulae, and Chinese patent medicines [174,175,176].
3.
CMM network pharmacology (CMM-NP) research
Screening marker drug targets is useful in identifying CAIs and assessing their mechanisms of action. However, this approach alone is insufficient to fully characterize the efficacy and mechanisms of CMM, which exerts its therapeutic effects through multiple CAIs, multiple targets, and multiple pathways. Given the complexity of the signaling networks involved in CMM’s treatment of diseases, a network-based approach is essential for CMM research. In 2007, Li et al. utilized bioinformatics to construct the first biomolecular network for Cold/Hot syndrome in TCM, revealing the network’s regulatory effects of formulae for Cold/Hot syndrome [177]. In the same year, “network pharmacology” was introduced by Andrew L. Hopkins, a pharmacologist at Dundee University in the UK [178].
In the following decades, the application of network pharmacology in CMM research grew rapidly, leading to the emergence of a specialized field known as CMM-NP. A search using the terms “network pharmacology” and “Chinese medicine” in Web of Science reveals over 4700 publications. In the CNKI, searching for “network pharmacology” yields over 18,600 publications in Chinese, more than 16,000 of which are from the field of traditional Chinese pharmacy. These statistics demonstrate that network pharmacology, as a bioinformatics-based research method, has become a pivotal tool in CMM research since its inception.
CMM-NP research generally follows several steps: (1) the identification of CAIs and their potential targets through literature mining and database screening (e.g., the TCM databases listed in Table 2 and other public databases); (2) the use of bioinformatics tools and databases such as GeneCards and OMIM (http://www.omim.org/, accessed on 27 February 2025) to obtain disease targets and construct disease–CAI–target networks; (3) the construction of PPI networks using target data from databases like STRING (https://cn.string-db.org/, accessed on 27 February 2025) to explore potential molecular mechanisms; functional annotation, including GO and Kyoto Encyclopedia of Genes and Genomes (KEGG) pathway enrichment analysis, is then conducted to investigate the biological significance of key targets and related pathways; (4) molecular simulations are employed to validate the interactions between key targets and CAIs, providing theoretical support for an understanding of the mechanisms of CMM.
Compared to the single-omics techniques used to screen drug targets for CAIs, CMM-NP offers a significant throughput advantage, as it can screen multiple CAIs, targets, and pathways simultaneously. These CAIs, targets, and pathways form a complex network, facilitating the systematic exploration of the pharmacodynamics, material basis, and mechanisms of action of CMM [179,180]. Peng et al. applied CMM-NP and experimental investigations to explore the material basis and potential molecular mechanisms of a CMM pair containing Astragali radix and Spatholobi caulis in the treatment of bone marrow suppression. They identified 36 CAIs from the CMM pair and screened eight potential drug targets, suggesting that the CMM pair may be used to treat myelosuppression through various biological processes [181]. Similarly, CMM-NP has been applied in the study of CMM formulae and Chinese patent medicines. In these cases, different CMM herbs are considered as additional nodes in the “disease–CMM herb–CAI–target–pathway” network, providing a novel approach to uncovering and visualizing the underlying interaction networks of CMM in the treatment of multifactorial diseases [182,183].
The holistic and systematic nature of CMM-NP aligns with the overall concept and dialectical treatment of CMM, making it widely applicable for the screening of CAIs, the prediction of targets, and the interpretation of the mechanisms of action of CMM. This approach holds great promise in uncovering new insights in the field of CMM research. With the advent of advanced techniques, particularly omics and AI, CMM-NP has been refined and deeply implemented, demonstrating its potential as the next paradigm in drug discovery [179,180].
4.
CMM toxicity and quality research
Researchers have developed a technique akin to CMM-NP, known as network toxicology, which focuses on toxic CAIs and their associated targets as nodes in the network [184]. This approach typically integrates omics, molecular simulations, and AI to predict the toxicity of CMM and explain the underlying mechanisms through network topology analysis. In the past five years, most network toxicology research in CMM has centered on predicting cardiotoxicity, hepatotoxicity, neurotoxicity, reproductive toxicity, nephrotoxicity, and brain toxicity [185,186,187,188,189,190]. Beyond network toxicology, several bioinformatics-based methods have been developed to predict CMM toxicity, falling into three main categories: (1) traditional QSPR/QSAR models based on computational chemical descriptors, (2) AI models based on available toxicity data from databases and the literature, and (3) AI models based on traditional CMM toxicity theories documented in ancient texts [191,192,193].
Toxicity is a key indicator in evaluating CMM’s safety, while another critical indicator is the CMM quality. Currently, adulteration is widespread in the CMM industry, making the rapid and accurate identification of CMM essential for quality control [194]. Fortunately, CADD technologies, particularly those based on bioinformatics and AI, have enabled significant progress in CMM quality studies. Genomic, transcriptomic, and proteomic data for many CMM herbs are recorded in the GenBank database, greatly assisting in the identification of CMM herbs. For example, DNA barcoding, a bioinformatics-based technology, has gained significant attention in recent years and is widely used for the identification and characterization of CMM herbs. It has also been included in the Chinese Pharmacopoeia [195]. In the CMM industry, CMM herbs with excellent quality, produced in a specific region, are usually referred to as geo-authentic (Daodi in Chinese) herbs. For a long time, the formation mechanisms of geo-authentic herbs have posed challenges for traditional experimental research, but bioinformatics is expected to reduce these difficulties. By analyzing the omics data of CMM herbs, bioinformatics can help to reveal the genetic background and germplasm resource differences of geo-authentic herbs, explore the influence of the environment on gene expression, and identify metabolic markers that are characteristic of production areas [196,197].
The analysis of metabolic markers, often serving as CAIs, is crucial in ensuring CMM quality. In 2016, academician Chang-xiao Liu introduced the concept of quality markers (Q-markers), which are compounds that serve as comprehensive quality indicators [198]. Unlike ordinary CAIs, Q-markers should not only represent the pharmacological activity of CMM but also offer additional characteristics, such as testability, specificity, traceability, and TCM relevance [31]. Since the introduction of the Q-marker concept, research in this area has grown rapidly, contributing positively to the quality evaluation and control of CMM. Through a systematic literature review, we found that CADD techniques, particularly bioinformatics methods, are frequently employed in Q-marker research for CMM. These methods primarily screen key CAIs through CMM-NP and omics, combined with traditional experiments to evaluate whether the CAIs meet the principles of Q-markers [31]. Some studies have even introduced AI algorithms for the automated prediction of Q-markers, although these efforts are still in the preliminary stage [199].

3.2. Data Mining

3.2.1. Brief Introduction to Data Mining

Data mining is the process of extracting meaningful information and patterns from large datasets, making it a crucial component in the fields of data science and computer science. It integrates techniques from statistics, AI, and database management to uncover potential knowledge, relationships, and trends. In the context of CADD, data mining plays a vital role in conjunction with cheminformatics and bioinformatics. This is particularly relevant due to the complexity and vastness of chemical and biological data related to drugs, requiring sophisticated scientific data processing techniques to extract valuable insights. Data mining provides the necessary methodology for this process, with key steps including data collection, data cleaning, data analysis, and data visualization (Figure 6).

3.2.2. Application of Data Mining in CMM Research

In fact, data mining is integral to many of the applications discussed in the previous section. Its use in conjunction with computational chemistry, cheminformatics, and bioinformatics for CMM research is widespread. For instance, data mining plays a crucial role in the establishment of TCM databases, CMM-NP research, multi-omics studies of CMM, and toxicology studies of CMM. In this section, we focus on other significant aspects of data mining in CMM research.
Analyzing the usage patterns of CMM is a common task in data mining for CMM research. By examining these patterns, researchers can gain insights into the interactions between different CMM herbs, optimizing CMM formulae, improving their therapeutic effects, and guiding clinical practice [200,201,202]. The primary raw data sources include the literature, case reports, books, and TCM databases (Figure 6 and Table 2). However, raw data often contain noise due to various factors: (1) homonyms (i.e., different CMM herbs with the same name) and synonyms (i.e., the same CMM herb with different names) are common in TCM clinical practice; (2) diseases in modern medicine may have various names in traditional medicine, referred to as “syndromes” (Zhenghou in Chinese); and (3) raw data may include duplicates and errors. To ensure the usability of the data, multiple rounds of data cleaning by different individuals are essential.
Data analysis involves statistical techniques, such as frequency analysis, cluster analysis, and association rule mining (Figure 6). This process often utilizes specialized software, including common tools like IBM SPSS Statistics 26 and Excel 2016, as well as dedicated platforms for the analysis of CMM formulae data, such as the TCM Inheritance Support System Software Platform V2.5, the Ancient and Modern Medical Cases Cloud Platform V2.3.5, TCM Miner (https://www.tcmminer.com/, accessed on 27 February 2025), and ITCMDAS (https://gitee.com/serendipity_LB/itcmdas, accessed on 27 February 2025) [203,204,205,206]. Based on the results, researchers typically use software or programming languages (e.g., Python 3.10 and R 4.4.3) to visualize the findings through charts (Figure 6).
Data mining is also frequently applied in large-scale case analyses of TCM clinical practice, which significantly enhances medical quality, supports medical education, and informs clinical practice (Figure 6). Guo et al. used data mining to study a ten-year dataset and found that younger people with higher education levels in China are more likely to use CMM when ill [207]. Similarly, Zhou et al. analyzed the follow-up data of 3850 osteoarthritis patients and found that CMM treatment significantly improved the immune–inflammatory indices of these patients [208]. Moreover, Lam et al. conducted a data mining survey of 54 clinical studies and concluded that CMM is an effective adjuvant therapy for pediatric cancer patients [209]. Although there is no direct correlation between these studies and CMM, CMM represents the drugs used in TCM clinical practice, and these studies have indirectly guided the clinical application of and scientific research on CMM.

4. Advantages and Challenges of CADD in CMM Research

4.1. Advantages of CADD in CMM Research

The growing application of CADD in CMM research underscores the transformative potential of modern computational tools in advancing the understanding and development of ancient medical practices. This potential is evident in several key areas that highlight the advantages of CADD in CMM research (Figure 7).

4.1.1. Enhancing the Accuracy and Reliability of CMM Research

Numerous case studies discussed above demonstrate that computational chemistry can yield predictive outcomes that significantly enhance the study of CAIs. The mutual validation between computational chemistry and experimental results in various scenarios strengthens the accuracy and reliability of CMM research. More importantly, the multi-ingredient, multi-target, and multi-pathway nature of CMM herbs, formulae, and Chinese patent medicines is difficult to verify through traditional experimental methods. However, informatics and AI approaches align with TCM’s holistic and systemic perspectives, offering a more robust and comprehensive understanding. Through extensive research cases, we have illustrated how these CADD techniques improve the accuracy and reliability of CMM research.
CADD also contributes to the interpretation of TCM theories, further enhancing the accuracy and reliability of CMM research. For example, TCM theory holds that CMM has properties such as four Qi (Siqi in Chinese); five flavors (Wuwei in Chinese); ascending, descending, floating, and sinking (Shengjiang Fuchen in Chinese); meridian entry (Guijing in Chinese); and toxicity. These properties, essential in determining the efficacy and application of CMM, have traditionally been derived from ancient texts and are challenging to explain using modern scientific principles. However, informatics and AI have successfully been employed to develop predictive models for CMM properties [210,211,212]. These models play a crucial role in bridging the gap between traditional knowledge and contemporary scientific understanding, thus guiding the direction and logic of CMM research.

4.1.2. Improving the Efficiency and Reducing the Cost of CMM Research

CADD has been applied across various aspects of CMM research. While specific statistical data may be lacking, it is widely recognized that CADD effectively enhances the efficiency and reduces the costs associated with CMM research [9]. In fact, the application of CADD in WM research is more extensive and supported by more comprehensive data, which to some extent highlights the advantages of CADD in CMM research. For example, traditional drug development cycles are lengthy, typically spanning 10–15 years and costing USD 2–3 billion [213]. With the help of CADD, researchers can screen millions of molecules in just a few days and identify preclinical candidates within a year [214]. One study even estimated that AI could provide over 50 new therapies in 10 years, saving USD 50 billion in preclinical costs alone [215].
At present, traditional experimental methods still dominate CMM research, and the application of CADD is relatively limited compared to its widespread use in WM research. However, the inherent complexity of CMM, involving multiple ingredients, targets, and pathways, makes traditional experimental methods time-consuming and labor-intensive [216]. Therefore, CADD holds great promise in CMM research and is expected to evolve into a research methodology that is as significant as traditional experimental approaches.

4.1.3. Promoting the Modernization and Internationalization of CMM Research

Despite the growing global acceptance of CMM, it remains regrettable that no CMM product has been approved by the FDA. Although several have entered clinical trials in recent years, with claims of potential FDA approval, they have all failed [217]. In contrast, the CMM market in China is booming, with its scale approaching CNY 1 trillion. The China Center for Drug Evaluation processed 2569 CMM registration applications in 2023, marking a year-over-year increase of 75.24% [218,219]. This discrepancy arises from the significant differences between CMM and modern chemical medicines in terms of theory, source, formulation, manufacturing, and clinical evaluation. These factors hinder the modernization and internationalization of CMM.
CADD plays a dual role in this context: on the one hand, it helps to clarify the material basis and mechanisms of CMM in treating diseases; on the other hand, it encourages CMM research to align with the processes and standards of international drug development. The data obtained from CADD in CMM research can be compared with those from WM, facilitating the modernization and internationalization of CMM. Already, CADD has helped numerous WM products to gain FDA approval. As its application in CMM research continues to expand, it is likely that, in the future, CMM products will also gain FDA approval with the assistance of CADD.

4.2. Challenges of CADD in CMM Research

The advantages of CADD in CMM research are very attractive; however, the practical application of CADD in CMM research has also highlighted several challenges, many of which were initially raised in WM research (Figure 7). These challenges present significant obstacles to the advancement of CADD in CMM research [10,220,221].

4.2.1. Inadequate Computational Accuracy and Computational Resources

While CADD contributes to the accuracy and reliability of CMM research, particularly when experimental results are available for comparison, the general accuracy and reliability of CADD remains a subject of ongoing debate. In 1998, when Walter Kohn and John A. Pople were awarded the Nobel Prize in Chemistry for the development of DFT, the Royal Swedish Academy of Sciences stated, “chemistry is no longer a purely experimental science” [222]. Over the years, despite significant efforts by computational chemistry researchers to improve algorithms’ efficiency and accuracy, traditional experimental results continue to be regarded as the gold standard within the academic community. The most direct method of evaluating the accuracy of computational chemistry algorithms remains their consistency with experimental results. It is observed that an increasing number of researchers are adopting in silico approaches for a deeper understanding of experimental results. This has resulted in a growing body of high-quality academic literature that integrates computation with experimentation. Such a trend is expected to continue in academic research, thereby promoting the application of CADD in CMM research.
As a major country conducting CMM research, China, with its late development in the semiconductor field, heavily relies on imported computational resources, including central processing units (CPUs) and graphics processing units (GPUs). Currently, the global demand for computational resources is extremely high, fueled by the growing popularity of AI. Studies have shown that the high cost of computational resources often poses a significant challenge for researchers engaged in computational chemistry and AI [223]. In recent years, researchers have focused on minimizing computational resource usage while still achieving desired results. DeepSeek (https://www.deepseek.com/, accessed on 27 February 2025), a generative AI model developed in China, has optimized its algorithms to deliver results comparable to those of well-known generative AI models, such as ChatGPT (https://chatgpt.com/, accessed on 27 February 2025), while using significantly fewer computational resources [224].
On the other hand, the continued promise of computer-aided technologies for life sciences is evident. In 2024, the Nobel Prize in Chemistry was awarded to David Baker, Demis Hassabis, and John M. Jumper for their pioneering contributions to protein structure prediction, particularly their groundbreaking integration of AI and computational chemistry to advance protein structure research. This underscores the academic community’s confidence in the future potential of CADD. With the rapid advancement of AI, CADD has made significant strides in terms of computational accuracy. For instance, AlphaFold 3, the latest protein prediction tool, now predicts protein structures with remarkable precision [225]. Overall, although the challenge of inadequate computational accuracy and resources may complicate the widespread adoption of CADD in CMM research, it is likely that, in the near future, CADD will become as integral to CMM research as traditional experimental methods.

4.2.2. Difficulties in Data Collection and Data Quality Control

The collection of data is central to informatics, which is an essential component of CADD. In terms of the database quantity, the number of CMM databases is significantly smaller than that of chemical and biological databases, and this disparity extends to other data sources, such as the literature, books, and medical cases. As a result, the application of CADD in CMM research faces inherent disadvantages in terms of data collection. CMM is characterized by its multi-ingredient, multi-target, and multi-pathway nature, often used in complex formulae in clinical practice. This complexity is reflected in the data types, which include source materials, chemical constituents, properties, efficacy, and pharmacological activity. Moreover, many of the CMM data come from ancient texts spanning various historical periods and medical schools of thought, requiring considerable human and material resources to organize and extract the information. These factors make data collection a significant challenge in CMM research.
Due to these difficulties in data collection, ensuring data quality becomes equally challenging. Our team previously compared the ingredients of the same CMM recorded in different databases and cross-referenced them with experimental data, revealing significant discrepancies [31,137]. More concerning is the fact that some TCM databases contain nearly identical data without citing specific sources, often integrated from other databases. Aside from databases, other data sources also present challenges. For example, limitations in the ancient knowledge of medicinal substances, differences in medical schools, regional and temporal variations, and constraints in recording methods have led to discrepancies between ancient records and modern scientific understanding. In some cases, records from different ancient texts even contradict one another. While it is essential for researchers to maintain high standards in the collection of primary data, we must actively employ rapidly evolving informatics and AI methods to address the difficulties in data collection and improve data quality control.

4.2.3. Insufficient Adaptability and Interpretability of AI Models

Although AI has significantly improved the efficiency and accuracy of CADD, the adaptability and interpretability of AI models remain key challenges. The current AI algorithms used in CMM research heavily rely on large volumes of high-quality data for training. However, as previously discussed, data collection and quality control in CMM research are particularly challenging, which directly impacts the effectiveness and usability of model training. For instance, models often perform well only for a limited set of strictly selected ingredients and targets, making it difficult to generalize them to the complex mechanisms of CMM formulae.
AI models are often referred to as “black box” models, where researchers find it difficult to explain the biological mechanisms underlying specific prediction results [215]. In CMM-related CADD, this “black box” issue exacerbates the trust barrier that researchers face regarding a model’s predictions. Given the considerable gap between traditional CMM theory and modern scientific understanding, ensuring that a model’s output is both interpretable and verifiable in clinical applications remains a major challenge. Although some explainable AI techniques have been proposed, their application in CMM research is still in the exploratory phase and requires further development in the future [226].

5. Conclusions

CADD in CMM research primarily spans two key fields, computational chemistry and informatics, with the rise of AI offering new opportunities for both areas. This paper systematically reviews the foundational theories and research methods across these two domains, presenting numerous case studies directly tied to CMM research. Our analysis indicates that CADD holds significant promise in advancing CMM research, aligning well with national policies and the current trajectory of modern scientific development, particularly in the field of AI. While CADD, as a rapidly evolving technology, faces certain challenges, it is positioned to play an increasingly pivotal role in the future of CMM research.

Author Contributions

Conceptualization, B.C. and Y.Z.; data curation, S.L., H.X. and X.L.; writing—original draft preparation, B.C.; writing—review and editing, X.L. and Y.Z.; visualization, B.C.; supervision, B.C. and Y.Z.; project administration, B.C. and Y.Z.; funding acquisition, B.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (82304707), the Hubei Provincial Natural Science Foundation of China (2023AFB373), and the Doctoral Research Initiation Grant of Hubei University of Technology (XJ2022004001).

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

The following abbreviations are used in this manuscript:
ADMETabsorption, distribution, metabolism, excretion, and toxicity
AIartificial intelligence
AIDDartificial intelligence drug design
CADDcomputer-aided drug design
CAIsChinese materia medica active ingredients
CDcircular dichroism
CMMChinese materia medica
CMM-NPChinese materia medica network pharmacology
CNKIChina National Knowledge Infrastructure
CPUscentral processing units
DEGsdifferentially expressed genes
DFTdensity functional theory
DLdeep learning
ESPmolecular surface electrostatic potentials
FDAUS Food and Drug Administration
GOgene ontology
GPUsgraphics processing units
GWASgenome-wide association studies
HIVhuman immunodeficiency virus
HOMOhighest occupied molecular orbital
InChIInternational Chemical Identifier
IRinfrared
IRCintrinsic reaction coordinates
KRRkernel ridge regression
KEGGKyoto Encyclopedia of Genes and Genomes
LBDDligand-based drug design
LUMOlowest unoccupied molecular orbital
MDmolecular dynamics
MDSmolecular dynamics simulation
MLmachine learning
MMmolecular mechanics
MM/PB(GB)SAmolecular mechanics Poisson–Boltzmann/generalized Born surface area
MSmass spectrometry
NMRnuclear magnetic resonance
NPAnatural population analysis
NSFCNational Natural Science Foundation of China
NPsnatural products
PPIprotein–protein interaction
QCquantum chemistry
QCCquantum chemical calculation
QMquantum mechanics
QM/MMquantum mechanics/molecular mechanics
Q-markersquality markers
QSARquantitative structure–activity relationship
QSPRquantitative structure–property relationship
SBDDstructure-based drug design
SMILESsimplified molecular input line entry system
TCMtraditional Chinese medicine
UV–Visultraviolet–visible
WGCNAweighted gene co-expression network analysis
WMWestern medicine

References

  1. Li, C.; Sun, H.; Kong, R. Traditional Chinese Medicine and Self-Molding of National Image. Asia-Pac. Tradit. Med. 2024, 20, 6–9. [Google Scholar]
  2. Cyranoski, D. Why Chinese Medicine Is Heading for Clinics around the World. Nature 2018, 561, 448–450. [Google Scholar] [CrossRef]
  3. Bi, M.; Xu, H.; Han, L.; Sun, R. Stable Support from National Nature Science Foundation for Innovative Development of Basic Research in Traditional Chinese Medicine. Bull. Natl. Nat. Sci. Found. China 2024, 38, 378–382. [Google Scholar]
  4. Zhou, M.C.; Fei, Y.T.; Lai, X.Z.; Lan, J.; Liu, B.; Wang, Z.W.; Fang, H.; Liu, J.P.; Rong, H.G. Progress and Challenges in Integrated Traditional Chinese and Western Medicine in China from 2002 to 2021. Front. Pharmacol. 2024, 15, 1425940. [Google Scholar] [CrossRef]
  5. Xu, Z.; Dang, Y.; Chen, X.; Hai; Yao, W.; Kou, W.; Zhang, J.; Shi, J.; Dong, Y.; Li, J. Quercetin 7-Rhamnoside from Sorbaria Sorbifolia Exerts Anti-Hepatocellular Carcinoma Effect via DHRS13/Apoptotic Pathway. Phytomedicine 2024, 135, 156031. [Google Scholar] [CrossRef] [PubMed]
  6. Fu, S.; Song, X.; Tang, X.; Qian, X.; Du, Z.; Hu, Y.; Xu, X.; Zhang, M. Synergistic Effect of Constituent Drugs of Baibutang on Improving Yin-Deficiency Pulmonary Fibrosis in Rats. J. Ethnopharmacol. 2023, 306, 116050. [Google Scholar] [CrossRef] [PubMed]
  7. Humblet, C.; Marshall, G.R. Three-Dimensional Computer Modeling as an Aid to Drug Design. Drug Dev. Res. 1981, 1, 409–434. [Google Scholar] [CrossRef]
  8. Merz, K.M.; Ringe, D.; Reynolds, C.H. Drug Design: Structure- and Ligand-Based Approaches; Cambridge University Press: Cambridge, UK, 2010. [Google Scholar]
  9. Lin, Y.; Zhang, Y.; Wang, D.; Yang, B.; Shen, Y.Q. Computer Especially AI-Assisted Drug Virtual Screening and Design in Traditional Chinese Medicine. Phytomedicine 2022, 107, 154481. [Google Scholar] [CrossRef] [PubMed]
  10. Vemula, D.; Jayasurya, P.; Sushmitha, V.; Kumar, Y.N.; Bhandari, V. CADD, AI and ML in Drug Discovery: A Comprehensive Review. Eur. J. Pharm. Sci. 2023, 181, 106324. [Google Scholar] [CrossRef]
  11. Levine, I.N.; Busch, D.H.; Shull, H. Quantum Chemistry; Pearson Prentice Hall: Upper Saddle River, NJ, USA, 2009; Volume 6. [Google Scholar]
  12. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Petersson, G.A.; Nakatsuji, H.; et al. Gaussian 09, Revision C.01; Gaussian, Inc.: Wallingford, CT, USA, 2009. [Google Scholar]
  13. Neese, F.; Wennmohs, F.; Becker, U.; Riplinger, C. The ORCA Quantum Chemistry Program Package. J. Chem. Phys. 2020, 152, 224108. [Google Scholar] [CrossRef]
  14. Bannwarth, C.; Caldeweyher, E.; Ehlert, S.; Hansen, A.; Pracht, P.; Seibert, J.; Spicher, S.; Grimme, S. Extended Tight-Binding Quantum Chemistry Methods. Wiley Interdiscip. Rev. Comput. Mol. Sci. 2021, 11, e1493. [Google Scholar] [CrossRef]
  15. Zhang, K.T.; Huang, Z.P.; Xu, X.R.; Li, S.H.; Xu, Y.X.; Zhao, Q.; Zhang, X.M. Two New Diketopiperazines from the Cordyceps Fungus Samsoniella Sp. XY4. J. Antibiot. 2023, 76, 735–740. [Google Scholar] [CrossRef]
  16. Gan, L.S.; Zheng, Y.L.; Mo, J.X.; Liu, X.; Li, X.H.; Zhou, C.X. Sesquiterpene Lactones from the Root Tubers of Lindera Aggregata. J. Nat. Prod. 2009, 72, 1497–1501. [Google Scholar] [CrossRef]
  17. Chang, T.; Liang, J.; Wei, D.; Cui, H.L. Molecular and Crystalline Vibration Characteristics of Baicalin Investigated by Terahertz Spectroscopy and Density Functional Theory. IEEE Trans. Terahertz Sci. Technol. 2022, 12, 611–618. [Google Scholar] [CrossRef]
  18. Lu, S.; Gao, Z.; He, D.; Zhang, L.; Ke, G. Data-Driven Quantum Chemical Property Prediction Leveraging 3D Conformations with Uni-Mol+. Nat. Commun. 2024, 15, 7104. [Google Scholar] [CrossRef]
  19. Stephens, P.J.; McCann, D.M.; Devlin, F.J.; Smith, A.B. Determination of the Absolute Configurations of Natural Products via Density Functional Theory Calculations of Optical Rotation, Electronic Circular Dichroism, and Vibrational Circular Dichroism:  The Cytotoxic Sesquiterpene Natural Products Quadrone, Suberosenone, Suberosanone, and Suberosenol a Acetate. J. Nat. Prod. 2006, 69, 1055–1064. [Google Scholar] [PubMed]
  20. Yi, Y.; Adrjan, B.; Wlodarz, J.; Li, J.; Jackowski, K.; Roszak, S. NMR Measurements and DFT Studies of Nuclear Magnetic Shielding in Emodin and Chuanxiongzine Molecules. J. Mol. Struct. 2018, 1166, 304–310. [Google Scholar] [CrossRef]
  21. Gao, Y.; Hu, Z.; Wu, J.; Ning, Z.; Jian, J.; Zhao, T.; Liang, X.; Yang, X.; Yang, Z.; Zhao, Q.; et al. Size-Tunable Au@ag Nanoparticles for Colorimetric and SERS Dual-Mode Sensing of Palmatine in Traditional Chinese Medicine. J. Pharm. Biomed. Anal. 2019, 174, 123–133. [Google Scholar] [CrossRef]
  22. Zhang, G.; Chen, Y.; Liu, F.; Huang, J.; Li, P.; Wang, B.; Zhao, W.; Chen, M.; Xu, S.; Guan, F.; et al. Comprehensive Investigation of Structural Properties (X-Ray Diffraction, IR, Hirshfeld, MEP and FMOs) and in Silico Screening of Potential Biological Activity of Euphorbia Factor L1. J. Mol. Struct. 2021, 1246, 131237. [Google Scholar] [CrossRef]
  23. Gupta, D.; Ranjan, R.; Shukla, M. Molecular Interaction of Curcumin with Silver Nanocluster: A DFT Study. Vib. Spectrosc. 2023, 129, 103604. [Google Scholar] [CrossRef]
  24. Wang, S.; Lin, C.; Zhao, L.; Gong, X.; Zhang, M.; Zhang, H.; Hu, P. Identifying Isomers in Chinese Traditional Medicine via Density Functional Theory and Ion Fragmentation Simulation Software QCxMS. J. Chromatogr. A 2024, 1730, 465122. [Google Scholar] [CrossRef] [PubMed]
  25. Li, X.; Zeng, J.; Cai, R.; Li, C. New UHPLC-Q-Orbitrap MS/MS-Based Library-Comparison Method Simultaneously Distinguishes 22 Phytophenol Isomers from Desmodium Styracifolium. Microchem. J. 2023, 191, 108938. [Google Scholar] [CrossRef]
  26. Hu, G.; Qiu, M. Machine Learning-Assisted Structure Annotation of Natural Products Based on MS and NMR Data. Nat. Prod. Rep. 2023, 40, 1735–1753. [Google Scholar] [CrossRef]
  27. Tsai, Y.H.; Amichetti, M.; Zanardi, M.M.; Grimson, R.; Daranas, A.H.; Sarotti, A.M. ML-J-DP4: An Integrated Quantum Mechanics-Machine Learning Approach for Ultrafast NMR Structural Elucidation. Org. Lett. 2022, 24, 7487–7491. [Google Scholar] [CrossRef]
  28. Zeng, Y.; Qian, H.; Chen, X.; Li, Z.; Yu, S.; Xiao, X. Thermodynamic Estimate of pKa Values of the Carboxylic Acids in Aqueous Solution with the Density Functional Theory. Chin. J. Chem. 2010, 28, 727–733. [Google Scholar] [CrossRef]
  29. Guan, D.; Lui, R.; Matthews, S. LogP Prediction Performance with the SMD Solvation Model and the M06 Density Functional Family for SAMPL6 Blind Prediction Challenge Molecules. J. Comput. Aided Mol. Des. 2020, 34, 511–522. [Google Scholar] [CrossRef]
  30. Yang, Y.; Tang, W.; Liu, S.; Han, D.; Liu, Y.; Gong, J. Solubility of Benzoin in Three Binary Solvent Mixtures and Investigation of Intermolecular Interactions by Molecular Dynamic Simulation. J. Mol. Liq. 2017, 243, 472–483. [Google Scholar] [CrossRef]
  31. Chen, B.; Liu, S.; Li, X.; Li, C.; Cai, R.; Zeng, J.; Hu, Y.; Su, J.; Chen, S. Reconstruction of Quality Marker System for Ginkgo Folium Tablet Using UHPLC-Q-orbitrap MS, Quantum Chemical Calculation, Network Pharmacology, and Molecular Simulation. Phytochem. Anal. 2024, 35, 1–15. [Google Scholar] [CrossRef]
  32. Chen, B.; Su, J.; Hu, Y.; Liu, S.; Ouyang, X.; Cai, R.; Li, X. Antioxidant Mechanisms and Products of Four 4′,5,7-Trihydroxyflavonoids with Different Structural Types. RSC Med. Chem. 2022, 14, 173–182. [Google Scholar] [CrossRef]
  33. Bredas, J.L. Mind the Gap! Mater. Horiz. 2014, 1, 17–19. [Google Scholar] [CrossRef]
  34. Ma, S.; McGann, M.; Enyedy, I.J. The Influence of Calculated Physicochemical Properties of Compounds on Their ADMET Profiles. Bioorg. Med. Chem. Lett. 2021, 36, 127825. [Google Scholar] [CrossRef] [PubMed]
  35. He, Y.N.; Chen, L.M.; Liu, Y.; Ma, H.Y.; Hu, Q.; Cao, Z.X.; Han, L.; Xu, R.C.; Yang, M.; Tian, Y.; et al. New Understanding of Aconitine Hydrolysis Pathway: Isolation, Identification and Toxicity Evaluation Based on Intermediate Products. Arabian J. Chem. 2022, 15, 104255. [Google Scholar]
  36. Su, J.; Li, D.; Hu, Y.; You, X.; Guo, X.; Li, X.; Chen, B. A Novel C6-Sulfonated Celastrol Analog as a Tyrosinase and Melanin Inhibitor: Design, Synthesis, Biological Evaluation and Molecular Simulation. J. Mol. Struct. 2023, 1283, 135288. [Google Scholar] [CrossRef]
  37. Halabi Diaz, A.; Duque-Noreña, M.; Chamorro, E. Unveiling an Electronic LogP Analogue within the Conceptual Density Functional Theory Framework. Chem. Phys. 2024, 584, 112346. [Google Scholar] [CrossRef]
  38. Neugebauer, H.; Bohle, F.; Bursch, M.; Hansen, A.; Grimme, S. Benchmark Study of Electrochemical Redox Potentials Calculated with Semiempirical and DFT Methods. J. Phys. Chem. A 2020, 124, 7166–7176. [Google Scholar] [CrossRef]
  39. Rice, B.M.; Byrd, E.F.C. Evaluation of Electrostatic Descriptors for Predicting Crystalline Density. J. Comput. Chem. 2013, 34, 2146–2151. [Google Scholar] [CrossRef]
  40. Byrd, E.F.C.; Rice, B.M. Improved Prediction of Heats of Formation of Energetic Materials Using Quantum Mechanical Calculations. J. Phys. Chem. A 2006, 110, 1005–1013. [Google Scholar] [CrossRef]
  41. Murray, J.S.; Lane, P.; Brinck, T.; Paulsen, K.; Grice, M.E.; Politzer, P. Relationships of Critical Constants and Boiling Points to Computed Molecular Surface Properties. J. Phys. Chem. 1993, 97, 9369–9373. [Google Scholar] [CrossRef]
  42. Afzal, M.A.F.; Hachmann, J. Benchmarking DFT Approaches for the Calculation of Polarizability Inputs for Refractive Index Predictions in Organic Polymers. Phys. Chem. Chem. Phys. 2019, 21, 4452–4460. [Google Scholar] [CrossRef]
  43. Li, J.; Meng, J. Nuclear Magnetic Moments in Covariant Density Functional Theory. Front. Phys. 2018, 13, 132109. [Google Scholar] [CrossRef]
  44. Dennington, R.; Keith, T.A.; Millam, J.M. GaussView, Version 6; Semichem Inc.: Shawnee Mission, KS, USA, 2016. [Google Scholar]
  45. Lu, T. A Comprehensive Electron Wavefunction Analysis Toolbox for Chemists, Multiwfn. J. Chem. Phys. 2024, 161, 082503. [Google Scholar] [CrossRef] [PubMed]
  46. Liu, J.; Guo, H.; Zhou, J.; Wang, Y.; Yan, H.; Jin, R.; Tang, Y. Evodiamine and Rutaecarpine as Potential Anticancer Compounds: A Combined Computational Study. Int. J. Mol. Sci. 2022, 23, 11513. [Google Scholar] [CrossRef]
  47. Zhang, R.; Asikaer, A.; Chen, Q.; Wang, F.; Lan, J.; Liu, Y.; Hu, L.; Zhao, H.; Duan, H. Network Pharmacology and in Vitro Experimental Verification Unveil Glycyrrhizin from Glycyrrhiza Glabra Alleviates Acute Pancreatitis via Modulation of MAPK and STAT3 Signaling Pathways. BMC Complement. Med. Ther. 2024, 24, 58. [Google Scholar] [CrossRef]
  48. Lim, W.Z.; Cheng, P.G.; Abdulrahman, A.Y.; Teoh, T.C. The Identification of Active Compounds in Ganoderma Lucidum Var. Antler Extract Inhibiting Dengue Virus Serine Protease and Its Computational Studies. J. Biomol. Struct. Dyn. 2020, 38, 4273–4288. [Google Scholar] [CrossRef]
  49. Du, H.F.; Li, L.; Zhang, Y.H.; Wang, X.; Zhou, C.Y.; Zhu, H.J.; Pittman, C.U.; Shou, J.W.; Cao, F. The First Dimeric Indole-Diterpenoids from a Marine-Derived Penicillium Sp. Fungus and Their Potential for Anti-Obesity Drugs. Mar. Life Sci. Technol. 2025, 7, 120–131. [Google Scholar] [CrossRef]
  50. Zhao, W.; Cui, H.; Liu, K.; Yang, X.; Xing, S.; Li, W. Unveiling Anti-Diabetic Potential of Baicalin and Baicalein from Baikal Skullcap: LC–MS, In Silico, and In Vitro Studies. Int. J. Mol. Sci. 2024, 25, 3654. [Google Scholar] [CrossRef] [PubMed]
  51. Paukku, Y.; Rasulev, B.; Syrov, V.; Khushbaktova, Z.; Leszczynski, J. Structure-hepatoprotective Activity Relationship Study of Sesquiterpene Lactones: A QSAR Analysis. Int. J. Quantum Chem. 2009, 109, 17–27. [Google Scholar] [CrossRef]
  52. Qian, J.Z.; Wang, B.C.; Fan, Y.; Tan, J.; Yang, X. QSAR Study of Flavonoid-Metal Complexes and Their Anticancer Activities. J. Struct. Chem. 2015, 56, 338–345. [Google Scholar] [CrossRef]
  53. Wang, L.; Ding, J.; Pan, L.; Cao, D.; Jiang, H.; Ding, X. Quantum Chemical Descriptors in Quantitative Structure–Activity Relationship Models and Their Applications. Chemom. Intell. Lab. Syst. 2021, 217, 104384. [Google Scholar] [CrossRef]
  54. El Rhabori, S.; El Aissouq, A.; Chtita, S.; Khalil, F. QSAR, Molecular Docking and ADMET Studies of Quinoline, Isoquinoline and Quinazoline Derivatives against Plasmodium falciparum Malaria. Struct. Chem. 2023, 34, 585–603. [Google Scholar] [CrossRef]
  55. Hornig, M.; Klamt, A. COSMOfrag:  A Novel Tool for High-Throughput ADME Property Prediction and Similarity Screening Based on Quantum Chemistry. J. Chem. Inf. Model. 2005, 45, 1169–1177. [Google Scholar] [CrossRef] [PubMed]
  56. Silva-Júnior, E.F.; Aquino, T.M.; Araujo-Junior, J.X. Quantum Mechanical (QM) Calculations Applied to ADMET Drug Prediction: A Review. Curr. Drug Metab. 2017, 18, 511–526. [Google Scholar] [CrossRef]
  57. Lim, M.A.; Yang, S.; Mai, H.; Cheng, A.C. Exploring Deep Learning of Quantum Chemical Properties for Absorption, Distribution, Metabolism, and Excretion Predictions. J. Chem. Inf. Model. 2022, 62, 6336–6341. [Google Scholar] [CrossRef]
  58. Tropsha, A.; Isayev, O.; Varnek, A.; Schneider, G.; Cherkasov, A. Integrating QSAR Modelling and Deep Learning in Drug Discovery: The Emergence of Deep QSAR. Nat. Rev. Drug Discov. 2024, 23, 141–155. [Google Scholar] [CrossRef] [PubMed]
  59. Zhang, Q.; Liang, J.; Li, X.; Li, X.; Xia, B.; Shi, M.; Zeng, J.; Huang, H.; Yang, L.; He, J. Exploring Antithrombotic Mechanisms and Effective Constituents of Lagopsis Supina Using an Integrated Strategy Based on Network Pharmacology, Molecular Docking, Metabolomics, and Experimental Verification in Rats. J. Ethnopharmacol. 2025, 336, 118717. [Google Scholar] [CrossRef] [PubMed]
  60. Hehre, W.J. A Guide to Molecular Mechanics and Quantum Chemical Calculations; Wavefunction, Inc.: Irvine, CA, USA, 2003. [Google Scholar]
  61. Karplus, M.; Petsko, G.A. Molecular Dynamics Simulations in Biology. Nature 1990, 347, 631–639. [Google Scholar] [CrossRef]
  62. Hornak, V.; Abel, R.; Okur, A.; Strockbine, B.; Roitberg, A.; Simmerling, C. Comparison of Multiple Amber Force Fields and Development of Improved Protein Backbone Parameters. Proteins Struct. Funct. Bioinf. 2006, 65, 712–725. [Google Scholar] [CrossRef]
  63. Rudrapal, M.; Issahaku, A.R.; Agoni, C.; Bendale, A.R.; Nagar, A.; Soliman, M.E.S.; Lokwani, D. In Silico Screening of Phytopolyphenolics for the Identification of Bioactive Compounds as Novel Protease Inhibitors Effective against SARS-CoV-2. J. Biomol. Struct. Dyn. 2022, 40, 10437–10453. [Google Scholar] [CrossRef]
  64. Begum, S.; Shadrack, D.M.; Joseph, F.M.; Ndensendo, V.M.K. Molecular Dynamics Simulation of Bioactive Compounds of Withania Somnifera Leaf Extract as DNA Gyrase Inhibitor. J. Biomol. Struct. Dyn. 2022, 40, 9279–9286. [Google Scholar] [CrossRef]
  65. Ponder, J.W.; Wu, C.; Ren, P.; Pande, V.S.; Chodera, J.D.; Schnieders, M.J.; Haque, I.; Mobley, D.L.; Lambrecht, D.S.; DiStasio, R.A.; et al. Current Status of the AMOEBA Polarizable Force Field. J. Phys. Chem. B 2010, 114, 2549–2564. [Google Scholar] [CrossRef]
  66. Trenzado, J.L.; Benito, C.; Atilhan, M.; Aparicio, S. Hydrophobic Deep Eutectic Solvents Based on Cineole and Organic Acids. J. Mol. Liq. 2023, 377, 121322. [Google Scholar] [CrossRef]
  67. Vanommeslaeghe, K.; Raman, E.; MacKerell, A. Automation of the CHARMM General Force Field (CGenFF) II: Assignment of Bonded Parameters and Partial Atomic Charges. J. Chem. Inf. Model. 2012, 52, 3155–3168. [Google Scholar] [CrossRef] [PubMed]
  68. Karagiannis, T.C.; Ververis, K.; Liang, J.J.; Pitsillou, E.; Kagarakis, E.A.; Yi, D.T.; Xu, V.; Hung, A.; El-Osta, A. Investigation of the Anti-Inflammatory Properties of Bioactive Compounds from Olea Europaea: In Silico Evaluation of Cyclooxygenase Enzyme Inhibition and Pharmacokinetic Profiling. Molecules 2024, 29, 3502. [Google Scholar] [CrossRef] [PubMed]
  69. Zuhri, U.M.; Purwaningsih, E.H.; Fadilah, F.; Yuliana, N.D. Network Pharmacology Integrated Molecular Dynamics Reveals the Bioactive Compounds and Potential Targets of Tinospora crispa Linn. as Insulin Sensitizer. PLoS ONE 2022, 17, e0251837. [Google Scholar] [CrossRef]
  70. Vanommeslaeghe, K.; Hatcher, E.; Acharya, C.; Kundu, S.; Zhong, S.; Shim, J.; Darian, E.; Guvench, O.; Lopes, P.; Vorobyov, I.; et al. CHARMM General Force Field: A Force Field for Drug-like Molecules Compatible with the CHARMM All-atom Additive Biological Force Fields. J. Comput. Chem. 2010, 31, 671–690. [Google Scholar] [CrossRef] [PubMed]
  71. Subramani, N.K.; Venugopal, S. Molecular Docking and Dynamic Simulation Studies of Bioactive Compounds from Traditional Medicinal Compounds against Exfoliative Toxin B from Staphylococcus Aureus. J. Pharmacol. Pharmacother. 2024, 15, 316–326. [Google Scholar] [CrossRef]
  72. Wang, J.; Wolf, R.M.; Caldwell, J.W.; Kollman, P.A.; Case, D.A. Development and Testing of a General Amber Force Field. J. Comput. Chem. 2004, 25, 1157–1174. [Google Scholar] [CrossRef]
  73. Mirza, F.J.; Zahid, S.; Amber, S.; Jabeen, H.; Asim, N.; Ali Shah, S.A. Multitargeted Molecular Docking and Dynamic Simulation Studies of Bioactive Compounds from Rosmarinus Officinalis against Alzheimer’s Disease. Molecules 2022, 27, 7241. [Google Scholar] [CrossRef]
  74. Yepes-Pérez, A.F.; Herrera-Calderon, O.; Quintero-Saumeth, J. Uncaria Tomentosa (Cat’s Claw): A Promising Herbal Medicine against SARS-CoV-2/ACE-2 Junction and SARS-CoV-2 Spike Protein Based on Molecular Modeling. J. Biomol. Struct. Dyn. 2022, 40, 2227–2243. [Google Scholar] [CrossRef]
  75. Kirschner, K.N.; Yongye, A.B.; Tschampel, S.M.; González-Outeiriño, J.; Daniels, C.R.; Foley, B.L.; Woods, R.J. GLYCAM06: A Generalizable Biomolecular Force Field. Carbohydrates. J. Comput. Chem. 2008, 29, 622–655. [Google Scholar] [CrossRef]
  76. Jitapunkul, K.; Toochinda, P.; Lawtrakul, L. Molecular Dynamic Simulation Analysis on the Inclusion Complexation of Plumbagin with β-Cyclodextrin Derivatives in Aqueous Solution. Molecules 2021, 26, 6784. [Google Scholar] [CrossRef]
  77. Oostenbrink, C.; Villa, A.; Mark, A.E.; Van Gunsteren, W.F. A Biomolecular Force Field Based on the Free Enthalpy of Hydration and Solvation: The GROMOS Force-field Parameter Sets 53A5 and 53A6. J. Comput. Chem. 2004, 25, 1656–1676. [Google Scholar] [CrossRef] [PubMed]
  78. Mohamad Rosdi, M.N.; Mohd Arif, S.; Abu Bakar, M.H.; Razali, S.A.; Mohamed Zulkifli, R.; Ya’akob, H. Molecular Docking Studies of Bioactive Compounds from Annona Muricata Linn as Potential Inhibitors for Bcl-2, Bcl-w and Mcl-1 Antiapoptotic Proteins. Apoptosis 2018, 23, 27–40. [Google Scholar] [CrossRef]
  79. Gunasekaran, P.; Velmurugan, Y.; Arputharaj, D.S.; Savaridasson, J.K.; Hemamalini, M.; Venkatachalam, R. In Vitro Contraceptive Activities, Molecular Docking, Molecular Dynamics, MM-PBSA, Non-Covalent Interaction and DFT Studies of Bioactive Compounds from Aegle Marmelos. Linn., Leaves. Front. Chem. 2023, 11, 1096177. [Google Scholar] [CrossRef] [PubMed]
  80. Marrink, S.J.; Risselada, H.J.; Yefimov, S.; Tieleman, D.P.; De Vries, A.H. The MARTINI Force Field: Coarse Grained Model for Biomolecular Simulations. J. Phys. Chem. B 2007, 111, 7812–7824. [Google Scholar] [CrossRef]
  81. Laurella, L.C.; Moglioni, A.G.; Martini, M.F. Molecular Study of Endo and Phytocannabinoids on Lipid Membranes of Different Composition. Colloid. Surface. B 2023, 221, 113020. [Google Scholar] [CrossRef] [PubMed]
  82. Halgren, T.A. Merck Molecular Force Field. I. Basis, Form, Scope, Parameterization, and Performance of MMFF94. J. Comput. Chem. 1996, 17, 490–519. [Google Scholar] [CrossRef]
  83. Chung, H.M.; Hong, P.H.; Su, J.H.; Hwang, T.L.; Lu, M.C.; Fang, L.S.; Wu, Y.C.; Li, J.J.; Chen, J.J.; Wang, W.H. Bioactive Compounds from a Gorgonian Coral Echinomuricea Sp.(Plexauridae). Marine Drugs 2012, 10, 1169–1179. [Google Scholar] [CrossRef]
  84. Lolok, N.; Sumiwi, S.A.; Muhtadi, A.; Susilawati, Y.; Hendriani, R.; Ramadhan, D.S.F.; Levita, J.; Sahidin, I. Molecular Docking and Molecular Dynamics Studies of Bioactive Compounds Contained in Noni Fruit (Morinda Citrifolia L.) against Human Pancreatic α-Amylase. J. Biomol. Struct. Dyn. 2022, 40, 7091–7098. [Google Scholar] [CrossRef]
  85. Shaikh, S.; Ali, S.; Lim, J.H.; Chun, H.J.; Ahmad, K.; Ahmad, S.S.; Hwang, Y.C.; Han, K.S.; Kim, N.R.; Lee, E.J. Dipeptidyl Peptidase-4 Inhibitory Potentials of Glycyrrhiza Uralensis and Its Bioactive Compounds Licochalcone a and Licochalcone B: An in Silico and in Vitro Study. Front. Mol. Biosci. 2022, 9, 1024764. [Google Scholar] [CrossRef]
  86. Jorgensen, W.L.; Maxwell, D.S.; Tirado-Rives, J. Development and Testing of the OPLS All-Atom Force Field on Conformational Energetics and Properties of Organic Liquids. J. Am. Chem. Soc. 1996, 118, 11225–11236. [Google Scholar] [CrossRef]
  87. Patidar, K.; Deshmukh, A.; Bandaru, S.; Lakkaraju, C.; Girdhar, A.; Vr, G.; Banerjee, T.; Nayarisseri, A.; Singh, S.K. Virtual Screening Approaches in Identification of Bioactive Compounds Akin to Delphinidin as Potential HER2 Inhibitors for the Treatment of Breast Cancer. Asian Pac. J. Cancer Prev. 2016, 17, 2291–2295. [Google Scholar] [CrossRef] [PubMed]
  88. Murugesan, S.; Kottekad, S.; Crasta, I.; Sreevathsan, S.; Usharani, D.; Perumal, M.K.; Mudliar, S.N. Targeting COVID-19 (SARS-CoV-2) Main Protease through Active Phytocompounds of Ayurvedic Medicinal Plants–Emblica Officinalis (Amla), Phyllanthus Niruri Linn.(Bhumi Amla) and Tinospora Cordifolia (Giloy)–a Molecular Docking and Simulation Study. Comput. Biol. Med. 2021, 136, 104683. [Google Scholar] [CrossRef] [PubMed]
  89. Clark, M.; Cramer, R.D.; Van Opdenbosch, N. Validation of the General Purpose Tripos 5.2 Force Field. J. Comput. Chem. 1989, 10, 982–1012. [Google Scholar] [CrossRef]
  90. Aziz, M.; Ahmad, S.; Iqbal, M.N.; Khurshid, U.; Saleem, H.; Alamri, A.; Anwar, S.; Alamri, A.S.; Chohan, T.A. Phytochemical, Pharmacological, and in-Silico Molecular Docking Studies of Strobilanthes Glutinosus Nees: An Unexplored Source of Bioactive Compounds. S. Afr. J. Bot. 2022, 147, 618–627. [Google Scholar] [CrossRef]
  91. Morris, G.M.; Goodsell, D.S.; Halliday, R.S.; Huey, R.; Hart, W.E.; Belew, R.K.; Olson, A.J. Automated Docking Using a Lamarckian Genetic Algorithm and an Empirical Binding Free Energy Function. J. Comput. Chem. 1998, 19, 1639–1662. [Google Scholar] [CrossRef]
  92. Trott, O.; Olson, A.J. AutoDock Vina: Improving the Speed and Accuracy of Docking with a New Scoring Function, Efficient Optimization, and Multithreading. J. Comput. Chem. 2010, 31, 455–461. [Google Scholar] [CrossRef]
  93. Friesner, R.A.; Banks, J.L.; Murphy, R.B.; Halgren, T.A.; Klicic, J.J.; Mainz, D.T.; Repasky, M.P.; Knoll, E.H.; Shelley, M.; Perry, J.K.; et al. Glide: A New Approach for Rapid, Accurate Docking and Scoring. 1. Method and Assessment of Docking Accuracy. J. Med. Chem. 2004, 47, 1739–1749. [Google Scholar] [CrossRef]
  94. Vilar, S.; Cozza, G.; Moro, S. Medicinal Chemistry and the Molecular Operating Environment (MOE): Application of QSAR and Molecular Docking to Drug Discovery. Curr. Top. Med. Chem. 2008, 8, 1555–1572. [Google Scholar] [CrossRef]
  95. Ewing, T.J.; Makino, S.; Skillman, A.G.; Kuntz, I.D. DOCK 4.0: Search Strategies for Automated Molecular Docking of Flexible Molecule Databases. J. Comput. Aided Mol. Des. 2001, 15, 411–428. [Google Scholar] [CrossRef]
  96. Case, D.A.; Cheatham III, T.E.; Darden, T.; Gohlke, H.; Luo, R.; Merz Jr, K.M.; Onufriev, A.; Simmerling, C.; Wang, B.; Woods, R.J. The Amber Biomolecular Simulation Programs. J. Comput. Chem. 2005, 26, 1668–1688. [Google Scholar] [CrossRef] [PubMed]
  97. Van Der Spoel, D.; Lindahl, E.; Hess, B.; Groenhof, G.; Mark, A.E.; Berendsen, H.J.C. GROMACS: Fast, Flexible, and Free. J. Comput. Chem. 2005, 26, 1701–1718. [Google Scholar] [CrossRef]
  98. Christen, M.; Hünenberger, P.H.; Bakowies, D.; Baron, R.; Bürgi, R.; Geerke, D.P.; Heinz, T.N.; Kastenholz, M.A.; Kräutler, V.; Oostenbrink, C.; et al. The GROMOS Software for Biomolecular Simulation: GROMOS05. J. Comput. Chem. 2005, 26, 1719–1751. [Google Scholar] [CrossRef] [PubMed]
  99. Zheng, M.; Liu, X.; Xu, Y.; Li, H.; Luo, C.; Jiang, H. Computational Methods for Drug Design and Discovery: Focus on China. Trends Pharmacol. Sci. 2013, 34, 549–559. [Google Scholar] [CrossRef] [PubMed]
  100. Lu, F.; Wang, D.; Li, R.L.; He, L.Y.; Ai, L.; Wu, C.J. Current Strategies and Technologies for Finding Drug Targets of Active Components from Traditional Chinese Medicine. Front. Biosci. 2021, 26, 572–589. [Google Scholar] [CrossRef]
  101. Zhang, N.; Li, M.; Chen, X.; Wang, Y.; Wu, G.; Hu, G.; Wu, H. Solution Structure of BmKK2, a New Potassium Channel Blocker from the Venom of Chinese Scorpion Buthus Martensi Karsch. Proteins Struct. Funct. Bioinf. 2004, 55, 835–845. [Google Scholar] [CrossRef]
  102. Gao Weina; Li Yun; Zhang Rui; Gao Hui; Xu Weiren; Li Aixiu; Du Qishi; Zhang Xin; Wei Dongqing Screening of new HIV inhibitors based on the database of traditional Chinese medicine. Acta Pharm. Sin. 2006, 41, 241–246.
  103. Fan, J.; Fu, A.; Zhang, L. Progress in Molecular Docking. Quant. Biol. 2019, 7, 83–89. [Google Scholar] [CrossRef]
  104. Paggi, J.M.; Pandit, A.; Dror, R.O. The Art and Science of Molecular Docking. Annu. Rev. Biochem. 2024, 93, 389–410. [Google Scholar] [CrossRef]
  105. Zhou, G.; Rusnac, D.-V.; Park, H.; Canzani, D.; Nguyen, H.M.; Stewart, L.; Bush, M.F.; Nguyen, P.T.; Wulff, H.; Yarov-Yarovoy, V.; et al. An Artificial Intelligence Accelerated Virtual Screening Platform for Drug Discovery. Nat. Commun. 2024, 15, 7761. [Google Scholar] [CrossRef]
  106. Noor, F.; Junaid, M.; Almalki, A.H.; Almaghrabi, M.; Ghazanfar, S.; Tahir ul Qamar, M. Deep Learning Pipeline for Accelerating Virtual Screening in Drug Discovery. Sci. Rep. 2024, 14, 28321. [Google Scholar] [CrossRef] [PubMed]
  107. Mitchell, M.J.; McCammon, J.A. Free Energy Difference Calculations by Thermodynamic Integration: Difficulties in Obtaining a Precise Value. J. Comput. Chem. 1991, 12, 271–275. [Google Scholar] [CrossRef]
  108. Jorgensen, W.L.; Thomas, L.L. Perspective on Free-Energy Perturbation Calculations for Chemical Equilibria. J. Chem. Theory Comput. 2008, 4, 869–876. [Google Scholar] [CrossRef]
  109. Hansson, T.; Marelius, J.; Åqvist, J. Ligand Binding Affinity Prediction by Linear Interaction Energy Methods. J. Comput. Aided Mol. Des. 1998, 12, 27–35. [Google Scholar] [CrossRef] [PubMed]
  110. Genheden, S.; Ryde, U. The MM/PBSA and MM/GBSA Methods to Estimate Ligand-Binding Affinities. Expert Opin. Drug Discov. 2015, 10, 449–461. [Google Scholar] [CrossRef]
  111. Li, Y.; Zhang, Y.; Wu, X.; Gao, Y.; Guo, J.; Tian, Y.; Lin, Z.; Wang, X. Discovery of Natural 15-LOX Small Molecule Inhibitors from Chinese Herbal Medicine Using Virtual Screening, Biological Evaluation and Molecular Dynamics Studies. Bioorg. Chem. 2021, 115, 105197. [Google Scholar] [CrossRef]
  112. Wang, J.; Gao, S.; Shi, J.; Cao, H.; Ye, T.; Yue, M.; Ye, F.; Fu, Y. Virtual Screening Based on Pharmacophore Model for Developing Novel HPPD Inhibitors. Pestic. Biochem. Physiol. 2022, 184, 105109. [Google Scholar] [CrossRef]
  113. Zhang, Y.; Huang, M.; Wang, Q.; Zhang, X.; Peng, J.; Zhang, Y.; Wu, Q.; Duan, J.; Mao, X.; Tang, Z.; et al. Insights into the Penetration of PhACs in TCM during Ultrafiltration: Effects of Fouling Mechanisms and Intermolecular Interactions. Sep. Purif. Technol. 2022, 295, 121205. [Google Scholar] [CrossRef]
  114. Yin, Q.; Shi, X.; Ding, H.; Dai, X.; Wan, G.; Qiao, Y. Interactions of Borneol with DPPC Phospholipid Membranes: A Molecular Dynamics Simulation Study. Int. J. Mol. Sci. 2014, 15, 20365–20381. [Google Scholar] [CrossRef]
  115. Dai, X.; Wang, R.; Wu, Z.; Guo, S.; Yang, C.; Ma, L.; Chen, L.; Shi, X.; Qiao, Y. Permeation-enhancing Effects and Mechanisms of Borneol and Menthol on Ligustrazine: A Multiscale Study Using in Vitro and Coarse-grained Molecular Dynamics Simulation Methods. Chem. Biol. Drug Des. 2018, 92, 1830–1837. [Google Scholar] [CrossRef]
  116. Huang, W.; Yang, Y.; Wen, W.; Luo, Y.; Wu, J.; Xiang, L.; Hu, Y.; Xu, S.; Chen, S.; Wang, P. Vanillin Enhances the Passive Transport Rate and Absorption of Drugs with Moderate Oral Bioavailability in Vitro and in Vivo by Affecting the Membrane Structure. Food Funct. 2020, 11, 700–710. [Google Scholar] [CrossRef]
  117. Zhou, K.; Liu, Z.; Fan, R.; Zhao, M.; Luo, L.; Wang, Y.; Jiang, Y.; Lu, Z.; Tang, J.; Luo, A.; et al. A New Methodology of Understanding the Mechanism of High Shear Wet Granulation Based on Experiment and Molecular Dynamics Simulation. Int. J. Pharm. 2023, 638, 122923. [Google Scholar] [CrossRef] [PubMed]
  118. Shen, W.; Pan, S.; Li, J.; Ding, X.; Li, J.; Xu, J.; Qiu, Y.; Xu, W. Formulation Design and Evaluation of Ginsenoside Compound K Nanostructured Lipid Carriers Based on Molecular Dynamics Simulations. J. Pharm. Innov. 2024, 19, 1–15. [Google Scholar] [CrossRef]
  119. Dai, X.; Shi, X.; Ding, H.; Yin, Q.; Qiao, Y. Dissipative Particle Dynamics Simulation of Ginsenoside Ro Vesicular Solubilization Systems. J. Comput. Theor. Nanosci. 2014, 11, 2046–2054. [Google Scholar] [CrossRef]
  120. Lei, C.; Wen, J.; Sun, Y.; Ren, M.; Qiao, R.; Li, C. Self-Assembled Herbal Hydrogel for Rectal Administration Therapy in Ulcerative Colitis. Chem. Eng. J. 2025, 503, 158477. [Google Scholar] [CrossRef]
  121. Zhang, X.; Dong, X.; Zhang, R.; Hao, D.; Zhang, J.; Shen, Y.; Chai, X.; Wang, H.; Wang, Y.; Wang, Y. Unraveling the Mechanism of the Supramolecular Self-Assembly during the in Vivo Metabolism of Geniposide from Chinese Medicine. Mater. Des. 2023, 225, 111546. [Google Scholar] [CrossRef]
  122. Guo, X.; Luo, W.; Wu, L.; Zhang, L.; Chen, Y.; Li, T.; Li, H.; Zhang, W.; Liu, Y.; Zheng, J.; et al. Natural Products from Herbal Medicine Self-assemble into Advanced Bioactive Materials. Adv. Sci. 2024, 11, 2403388. [Google Scholar] [CrossRef]
  123. Prašnikar, E.; Ljubič, M.; Perdih, A.; Borišek, J. Machine Learning Heralding a New Development Phase in Molecular Dynamics Simulations. Artif. Intell. Rev. 2024, 57, 102. [Google Scholar] [CrossRef]
  124. Warshel, A.; Levitt, M. Theoretical Studies of Enzymic Reactions: Dielectric, Electrostatic and Steric Stabilization of the Carbonium Ion in the Reaction of Lysozyme. J. Mol. Biol. 1976, 103, 227–249. [Google Scholar] [CrossRef]
  125. Kühne, T.D.; Iannuzzi, M.; Del Ben, M.; Rybkin, V.V.; Seewald, P.; Stein, F.; Laino, T.; Khaliullin, R.Z.; Schütt, O.; Schiffmann, F.; et al. CP2K: An Electronic Structure and Molecular Dynamics Software Package—Quickstep: Efficient and Accurate Electronic Structure Calculations. J. Chem. Phys. 2020, 152, 194103. [Google Scholar] [CrossRef]
  126. Xue, Q.; Su, X.; Yu, W.; Liu, J.; Hou, K.; Wang, C. Efficient Production of Neohesperidin Enabled by Protein Engineering of Rhamnosyltransferase Cm1,2RhaT. ACS Sustain. Chem. Eng. 2024, 12, 1960–1972. [Google Scholar] [CrossRef]
  127. Lin, H.; Zhang, Y.; Pezeshki, S.; Duster, A.W.; Wang, B.; Wu, X.P.; Zheng, S.W.; Gagliardi, L.; Truhlar, D.G. QMMM 2023: A Program for Combined Quantum Mechanical and Molecular Mechanical Modeling and Simulations. Comput. Phys. Commun. 2024, 295, 108987. [Google Scholar] [CrossRef]
  128. Yi, Y.; Zhang, M.; Xue, H.; Yu, R.; Bao, Y.O.; Kuang, Y.; Chai, Y.; Ma, W.; Wang, J.; Shi, X.; et al. Schaftoside Inhibits 3CLpro and PLpro of SARS-CoV-2 Virus and Regulates Immune Response and Inflammation of Host Cells for the Treatment of COVID-19. Acta Pharm. Sin. B 2022, 12, 4154–4164. [Google Scholar] [CrossRef]
  129. Bodun, D.S.; Omoboyowa, D.A.; Omotuyi, O.I.; Olugbogi, E.A.; Balogun, T.A.; Ezeh, C.J.; Omirin, E.S. QSAR-Based Virtual Screening of Traditional Chinese Medicine for the Identification of Mitotic Kinesin Eg5 Inhibitors. Comput. Biol. Chem. 2023, 104, 107865. [Google Scholar] [CrossRef] [PubMed]
  130. Lou, T.; Li, A.; Xu, H.; Pan, J.; Xing, B.; Wu, R.; Dickschat, J.S.; Yang, D.; Ma, M. Structural Insights into Three Sesquiterpene Synthases for the Biosynthesis of Tricyclic Sesquiterpenes and Chemical Space Expansion by Structure-Based Mutagenesis. J. Am. Chem. Soc. 2023, 145, 8474–8485. [Google Scholar] [CrossRef] [PubMed]
  131. Luo, Y.; Ma, X.; Qiu, Y.; Lu, Y.; Shen, S.; Li, Y.; Gao, H.; Chen, K.; Zhou, J.; Hu, T.; et al. Structural and Catalytic Insight into the Unique Pentacyclic Triterpene Synthase TwOSC. Angew. Chem. Int. Ed. 2023, 62, e202313429. [Google Scholar] [CrossRef]
  132. Liu, X.; Liu, Y.; Xu, X.; Huang, W.; Yan, Y.; Wang, Y.; Tian, W.; Mo, T.; Cui, X.; Li, J.; et al. Molecular Characterization and Structure Basis of a Malonyltransferase with Both Substrate Promiscuity and Catalytic Regiospecificity from Cistanche Tubulosa. Acta Pharm. Sin. B 2024, 14, 2333–2348. [Google Scholar] [CrossRef]
  133. Gu, S.; Pei, J. Chinese Herbal Medicine Meets Biological Networks of Complex Diseases: A Computational Perspective. Evid.-Based Compl. Alt. 2017, 2017, 7198645. [Google Scholar] [CrossRef]
  134. Matsuo, T.; Tsugawa, H.; Miyagawa, H.; Fukusaki, E. Integrated Strategy for Unknown EI-MS Identification Using Quality Control Calibration Curve, Multivariate Analysis, EI-MS Spectral Database, and Retention Index Prediction. Anal. Chem. 2017, 89, 6766–6773. [Google Scholar] [CrossRef]
  135. López-López, E.; Bajorath, J.; Medina-Franco, J.L. Informatics for Chemistry, Biology, and Biomedical Sciences. J. Chem. Inf. Model. 2021, 61, 26–35. [Google Scholar] [CrossRef]
  136. Fan, M.; Jin, C.; Li, D.; Deng, Y.; Yao, L.; Chen, Y.; Ma, Y.L.; Wang, T. Multi-Level Advances in Databases Related to Systems Pharmacology in Traditional Chinese Medicine: A 60-Year Review. Front. Pharmacol. 2023, 14, 1289901. [Google Scholar] [CrossRef] [PubMed]
  137. Chen, B.; Liu, S.; Li, X.; Cai, R.; Li, C.; Hu, Y.; Su, J.; Lei, T. Database-aided Ultrahigh-performance Liquid Chromatography Q-exactive-orbitrap Tandem Mass Spectrometry Putatively Identifies 16 Unexpected Compounds and Three Anticounterfeiting Pharmacopoeia Quality Markers for Perillae Fructus. Rapid Commun. Mass Spectrom. 2024, 38, e9762. [Google Scholar] [CrossRef] [PubMed]
  138. Liu, Z.; Guo, F.; Wang, Y.; Li, C.; Zhang, X.; Li, H.; Diao, L.; Gu, J.; Wang, W.; Li, D.; et al. BATMAN-TCM: A Bioinformatics Analysis Tool for Molecular mechANism of Traditional Chinese Medicine. Sci. Rep. 2016, 6, 21146. [Google Scholar] [CrossRef]
  139. Tao, W.; Li, B.; Gao, S.; Bai, Y.; Shar, P.A.; Zhang, W.; Guo, Z.; Sun, K.; Fu, Y.; Huang, C.; et al. CancerHSP: Anticancer Herbs Database of Systems Pharmacology. Sci. Rep. 2015, 5, 11481. [Google Scholar] [CrossRef]
  140. Huang, J.; Zheng, Y.; Wu, W.; Xie, T.; Yao, H.; Pang, X.; Sun, F.; Ouyang, L.; Wang, J. CEMTDD: The Database for Elucidating the Relationships among Herbs, Compounds, Targets and Related Diseases for Chinese Ethnic Minority Traditional Drugs. Oncotarget 2015, 6, 17675–17684. [Google Scholar] [CrossRef]
  141. Hou, D.; Lin, H.; Feng, Y.; Zhou, K.; Li, X.; Yang, Y.; Wang, S.; Yang, X.; Wang, J.; Zhao, H.; et al. CMAUP Database Update 2024: Extended Functional and Association Information of Useful Plants for Biomedical Research. Nucleic Acids Res. 2024, 52, D1508–D1518. [Google Scholar] [CrossRef] [PubMed]
  142. Sun, C.; Huang, J.; Tang, R.; Li, M.; Yuan, H.; Wang, Y.; Wei, J.M.; Liu, J. CPMCP: A Database of Chinese Patent Medicine and Compound Prescription. Database 2022, 2022, baac073. [Google Scholar] [CrossRef]
  143. Xu, H.Y.; Zhang, Y.Q.; Liu, Z.M.; Chen, T.; Lv, C.Y.; Tang, S.H.; Zhang, X.B.; Zhang, W.; Li, Z.Y.; Zhou, R.R.; et al. ETCM: An Encyclopaedia of Traditional Chinese Medicine. Nucleic Acids Res. 2019, 47, D976–D982. [Google Scholar] [CrossRef]
  144. Bultum, L.E.; Woyessa, A.M.; Lee, D. ETM-DB: Integrated Ethiopian Traditional Herbal Medicine and Phytochemicals Database. BMC Complement. Altern. Med. 2019, 19, 212. [Google Scholar] [CrossRef]
  145. Fang, S.; Dong, L.; Liu, L.; Guo, J.; Zhao, L.; Zhang, J.; Bu, D.; Liu, X.; Huo, P.; Cao, W.; et al. HERB: A High-Throughput Experiment- and Reference-Guided Database of Traditional Chinese Medicine. Nucleic Acids Res. 2021, 49, D1197–D1206. [Google Scholar] [CrossRef]
  146. Ye, H.; Ye, L.; Kang, H.; Zhang, D.; Tao, L.; Tang, K.; Liu, X.; Zhu, R.; Liu, Q.; Chen, Y.Z.; et al. HIT: Linking Herbal Active Ingredients to Targets. Nucleic Acids Res. 2011, 39, D1055–D1059. [Google Scholar] [CrossRef] [PubMed]
  147. Ye, Y.; Liang, D.; Yi, J.; Jin, S.; Zeng, Z. IGTCM: An Integrative Genome Database of Traditional Chinese Medicine Plants. Plant Genome 2023, 16, e20317. [Google Scholar] [CrossRef] [PubMed]
  148. Tian, S.; Zhang, J.; Yuan, S.; Wang, Q.; Lv, C.; Wang, J.; Fang, J.; Fu, L.; Yang, J.; Zu, X.; et al. Exploring Pharmacological Active Ingredients of Traditional Chinese Medicine by Pharmacotranscriptomic Map in ITCM. Briefings Bioinf. 2023, 24, bbad027. [Google Scholar] [CrossRef]
  149. Li, X.; Ren, J.; Zhang, W.; Zhang, Z.; Yu, J.; Wu, J.; Sun, H.; Zhou, S.; Yan, K.; Yan, X.; et al. LTM-TCM: A Comprehensive Database for the Linking of Traditional Chinese Medicine with Modern Medicine at Molecular and Phenotypic Levels. Pharmacol. Res. 2022, 178, 106185. [Google Scholar] [CrossRef]
  150. Chen, Q.; Springer, L.; Gohlke, B.O.; Goede, A.; Dunkel, M.; Abel, R.; Gallo, K.; Preissner, S.; Eckert, A.; Seshadri, L.; et al. SuperTCM: A Biocultural Database Combining Biological Pathways and Historical Linguistic Data of Chinese Materia Medica for Drug Development. Biomed. Pharmacother. 2021, 144, 112315. [Google Scholar] [CrossRef]
  151. Wu, Y.; Zhang, F.; Yang, K.; Fang, S.; Bu, D.; Li, H.; Sun, L.; Hu, H.; Gao, K.; Wang, W.; et al. SymMap: An Integrative Database of Traditional Chinese Medicine Enhanced by Symptom Mapping. Nucleic Acids Res. 2019, 47, D1110–D1117. [Google Scholar] [CrossRef] [PubMed]
  152. Chen, C.Y.C. TCM Database@taiwan: The World’s Largest Traditional Chinese Medicine Database for Drug Screening In Silico. PLoS ONE 2011, 6, e15939. [Google Scholar] [CrossRef]
  153. Lv, Q.; Chen, G.; He, H.; Yang, Z.; Zhao, L.; Zhang, K.; Chen, C.Y.C. TCMBank-the Largest TCM Database Provides Deep Learning-Based Chinese-Western Medicine Exclusion Prediction. Signal Transduct. Tar. 2023, 8, 127. [Google Scholar] [CrossRef]
  154. Chen, X.; Zhou, H.; Liu, Y.B.; Wang, J.F.; Li, H.; Ung, C.Y.; Han, L.Y.; Cao, Z.W.; Chen, Y.Z. Database of Traditional Chinese Medicine and Its Application to Studies of Mechanism and to Prescription Validation. Br. J. Pharmacol. 2006, 149, 1092–1103. [Google Scholar] [CrossRef]
  155. Xue, R.; Fang, Z.; Zhang, M.; Yi, Z.; Wen, C.; Shi, T. TCMID: Traditional Chinese Medicine Integrative Database for Herb Molecular Mechanism Analysis. Nucleic Acids Res. 2012, 41, D1089–D1095. [Google Scholar] [CrossRef]
  156. Liu, Z.; Cai, C.; Du, J.; Liu, B.; Cui, L.; Fan, X.; Wu, Q.; Fang, J.; Xie, L. TCMIO: A Comprehensive Database of Traditional Chinese Medicine on Immuno-Oncology. Front. Pharmacol. 2020, 11, 439. [Google Scholar] [CrossRef]
  157. Ren, Z.; Ren, Y.; Li, Z.; Xu, H. TCMM: A Unified Database for Traditional Chinese Medicine Modernization and Therapeutic Innovations. Comput. Struct. Biotec. 2024, 23, 1619–1630. [Google Scholar] [CrossRef]
  158. Zhang, R.; Yu, S.; 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]
  159. Zhang, L.X.; Dong, J.; Wei, H.; Shi, S.H.; Lu, A.P.; Deng, G.M.; Cao, D.S. TCMSID: A Simplified Integrated Database for Drug Discovery from Traditional Chinese Medicine. J. Cheminf. 2022, 14, 89. [Google Scholar] [CrossRef] [PubMed]
  160. 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. Cheminf. 2014, 6, 13. [Google Scholar] [CrossRef]
  161. Huang, L.; Wang, Q.; Duan, Q.; Shi, W.; Li, D.; Chen, W.; Wang, X.; Wang, H.; Chen, M.; Kuang, H.; et al. TCMSSD: A Comprehensive Database Focused on Syndrome Standardization. Phytomedicine 2024, 128, 155486. [Google Scholar] [CrossRef]
  162. Yang, P.; Lang, J.; Li, H.; Lu, J.; Lin, H.; Tian, G.; Bai, H.; Yang, J.; Ning, K. TCM-suite: A Comprehensive and Holistic Platform for Traditional Chinese Medicine Component Identification and Network Pharmacology Analysis. iMeta 2022, 1, e47. [Google Scholar] [CrossRef]
  163. Kim, S.K.; Lee, M.K.; Jang, H.; Lee, J.J.; Lee, S.; Jang, Y.; Jang, H.; Kim, A. TM-MC 2.0: An Enhanced Chemical Database of Medicinal Materials in Northeast Asian Traditional Medicine. BMC Complement. Med. Ther. 2024, 24, 40. [Google Scholar] [CrossRef] [PubMed]
  164. Li, B.; Ma, C.; Zhao, X.; Hu, Z.; Du, T.; Xu, X.; Wang, Z.; Lin, J. YaTCM: Yet Another Traditional Chinese Medicine Database for Drug Discovery. Comput. Struct. Biotechnol. J. 2018, 16, 600–610. [Google Scholar] [CrossRef]
  165. Sun, Y.; Tao, Q.; Cao, Y.; Yang, T.; Zhang, L.; Luo, Y.; Wang, L. Kaempferol Has Potential Anti-Coronavirus Disease 2019 (COVID-19) Targets Based on Bioinformatics Analyses and Pharmacological Effects on Endotoxin-Induced Cytokine Storm. Phytother. Res. 2023, 37, 2290–2304. [Google Scholar] [CrossRef]
  166. Kim, J.; Yoo, M.; Shin, J.; Kim, H.; Kang, J.; Tan, A.C. Systems Pharmacology-Based Approach of Connecting Disease Genes in Genome-Wide Association Studies with Traditional Chinese Medicine. Int. J. Genom. 2018, 2018, 7697356. [Google Scholar] [CrossRef] [PubMed]
  167. Dai, Y.; Qiang, W.; Gui, Y.; Tan, X.; Pei, T.; Lin, K.; Cai, S.; Sun, L.; Ning, G.; Wang, J.; et al. A Large-Scale Transcriptional Study Reveals Inhibition of COVID-19 Related Cytokine Storm by Traditional Chinese Medicines. Sci. Bull. 2021, 66, 884–888. [Google Scholar] [CrossRef]
  168. Zhong, Y.; Lee, K.; Deng, Y.; Ma, Y.; Chen, Y.; Li, X.; Wei, C.; Yang, S.; Wang, T.; Wong, N.J.; et al. Arctigenin Attenuates Diabetic Kidney Disease through the Activation of PP2A in Podocytes. Nat. Commun. 2019, 10, 4523. [Google Scholar] [CrossRef] [PubMed]
  169. Zhou, W.; Wu, J.; Zhang, J.; Liu, X.; Guo, S.; Jia, S.; Zhang, X.; Zhu, Y.; Wang, M. Integrated Bioinformatics Analysis to Decipher Molecular Mechanism of Compound Kushen Injection for Esophageal Cancer by Combining WGCNA with Network Pharmacology. Sci. Rep. 2020, 10, 12745. [Google Scholar] [CrossRef] [PubMed]
  170. Xu, X.; Yu, Y.; Yang, L.; Wang, B.; Fan, Y.; Ruan, B.; Zhang, X.; Dai, H.; Mei, W.; Jie, W.; et al. Integrated Analysis of Dendrobium Nobile Extract Dendrobin a against Pancreatic Ductal Adenocarcinoma Based on Network Pharmacology, Bioinformatics, and Validation Experiments. Front. Pharmacol. 2023, 14, 1079539. [Google Scholar] [CrossRef]
  171. Shannon, P.; Markiel, A.; Ozier, O.; Baliga, N.S.; Wang, J.T.; Ramage, D.; Amin, N.; Schwikowski, B.; Ideker, T. Cytoscape: A Software Environment for Integrated Models of Biomolecular Interaction Networks. Genome Res. 2003, 13, 2498–2504. [Google Scholar] [CrossRef]
  172. Zhang, Q.; Yu, H.; Qi, J.; Tang, D.; Chen, X.; Wan, J.; Li, P.; Hu, H.; Wang, Y.; Hu, Y. Natural Formulas and the Nature of Formulas: Exploring Potential Therapeutic Targets Based on Traditional Chinese Herbal Formulas. PLoS ONE 2017, 12, e0171628. [Google Scholar] [CrossRef]
  173. Guo, X.; Zhao, X.; Lu, X.; Zhao, L.; Zeng, Q.; Chen, F.; Zhang, Z.; Xu, M.; Feng, S.; Fan, T.; et al. A Deep Learning-Driven Discovery of Berberine Derivatives as Novel Antibacterial against Multidrug-Resistant Helicobacter Pylori. Signal Transduct. Tar. 2024, 9, 1–18. [Google Scholar] [CrossRef]
  174. Xu, Z.; Rasteh, A.M.; Dong, A.; Wang, P.; Liu, H. Identification of Molecular Targets of Hypericum Perforatum in Blood for Major Depressive Disorder: A Machine-Learning Pharmacological Study. Chin. Med. 2024, 19, 141. [Google Scholar] [CrossRef]
  175. Zhou, G.; Lin, Z.; Miao, Q.; Lin, L.; Wang, S.; Lu, K.; Zhang, Y.; Chu, Q.; Kong, W.; Wu, K.; et al. Mechanisms of QingRe HuoXue Formula in Atherosclerosis Treatment: An Integrated Approach Using Bioinformatics, Machine Learning, and Experimental Validation. Int. Immunopharmacol. 2024, 141, 112890. [Google Scholar] [CrossRef]
  176. Jin, M.; Ren, W.; Zhang, W.; Liu, L.; Yin, Z.; Li, D. Exploring the Underlying Mechanism of Shenyankangfu Tablet in the Treatment of Glomerulonephritis through Network Pharmacology, Machine Learning, Molecular Docking, and Experimental Validation. Drug Des. Dev. Ther. 2021, 15, 4585–4601. [Google Scholar] [CrossRef] [PubMed]
  177. Li, S. Framework and Practice of Network-Based Studies for Chinese Herbal Formula. J. Chin. Integr. Med. 2007, 5, 489–493. [Google Scholar] [CrossRef]
  178. Hopkins, A.L. Network Pharmacology. Nat. Biotechnol. 2007, 25, 1110–1111. [Google Scholar] [CrossRef]
  179. Zhao, L.; Zhang, H.; Li, N.; Chen, J.; Xu, H.; Wang, Y.; Liang, Q. Network Pharmacology, a Promising Approach to Reveal the Pharmacology Mechanism of Chinese Medicine Formula. J. Ethnopharmacol. 2023, 309, 116306. [Google Scholar] [CrossRef] [PubMed]
  180. Li, X.; Liu, Z.; Liao, J.; Chen, Q.; Lu, X.; Fan, X. Network Pharmacology Approaches for Research of Traditional Chinese Medicines. Chin. J. Nat. Med. 2023, 21, 323–332. [Google Scholar] [CrossRef]
  181. Peng, F.; Hong, W.; Wang, Y.; Peng, Y.; Fang, Z. Mechanism of Herb Pair Containing Astragali Radix and Spatholobi Caulis in the Treatment of Myelosuppression Based on Network Pharmacology and Experimental Investigation. J. Ethnopharmacol. 2024, 319, 117178. [Google Scholar] [CrossRef] [PubMed]
  182. Duan, Z.; Wang, Y.; Lu, Z.; Tian, L.; Xia, Z.Q.; Wang, K.; Chen, T.; Wang, R.; Feng, Z.; Shi, G.; et al. Wumei Wan Attenuates Angiogenesis and Inflammation by Modulating RAGE Signaling Pathway in IBD: Network Pharmacology Analysis and Experimental Evidence. Phytomedicine 2023, 111, 154658. [Google Scholar] [CrossRef]
  183. Shang, L.; Wang, Y.; Li, J.; Zhou, F.; Xiao, K.; Liu, Y.; Zhang, M.; Wang, S.; Yang, S. Mechanism of Sijunzi Decoction in the Treatment of Colorectal Cancer Based on Network Pharmacology and Experimental Validation. J. Ethnopharmacol. 2023, 302, 115876. [Google Scholar] [CrossRef]
  184. Zulkifli, M.H.; Abdullah, Z.L.; Mohamed Yusof, N.I.S.; Mohd Fauzi, F. In Silico Toxicity Studies of Traditional Chinese Herbal Medicine: A Mini Review. Curr. Opin. Struc. Biol. 2023, 80, 102588. [Google Scholar] [CrossRef]
  185. Wang, M.; Shi, Y.; Yao, L.; Li, Q.; Wang, Y.; Fu, D. Potential Molecular Mechanisms and Drugs for Aconitine-Induced Cardiotoxicity in Zebrafish through RNA Sequencing and Bioinformatics Analysis. Med. Sci. Monit. 2020, 26, e924092. [Google Scholar] [CrossRef]
  186. Liao, Y.; Ding, Y.; Yu, L.; Xiang, C.; Yang, M. Exploring the Mechanism of Alisma Orientale for the Treatment of Pregnancy Induced Hypertension and Potential Hepato-Nephrotoxicity by Using Network Pharmacology, Network Toxicology, Molecular Docking and Molecular Dynamics Simulation. Front. Pharmacol. 2022, 13, 1027112. [Google Scholar] [CrossRef] [PubMed]
  187. Jiang, H.Y.; Gao, H.Y.; Li, J.; Zhou, T.Y.; Wang, S.T.; Yang, J.B.; Hao, R.R.; Pang, F.; Wei, F.; Liu, Z.G.; et al. Integrated Spatially Resolved Metabolomics and Network Toxicology to Investigate the Hepatotoxicity Mechanisms of Component D of Polygonum Multiflorum Thunb. J. Ethnopharmacol. 2022, 298, 115630. [Google Scholar] [CrossRef]
  188. Dai, J.; Liu, J.; Zhang, M.; Yu, Y.; Wang, J. Network Toxicology and Molecular Docking Analyses on Strychnine Indicate CHRM1 Is a Potential Neurotoxic Target. BMC Complement. Med. Ther. 2022, 22, 273. [Google Scholar] [CrossRef]
  189. Ge, J.C.; Qian, Q.; Gao, Y.H.; Zhang, Y.F.; Li, Y.X.; Wang, X.; Fu, Y.; Ma, Y.M.; Wang, Q. Toxic Effects of Tripterygium Glycoside Tablets on the Reproductive System of Male Rats by Metabolomics, Cytotoxicity, and Molecular Docking. Phytomedicine 2023, 114, 154813. [Google Scholar] [CrossRef] [PubMed]
  190. Zhang, S.; Li, H.; Li, X. Omics and bioinformatics studies of the brain toxicity of Sophorae Tonkinensis Radix et Rhizoma in mice. J. Shenyang Pharm. Univ. 2023, 40, 343–349. [Google Scholar]
  191. Chen, Z.; Zhao, M.; You, L.; Zheng, R.; Jiang, Y.; Zhang, X.; Qiu, R.; Sun, Y.; Pan, H.; He, T.; et al. Developing an Artificial Intelligence Method for Screening Hepatotoxic Compounds in Traditional Chinese Medicine and Western Medicine Combination. Chin. Med. 2022, 17, 58. [Google Scholar] [CrossRef]
  192. Li, X.; Wang, L.; Bu, R.; Wang, Y.; Zhang, F.; Chu, Y.; Li, Y.; Cai, T. Model-guided research strategies for safe use of traditional Chinese medicine: Quantitative toxicology of traditional Chinese medicine. Chin. Tradit. Herb. Drugs 2023, 54, 359–366. [Google Scholar]
  193. Shen, P.; Sun, D.; Zhou, W.; Gao, Y. Research progress in toxicity prediction of traditional Chinese medicines. Chin. J. Pharmacovigil. 2023, 20, 473–479. [Google Scholar]
  194. Han, J.; Pang, X.; Liao, B.; Yao, H.; Song, J.; Chen, S. An Authenticity Survey of Herbal Medicines from Markets in China Using DNA Barcoding. Sci. Rep. 2016, 6, 18723. [Google Scholar] [CrossRef]
  195. Zhu, S.; Liu, Q.; Qiu, S.; Dai, J.; Gao, X. DNA Barcoding: An Efficient Technology to Authenticate Plant Species of Traditional Chinese Medicine and Recent Advances. Chin. Med. 2022, 17, 112. [Google Scholar] [CrossRef]
  196. Su, J.; Wang, Y.; Bai, M.; Peng, T.; Li, H.; Xu, H.J.; Guo, G.; Bai, H.; Rong, N.; Sahu, S.K.; et al. Soil Conditions and the Plant Microbiome Boost the Accumulation of Monoterpenes in the Fruit of Citrus Reticulata ‘Chachi. ’ Microbiome 2023, 11, 61. [Google Scholar] [CrossRef] [PubMed]
  197. Zhao, L.; Shi, M.; Zhang, Q.; Qin, L.; Sun, Y. Research Progress on Quality Characteristics and Formation Mechanism of Genuine Medicinal Materials. Chin. Tradit. Herb. Drugs 2022, 53, 6931–6947. [Google Scholar]
  198. Liu, C.; Liu, L.; Guo, D. Quality Marker of TCMs: Concept and Applications. Phytomedicine 2018, 44, 85–86. [Google Scholar] [CrossRef] [PubMed]
  199. Li, Y.; Fan, J.; Cheng, X.; Jin, H.; Wang, Y.; Wei, F.; An, F.; Ma, S. New Revolution for Quality Control of TCM in Industry 4.0: Focus on Artificial Intelligence and Bioinformatics. Trends Anal. Chem. 2024, 181, 118023. [Google Scholar] [CrossRef]
  200. Ma, L.; Xu, E.; Bai, M.; Miao, M. Analysis of Characteristics of Traditional Chinese Medicine External Treatment for Dyspepsia Based on Data Mining. AIP Conf. Proc. 2020, 2208, 020010. [Google Scholar]
  201. Wu, Y.; Bai, L.; Miao, M. Analysis of Characteristics of Traditional Chinese Medicines for Treating Dysmenorrhea Based on Data Mining. DEStech Trans. Soc. Sci. Edu. Hum. Sci. 2019, 1, 28154. [Google Scholar] [CrossRef]
  202. Dai, L.; Lu, A.; Zhong, L.; Zheng, G.; Bian, Z. Chinese Herbal Medicine for Hyperlipidaemia: A Review Based on Data Mining from 1990 to 2016. Curr. Vasc. Pharmacol. 2017, 15, 520–531. [Google Scholar] [CrossRef]
  203. Tang, S.; Shen, D.; Lu, P.; Yang, H. Advances in researches made via traditional Chinese medicine inheritance support system. China J. Tradit. Chin. Med. Pharm. 2015, 30, 329–331. [Google Scholar]
  204. Wang, W.; Wang, L.; Jiang, S.S.; Ding, X.; Wei, H.J.; Liang, T.X.; Zhao, X.D.; Jiang, L.D. Study on the medication pattern of traditional Chinese medicine against new coronavirus pneumonia based on the cloud platform of ancient and modern medical cases. J. Beijing Univ. Tradit. Chin. Med. 2023, 39, 800–805. [Google Scholar]
  205. Wang, X.; Li, H.Y.; Kang, L.; Liu, J.; Xing, Y.H.; Yang, C.; Yang, L.; Li, X.Y.; Lei, L. Construction and Application of TCM Miner. Chin. J. Libr. Inf. Sci. Tradit. Chin. Med. 2021, 45, 1–6. [Google Scholar]
  206. Liu, B.; Liu, W.; Tang, B.; Meng, T.; Yao, J.B.; Hu, W. Design and Implementation of Intelligent Chinese Medicine Data Analysis System. Comput. Knowl. Technol. 2022, 18, 51–53. [Google Scholar]
  207. Guo, Y.; Wang, T.; Chen, W.; Kaptchuk, T.; Li, X.; Gao, X.; Yao, J.; Tang, X.; Xu, Z. Acceptability of Traditional Chinese Medicine in Chinese People Based on 10-Year’s Real World Study with Mutiple Big Data Mining. Front. Public Health 2022, 9, 811730. [Google Scholar] [CrossRef]
  208. Zhou, Q.; Liu, J.; Xin, L.; Fang, Y.; Hu, Y.; Qi, Y.; He, M.; Fang, D.; Chen, X.; Cong, C. Association between Traditional Chinese Medicine and Osteoarthritis Outcome: A 5-Year Matched Cohort Study. Heliyon 2024, 10, e26289. [Google Scholar] [CrossRef]
  209. Lam, C.; Peng, L.; Yang, L.; Chou, H.; Li, C.; Zuo, Z.; Koon, H.; Cheung, Y. Examining Patterns of Traditional Chinese Medicine Use in Pediatric Oncology: A Systematic Review, Meta-Analysis and Data-Mining Study. J. Integr. Med. 2022, 20, 402–415. [Google Scholar] [CrossRef] [PubMed]
  210. Wang, Z. A Study of Chinese Herbal Properties Based on Machine Learning. In Proceedings of the 2015 10th International Conference on Information, Communications and Signal Processing (ICICS), Singapore, 2–4 December 2015; IEEE: Piscataway, NJ, USA, 2015; pp. 1–5. [Google Scholar]
  211. Huang, S.H.; Tung, C.W.; Fülöp, F.; Li, J.H. Developing a QSAR Model for Hepatotoxicity Screening of the Active Compounds in Traditional Chinese Medicines. Food Chem. Toxicol. 2015, 78, 71–77. [Google Scholar] [CrossRef] [PubMed]
  212. Wang, Y.; Jafari, M.; Tang, Y.; Tang, J. Predicting Meridian in Chinese Traditional Medicine Using Machine Learning Approaches. PLoS Comput. Biol. 2019, 15, e1007249. [Google Scholar] [CrossRef]
  213. Ekins, S.; Puhl, A.C.; Zorn, K.M.; Lane, T.R.; Russo, D.P.; Klein, J.J.; Hickey, A.J.; Clark, A.M. Exploiting Machine Learning for End-to-End Drug Discovery and Development. Nat. Mater. 2019, 18, 435–441. [Google Scholar] [CrossRef] [PubMed]
  214. Chan, H.C.S.; Shan, H.; Dahoun, T.; Vogel, H.; Yuan, S. Advancing Drug Discovery via Artificial Intelligence. Trends Pharmacol. Sci. 2019, 40, 592–604. [Google Scholar] [CrossRef]
  215. Gholap, A.D.; Uddin, M.J.; Faiyazuddin, M.; Omri, A.; Gowri, S.; Khalid, M. Advances in Artificial Intelligence in Drug Delivery and Development: A Comprehensive Review. Comput. Biol. Med. 2024, 178, 108702. [Google Scholar] [CrossRef]
  216. Li, D.; Hu, J.; Zhang, L.; Li, L.; Yin, Q.; Shi, J.; Guo, H.; Zhang, Y.; Zhuang, P. Deep Learning and Machine Intelligence: New Computational Modeling Techniques for Discovery of the Combination Rules and Pharmacodynamic Characteristics of Traditional Chinese Medicine. Eur. J. Pharmacol. 2022, 933, 175260. [Google Scholar] [CrossRef]
  217. Writing Group of Recommendations of Expert Panel from Chinese Geriatrics Society on the Clinical Use of Compound Danshen Dripping Pills. Recommendations on the Clinical Use of Compound Danshen Dripping Pills. Chin. Med. J. 2017, 130, 972–978. [Google Scholar] [CrossRef] [PubMed]
  218. Foresight Industry Research Institute. Report of Market Prospective and Investment Strategy Planning on Traditional Chinese Medicine Industry (2024-2029); Foresight Industry Research Institute: Beijing, China, 2024. [Google Scholar]
  219. Yao Zhi Zixun. China Pharmaceutical R&D Blue Book (2024); Yao Zhi Zixun: Chongqing, China, 2024. [Google Scholar]
  220. Chen, W.; Liu, X.; Zhang, S.; Chen, S. Artificial Intelligence for Drug Discovery: Resources, Methods, and Applications. Mol. Ther. Nucl. Acids 2023, 31, 691–702. [Google Scholar] [CrossRef] [PubMed]
  221. Hirlekar, B.U.; Nuthi, A.; Singh, K.D.; Murty, U.S.; Dixit, V.A. An Overview of Compound Properties, Multiparameter Optimization, and Computational Drug Design Methods for PARP-1 Inhibitor Drugs. Eur. J. Med. Chem. 2023, 252, 115300. [Google Scholar] [CrossRef] [PubMed]
  222. Yang, J.; Li, W.H.; Wang, D. Machine Learning: The Trends of Developing High-Efficiency Single-Atom Materials. Chem. Catal. 2021, 1, 24–26. [Google Scholar] [CrossRef]
  223. Khandelwal, A.; Yun, T.; Nayak, N.V.; Merullo, J.; Bach, S.H.; Sun, C.; Pavlick, E. $100K or 100 Days: Trade-Offs When Pre-Training with Academic Resources. arXiv 2024, arXiv:2410.23261. [Google Scholar]
  224. Guo, D.; Zhu, Q.; Yang, D.; Xie, Z.; Dong, K.; Zhang, W.; Chen, G.; Bi, X.; Wu, Y.; Li, Y.K.; et al. DeepSeek-Coder: When the Large Language Model Meets Programming—The Rise of Code Intelligence. arXiv 2024, arXiv:2401.14196. [Google Scholar]
  225. Abramson, J.; Adler, J.; Dunger, J.; Evans, R.; Green, T.; Pritzel, A.; Ronneberger, O.; Willmore, L.; Ballard, A.J.; Bambrick, J.; et al. Accurate Structure Prediction of Biomolecular Interactions with AlphaFold 3. Nature 2024, 630, 493–500. [Google Scholar] [CrossRef]
  226. Dwivedi, R.; Dave, D.; Naik, H.; Singhal, S.; Omer, R.; Patel, P.; Qian, B.; Wen, Z.; Shah, T.; Morgan, G.; et al. Explainable AI (XAI): Core Ideas, Techniques, and Solutions. ACM Comput. Surv. 2023, 55, 1–33. [Google Scholar] [CrossRef]
Figure 1. A typical Chinese materia medica (CMM) research workflow and its possible results or purpose. CAIs: CMM active ingredients. TCM: traditional Chinese medicine.
Figure 1. A typical Chinese materia medica (CMM) research workflow and its possible results or purpose. CAIs: CMM active ingredients. TCM: traditional Chinese medicine.
Pharmaceutics 17 00315 g001
Figure 2. The workflow and application of quantum chemical calculation (QCC) in Chinese materia medica research. QSPR/QSAR: quantitative structure–property relationships and quantitative structure–activity relationships.
Figure 2. The workflow and application of quantum chemical calculation (QCC) in Chinese materia medica research. QSPR/QSAR: quantitative structure–property relationships and quantitative structure–activity relationships.
Pharmaceutics 17 00315 g002
Figure 3. The molecular dynamics simulation (MDS) process (13) for a typical ligand–receptor complex. Other systems to which MDS is applied in Chinese materia medica (CMM) research are also shown in (4). CAIs: CMM active ingredients.
Figure 3. The molecular dynamics simulation (MDS) process (13) for a typical ligand–receptor complex. Other systems to which MDS is applied in Chinese materia medica (CMM) research are also shown in (4). CAIs: CMM active ingredients.
Pharmaceutics 17 00315 g003
Figure 4. Representation of the quantum mechanics/molecular mechanics method for a ligand–receptor system.
Figure 4. Representation of the quantum mechanics/molecular mechanics method for a ligand–receptor system.
Pharmaceutics 17 00315 g004
Figure 5. Applications of cheminformatics and bioinformatics in Chinese materia medica (CMM) research. TCM: traditional Chinese medicine; CAIs: CMM active ingredients; Q-marker: quality marker.
Figure 5. Applications of cheminformatics and bioinformatics in Chinese materia medica (CMM) research. TCM: traditional Chinese medicine; CAIs: CMM active ingredients; Q-marker: quality marker.
Pharmaceutics 17 00315 g005
Figure 6. The data mining workflow in Chinese materia medica (CMM) research. TCM: traditional Chinese medicine.
Figure 6. The data mining workflow in Chinese materia medica (CMM) research. TCM: traditional Chinese medicine.
Pharmaceutics 17 00315 g006
Figure 7. Advantages and challenges of computer-aided drug design (CADD) in Chinese materia medica (CMM) research. The “↑” means promoting effect and the “↓” means reducing effect.
Figure 7. Advantages and challenges of computer-aided drug design (CADD) in Chinese materia medica (CMM) research. The “↑” means promoting effect and the “↓” means reducing effect.
Pharmaceutics 17 00315 g007
Table 1. Commonly used molecular force fields in computer-aided drug design related to Chinese materia medica research.
Table 1. Commonly used molecular force fields in computer-aided drug design related to Chinese materia medica research.
Force Field NameDescriptionApplicable SystemsCMM Research Cases
AMBER [62]A well-known and widely used force field for various systems. It contains many versions, such as AMBER84, AMBER86, AMBER94, AMBER96, AMBER98, AMBER99, AMBER03, AMBER03UA, AMBER99SB, AMBER99SB-ILDN, AMBER14SB, and AMBER19SB.Proteins, nucleic acids, and some organic small molecules.[63,64]
AMOEBA [65]It introduces polarization effects to more accurately describe intermolecular interactions. It addresses the limitations of the traditional fixed-charge force field and proposes a more complex polarization model to improve the accuracy of the description of molecular properties.Biological macromolecules and organic small molecules in solution environments.[66]
CGenFF [67]A fully CHARMM-compatible force field dedicated to the simulation of organic small molecules.Organic small molecules.[68,69]
CHARMM [70]It was originally dedicated to the CHARMM program. After being updated with various versions, such as CHARMM16, CHARMM19, CHARMM22, CHARMM27, and CHARMM36, it is now supported by many programs.Proteins, nucleic acids, phospholipids, and sugars.[71]
GAFF [72]It is fully compatible with the AMBER force field and can describe a variety of organic small molecules. It is a simple force field with better structural description accuracy than some complex force fields.Organic small molecules.[73,74]
GLYCAM [75]It is fully compatible with the AMBER force field and can be used in the AMBER program to research glycoproteins. It includes various versions, such as GLYCAM93, GLYCAM2000, GLYCAM06, and GLYCAM06-LP.Proteins and sugars.[76]
GROMOS [77]A force field with a simple energy functional and extensive applications. It contains many versions, most of which are supported only by the GROMOS program and the Gromacs program.Condensed-phase simulation of proteins, nucleic acids, sugars, phospholipids, and organic small molecules.[78,79]
MARTINI [80]It improves the computational efficiency by simplifying the representation of atoms, combining multiple atoms into a single “coarse-grained” particle.Large-scale biophysical systems such as membranes, biopolymers, and complex fluids.[81]
MM [82]A high-precision force field developed by the Merk Group for the simulation of organic molecules. It is suitable for conformational searches and unsuitable for condensed-phase simulations. It also includes various versions, such as MM1, MM2, MM3, MM4, MM+, and MM2X.Organic small molecules.[83]
MMFF94 [82]An improved version of the MM series force fields for calculations of organic molecules and condensed phases.Organic small molecules.[84,85]
OPLS [86]A force field that initially specialized in condensed-phase simulations. Its versions include OPLS-UA, OPLS-AA, and OPLS-AA/M. Starting with OPLS 2.0, the force field is exclusive to Schrödinger, Inc. and has been developed into various simulation systems.Proteins, sugars, and organic small molecules.[87,88]
Tripos [89]The force field parameters are carefully optimized to provide a high-precision description based on QCC and experimental data.Organic small molecules and proteins.[90]
Note: Force fields unrelated to computer-aided drug design are not listed, although they may be commonly used.
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

Chen, B.; Liu, S.; Xia, H.; Li, X.; Zhang, Y. Computer-Aided Drug Design in Research on Chinese Materia Medica: Methods, Applications, Advantages, and Challenges. Pharmaceutics 2025, 17, 315. https://doi.org/10.3390/pharmaceutics17030315

AMA Style

Chen B, Liu S, Xia H, Li X, Zhang Y. Computer-Aided Drug Design in Research on Chinese Materia Medica: Methods, Applications, Advantages, and Challenges. Pharmaceutics. 2025; 17(3):315. https://doi.org/10.3390/pharmaceutics17030315

Chicago/Turabian Style

Chen, Ban, Shuangshuang Liu, Huiyin Xia, Xican Li, and Yingqing Zhang. 2025. "Computer-Aided Drug Design in Research on Chinese Materia Medica: Methods, Applications, Advantages, and Challenges" Pharmaceutics 17, no. 3: 315. https://doi.org/10.3390/pharmaceutics17030315

APA Style

Chen, B., Liu, S., Xia, H., Li, X., & Zhang, Y. (2025). Computer-Aided Drug Design in Research on Chinese Materia Medica: Methods, Applications, Advantages, and Challenges. Pharmaceutics, 17(3), 315. https://doi.org/10.3390/pharmaceutics17030315

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