Hostname: page-component-669899f699-7tmb6 Total loading time: 0 Render date: 2025-04-26T00:19:08.778Z Has data issue: false hasContentIssue false

Computational methods for binding site prediction on macromolecules

Published online by Cambridge University Press:  12 March 2025

Igor Kozlovskii
Affiliation:
Constructor Knowledge Labs, Bremen, Germany School of Science, Constructor University Bremen gGmbH, Bremen, Germany Tetra D AG, Schaffhausen, Switzerland
Petr Popov*
Affiliation:
Constructor Knowledge Labs, Bremen, Germany School of Science, Constructor University Bremen gGmbH, Bremen, Germany Tetra D AG, Schaffhausen, Switzerland
*
Corresponding author: Petr Popov; Email: [email protected]
Rights & Permissions [Opens in a new window]

Abstract

Binding sites are key components of biomolecular structures, such as proteins and RNAs, serving as hubs for interactions with other molecules. Identification of the binding sites in macromolecules is essential for structure-based molecular and drug design. However, experimental methods for binding site identification are resource-intensive and time-consuming. In contrast, computational methods enable large-scale binding site identification, structure flexibility analysis, as well as assessment of intermolecular interactions within the binding sites. In this review, we describe recent advances in binding site identification using machine learning methods; we classify the approaches based on the encoding of the macromolecule information about its sequence, structure, template knowledge, geometry, and energetic characteristics. Importantly, we categorize the methods based on the type of the interacting molecule, namely, small molecules, peptides, and ions. Finally, we describe perspectives, limitations, and challenges of the state-of-the-art methods with an emphasis on deep learning-based approaches. These computational approaches aim to advance drug discovery by expanding the druggable genome through the identification of novel binding sites in pharmacological targets and facilitating structure-based hit identification and lead optimization.

Type
Review
Creative Commons
Creative Common License - CCCreative Common License - BYCreative Common License - NCCreative Common License - ND
This is an Open Access article, distributed under the terms of the Creative Commons Attribution-NonCommercial-NoDerivatives licence (http://creativecommons.org/licenses/by-nc-nd/4.0), which permits non-commercial re-use, distribution, and reproduction in any medium, provided that no alterations are made and the original article is properly cited. The written permission of Cambridge University Press must be obtained prior to any commercial use and/or adaptation of the article.
Copyright
© The Author(s), 2025. Published by Cambridge University Press

Introduction

Proteins are essential for many cellular functions, including enzymatic activity, structural support, transport, and cell signaling (Alberts, Reference Alberts2017). Structurally, they are large macromolecules composed of long chains of amino acids, which fold into unique three-dimensional shapes specific to each protein (Rodwell et al., Reference Rodwell, Bender and Botham2018). Their functional roles are driven by local intermolecular interactions within specific regions called binding sites. Binding sites play a crucial role in drug discovery. They serve as ‘hot spots’ on pharmacological targets where designed drug-like molecules bind. Identifying novel binding sites expands the ‘druggable genome’, offering new strategies for therapeutic development and drug discovery (Hopkins and Groom, Reference Hopkins and Groom2002). Drug-like molecules typically target either the orthosteric binding site, where proteins interact with their natural ligands, or distinct allosteric binding sites, which have garnered special interest. Allosteric sites show higher sequence variability between protein subtypes, enabling the design of more selective drug-like molecules compared to those targeting orthosteric binding sites (Changeux, Reference Changeux2013; Wagner et al., Reference Wagner2016; Lu et al., Reference Lu2018). Furthermore, the binding sites can be formed by several protein molecules at their interaction interface (Ferré et al., Reference Ferré2014; Wang et al., Reference Wang2018a), opening another opportunity for proximity-induced drug discovery (Békés et al., Reference Békés, Langley and Crews2022; Dewey et al., Reference Dewey2023; Liu and Ciulli, Reference Liu and Ciulli2023; Tan et al., Reference Tan2024). While proteins are the most common pharmacological targets, nucleic acids, particularly RNAs, are gaining increasing interest in structure-based drug design (Chen et al., Reference Chen2024; Tong et al., Reference Tong, Childs-Disney and Disney2024). RNA plays a vital role in gene regulation and information transfer, making it an appealing target for drug development (Warner et al., Reference Warner, Hajdin and Weeks2018). Like proteins, RNA molecules are highly structured and contain binding sites that can be modulated by small molecules (Yu et al., Reference Yu, Choi and Tu2020). Both proteins and nucleic acids are flexible macromolecules, adopting multiple conformations throughout their life cycle. Accordingly, binding sites are dynamic properties, influenced by the conformational changes of macromolecules (Laskowski et al., Reference Laskowski, Gerick and Thornton2009; Changeux and Christopoulos, Reference Changeux and Christopoulos2016). A single structure of a macromolecule represents only a single point of the complete conformational space. Therefore, it is possible to overlook binding sites in the static structures (Di Pietro et al., Reference Di Pietro2017; Sun et al., Reference Sun2020). A remarkable progress has been made in developing experimental methods for identifying binding sites, including fragment screening, site-directed tethering (Hardy and Wells, Reference Hardy and Wells2004; Ludlow et al., Reference Ludlow2015), antibody-based techniques (Lawson, Reference Lawson2012), small molecule microarrays (Doyle et al., Reference Doyle2016), hydrogen-deuterium exchange (Chalmers et al., Reference Chalmers2006), and site-directed mutagenesis (Gelis et al., Reference Gelis2012). However, experimental methods are often resource-intensive and may yield negative results. In contrast, computational approaches enable large-scale identification of binding sites, exploration of macromolecular flexibility, and the ability to assess how well chemical compounds fit into these sites.

While there are several articles describing binding site prediction methods (Laurie and Jackson, Reference Laurie and Jackson2006; Henrich et al., Reference Henrich2010; Leis et al., Reference Leis, Schneider and Zacharias2010; Chen et al., Reference Chen2011; Roche et al., Reference Roche, Brackenridge and McGuffin2015; Simões et al., Reference Simões2017; Zhao et al., Reference Zhao, Cao and Zhang2020; Liao et al., Reference Liao2022; Utgés and Barton, Reference Utgés and Barton2024), we found several gaps persisting in the literature. Specifically, most of the works focus only on protein–small molecule interactions, neglecting other important binding site interaction types, such as protein–peptide, nucleic acid–small molecules, or protein–ion. Furthermore, while deep learning-based approaches have gained popularity, there has been a limited discussion on their limitations, applicability, and interpretability compared to traditional methods. In this study, we provide a comprehensive review of computational methods for the prediction of binding sites. We eliminate existing gaps in the literature with a unified overview of computational techniques across diverse binding site interaction types. The computational methods are classified based on the types of binding sites they predict: (i) protein–small molecule binding sites (Section Protein–small molecule binding sites); (ii) protein–peptide binding sites (Section Protein–peptide binding sites); (iii) nucleic acid–small molecule binding sites (Section Nucleic acid–small molecule binding sites); and (iv) protein–ion binding sites (Section Protein–ion binding site prediction). For each type of binding site, the corresponding methods are further divided into categories based on the macromolecule input representation (sequence or structure) and algorithm type (template-based, geometric, energetic, machine learning-based, and deep learning-based). If available, we also list the benchmarks and the performance metrics of different methods for each category of binding sites. Finally, we conclude the review with a discussion of current challenges and future perspectives in the field.

Protein–small molecule binding sites

In this section, we describe computational methods for the prediction of small molecule binding sites on proteins. Small molecules are usually defined as molecules with a mass $ \le $ 500 Da (Benet et al., Reference Benet2016) designed to interact with biological targets to modulate their functions (Southey and Brunavs, Reference Southey and Brunavs2023). A binding site usually has specific geometry and physicochemical properties, making the corresponding protein region distinguishable from the rest of the protein surface. Thus, methods for predicting binding sites in proteins aim to identify such regions based on the input information (e.g., sequence, structure, or other type of information). This section is organized as follows: first, we divided methods into two large groups based on the representation of input protein as sequence or structure. After that, we further split all structure-based methods into five categories based on the type of algorithm they use: template-based, geometric, energetic, machine learning-based, and deep learning-based methods. Table 1 provides a list of computational method for prediction of small molecule binding sites on proteins along with type of approach they use.

Table 1. List of methods for prediction of protein–small molecule binding sites

Sequence-based

Sequence-based methods utilize sequence or sequence-driven information about proteins. In this problem setup, the model takes the protein sequence as input and outputs a score for each position in this sequence, indicating whether the amino acid residue at this position interacts with the ligand or not. Figure 1 provides a schematic representation of the overall pipeline of sequence-based methods. Some methods search for similar sequences in a template database of sequences with known binders and map binding information from them (Kauffman and Karypis, Reference Kauffman and Karypis2009; López et al., Reference López, Valencia and Tress2007; Yang et al., Reference Yang, Roy and Zhang2013). The idea is supported by the study suggesting, that in most cases the function of the unknown protein can be identified from its sequence or structure by homology (Yao et al., Reference Yao2003). However, for the correct work of template-based methods, there should be proteins with significant sequence identity to a query sequence in a database (Devos and Valencia, Reference Devos and Valencia2000; Wilson et al., Reference Wilson, Kreychman and Gerstein2000), as it was shown that proteins with <35–40% identity may not share the same biochemical function (Todd et al., Reference Todd, Orengo and Thornton2001). The majority of sequence-based methods generate descriptors for each position in the input sequence. The feature vector can be composed of multiple different descriptors: evolutionary information comprised of a position-specific scoring matrix (PSSM) or conservation score; tabular physicochemical properties of amino acid residues on specified position: hydrophobicity, polarity, solvation potential, residue interface propensities, net charge, average accessible surface area (ASA), values from the AAindex (Kawashima et al., Reference Kawashima2007) database, and others. Note, that one can also utilize structural features predicted from sequence using other tools (including ML ones) to generate more sophisticated feature vectors; for example, solvent accessible solvent area (SASA)(Dor and Zhou, Reference Dor and Zhou2007; Garg et al., Reference Garg, Kaur and Raghava2005; Yuan and Huang, Reference Yuan and Huang2004; Ahmad et al., Reference Ahmad, Gromiha and Sarai2003; Adamczak et al., Reference Adamczak, Porollo and Meller2004; Heffernan et al., Reference Heffernan2015), secondary structure information (Faraggi et al., Reference Faraggi2012; Yaseen and Li, Reference Yaseen and Li2014; Lin et al., Reference Lin2005; Bondugula and Xu, Reference Bondugula and Xu2007; Cheng et al., Reference Cheng, Sen, Jernigan and Kloczkowski2007; Pei and Grishin, Reference Pei and Grishin2004), dihedral angles (Wood and Hirst, Reference Wood and Hirst2005; Dor and Zhou, Reference Dor and Zhou2007; Xue et al., Reference Xue2008; Heffernan et al., Reference Heffernan2015; Lyons et al., Reference Lyons2014), etc. The feature vectors are then used as inputs into the machine learning algorithm. In some methods, feature vectors from several consecutive amino acid residues (typically between 7 and 25 residues) are processed together to form a new feature vector (Chen et al., Reference Chen, Huang and Gao2014, Reference Chen2015). One can feed the feature vectors to a classical classification ML model, such as support vector machine (SVM) (Cortes and Vapnik, Reference Cortes and Vapnik1995) or random forest (RF) (Ho, Reference Ho1995), which outputs probability scores for the amino acid residues to interact with the ligand (Kauffman and Karypis, Reference Kauffman and Karypis2009; Chen et al., Reference Chen, Huang and Gao2014, Reference Chen2015; Yu et al., Reference Yu2015; Lu et al., Reference Lu2019). Other methods are based on larger DL models, such as 1D-CNN, GRU, or LSTM, feeding the whole sequence at once (Cui et al., Reference Cui2019; Lee and Nam, Reference Lee and Nam2022) and also predicting a probability score for each position. Recently, large pre-trained language models have advanced many tasks in the field of natural language processing (NLP). Protein sequences can be viewed as a ‘sentence’ with amino acid residues as ‘words’, and approaches similar to ones from NLP can be applied to them. This idea brought the development of several transformer-based (Vaswani et al., Reference Vaswani2017) models, such as ESM (Lin et al., Reference Lin2023), ProtTrans (Elnaggar et al., Reference Elnaggar2021) or ProteinBert (Brandes et al., Reference Brandes2022). Most of these models utilize BERT-like (Devlin et al., Reference Devlin2018) architectures and were trained on huge databases in a self-supervised manner for the prediction of masked tokens in sequence. It was shown that such protein language models (PLM) can capture structural information, such as secondary structure or residue-residue contacts (Rives et al., Reference Rives2021). The sequence or amino acid residue embeddings derived from these models can be used as feature vectors in ML or DL models to predict different types of binding sites (Li et al., Reference Li2023d).

Figure 1. Schematic presentation of the sequence-based methods. The top part demonstrates the pipeline for a template-based approach: the target sequence is aligned against a database of template sequences with known binding residues, and the output binding residues are defined by the consensus score from the alignment. The bottom part demonstrates the pipeline for ML or DL methods. First, the feature vectors (e.g., sequence or physicochemical properties) or the embeddings (e.g., using language models) are calculated. Then, a method uses a moving window across the sequence and feeds feature vectors for each position into an ML or DL model outputting a binding score for each position, or utilizing a larger DL model to get binding scores for each position simultaneously.

Template-based

The template-based methods operate with a database of protein complexes with known binding sites (Figure 2). Then, for the query protein, they search for similar proteins and retrieve information about binding sites from them.

Some methods rely on a global comparison of a query structure against the template structures, then superimpose known ligands or positions of binding residue position from the identified similar templates (Brylinski and Skolnick, Reference Brylinski and Skolnick2008; Wass et al., Reference Wass, Kelley and Sternberg2010; Yang et al., Reference Yang, Roy and Zhang2013; Gao et al., Reference Gao2016). However, methods that rely on global comparison can miss non-conserved binding sites, as some proteins bind the same molecule at sites with different amino acid patterns (Moodie et al., Reference Moodie, Mitchell and Thornton1996; Denessiouk et al., Reference Denessiouk, Rantanen and Johnson2001). Other template-based methods incorporate more complicated local comparisons of substructures or surface patches (Figure 2). For example, one uses geometric hashing to compare two sets of graph vertices representing the query and template protein structures. These vertices can be centers of 3D cells (Rinaldis et al., Reference Rinaldis1998), surface residues (Schmitt et al., Reference Schmitt, Kuhn and Klebe2002), surface vertices (Rosen et al., Reference Rosen1998), surface patches (Shulman-Peleg et al., Reference Shulman-Peleg, Nussinov and Wolfson2004), conserved residues (Roy et al., Reference Roy, Yang and Zhang2012), or all atoms (Barker and Thornton, Reference Barker and Thornton2003; Gold and Jackson, Reference Gold and Jackson2006). Some methods utilize the maximum clique detection method (Bron and Kerbosch, Reference Bron and Kerbosch1973) to compare the two sets of residues (Lee and Im, Reference Lee and Im2013; Viet Hung et al., Reference Viet Hung2015; Konc and Janežič, Reference Konc and Janežič2010) or surfaces (Kinoshita and Nakamura, Reference Kinoshita and Nakamura2005). Other use sub-graph isomorphism (Ullmann, Reference Ullmann1976) for the comparison of two sets of pseudo-atoms representing residue side chains (Spriggs et al., Reference Spriggs, Artymiuk and Willett2003) or calculate root-mean-square-deviation (RMSD) between spatially neighboring sets of residues in the query and the template (Stark et al., Reference Stark, Sunyaev and Russell2003). There are also many approaches that use geometric methods to identify pockets in the query protein structure, and then provide a method to compare two binding sites. The comparison methods can be divided into alignment-based or alignment-free. The alignment-based methods calculate alignment for each pair of binding sites to estimate their similarity and are usually computationally demanding. On the other hand, alignment-free methods calculate translation- and rotation-invariant descriptors, which can be compared relatively fast. These methods are much faster than alignment-based approaches, but their results may be difficult to interpret (we refer the reader to this review on binding site comparison methods (Eguida and Rognan, Reference Eguida and Rognan2022)). Nonetheless, the template-based methods in general have higher interpretability, compared to the ML ones. However, the template-based methods are resource-consuming, as for each query protein one needs to screen the entire database, and the screening time increases as the database grows. They also strongly depend on the database itself – if the database lacks certain type of binding site, the method will not be able to identify such a binding site in a query.

Figure 2. Schematic presentation of the structure template-based methods. In the first stage, the target is screened against a database of template structures with known binding sites. In the second stage, the output prediction is obtained based on the most similar template structures with respect to the target.

Geometric

Geometric methods identify pockets from the protein shape by analyzing occupancy grids, surfaces, or probes, such as spheres placed around the protein. Figure 3 demonstrates a schematic overview of geometric methods for binding site detection.

SurfNet (Laskowski, Reference Laskowski1995) is one of the first geometric algorithms; it generates an occupancy grid for protein atoms and outlines the surface around the occupied voxels to determine the cavities as the binding sites. Many other methods are based on a very similar approach, which generates an occupancy grid and, for each empty grid point, calculates the fraction of directions that are enclosed by protein atoms or surfaces (Levitt and Banaszak, Reference Levitt and Banaszak1992; Hendlich et al., Reference Hendlich, Rippmann and Barnickel1997; Huang and Schroeder, Reference Huang and Schroeder2006; Weisel et al., Reference Weisel, Proschak and Schneider2007; Halgren, Reference Halgren2009; Marchand et al., Reference Marchand2021) (see Figure 3a). POCKET (Levitt and Banaszak, Reference Levitt and Banaszak1992) casts rays along three directions and determines, if a ray goes through protein–empty points–and then again protein; in this case, the point is considered to be in the pocket. Similarly, LIGSITE (Hendlich et al., Reference Hendlich, Rippmann and Barnickel1997) casts rays along seven directions (diagonals added), and if the number of intersections is higher than a threshold, the point is a pocket point. PocketPicker (Weisel et al., Reference Weisel, Proschak and Schneider2007) calculates buriedness for each grid point, scans into 30 directions with rays of length 10Å and width 0.9Å, and counts the number of intersections. SiteMap (Halgren, Reference Halgren2009) calculates the fraction of 110 rays striking the receptor within 8Å distance. SiteMap also calculates multiple descriptors to evaluate the druggability of the detected pocket. Similarly, CAVIAR (Marchand et al., Reference Marchand2021) casts rays in 14 directions, selects relevant grid points surrounded by protein atoms, and clusters the grid points forming a binding site. Another common approach involves generating two representations of the protein, corresponding to the spheres probes with two different radii placed around the protein (Peters et al., Reference Peters, Fauck and Frömmel1996; Kawabata and Go, Reference Kawabata and Go2007). More specifically, APROPOS (Peters et al., Reference Peters, Fauck and Frömmel1996) creates a Delaunay representation of a protein and, then, rolls spheres of two different radii over the structure to remove some of the sides. Shapes removed by small spheres and not removed by large ones are considered pockets (see Figure 3b). Similarly, PHECOM (Kawabata and Go, Reference Kawabata and Go2007) and (Masuya and Doi, Reference Masuya and Doi1995) roll spheres around protein atoms, and GHECOM (Kawabata, Reference Kawabata2010) and POCASA (Yu et al., Reference Yu2010) place spheres on a 3D grid. Small spheres are removed, if they intersect with large ones; after that, clusters of small spheres are considered as pockets. In a similar approach (Kim et al., Reference Kim2008), one generates inner and outer surface meshes through Voronoi diagrams with different probe radii; the pocket is then defined as the cavity between inner and outer meshes. Note, that multiple methods (Delaney, Reference Delaney1992; Kleywegt and Jones, Reference Kleywegt and Jones1994; Brady and Stouten, Reference Brady and Stouten2000) comprise an addition-removal iterative process, where at each step a buffer around the protein is added and some of the points are removed again until the pocket is identified or there is no change after the iteration (see Figure 3c). One can use grid-based approaches, where the flood fill algorithm is performed after each addition until the pocket points become enclosed and cannot be removed (Delaney, Reference Delaney1992; Kleywegt and Jones, Reference Kleywegt and Jones1994). DoGSite (Volkamer et al., Reference Volkamer2010) generates a 3D occupancy grid for the protein, then repeatedly applies Gaussian filters to remove points with values exceeding a specified threshold; the remaining grid points are clustered to form the binding site. Other heuristics can also be applied; for example, LISE (Xie and Hwang, Reference Xie and Hwang2012) generates sets of triangle motifs with assigned scores from protein atoms. Then, the method generates a 3D grid around the protein, and for each empty point, the sum of scores from triangles whose centers lie inside the voxel is assigned. In the next step, for each empty point, the score is recalculated as a sum of scores of all empty points within a sphere of radius 11Å. Finally, top-score points are selected as final pocket centers. Another example is PASS (Brady and Stouten, Reference Brady and Stouten2000), which adds spheres around triplets of protein atoms and filters them until no spheres can be added. There are many other types of geometric approaches that treat protein structures as 3D object and apply geometry-based algorithms to detect binding sites. CAST (Binkowski et al., Reference Binkowski, Naghibzadeh and Liang2003; Liang et al., Reference Liang, Woodward and Edelsbrunner1998) creates a Delaunay representation of a protein and then applies a flow theory to determine pockets. In Del Carpio et al. (Reference Del Carpio, Takahashi and Sasaki1993), the authors utilized an iterative process in which they first calculate a protein center, identify the closest surface atom, and flag all surface atoms in sight from this atom. Further, the next closest unflagged surface atom is selected and the process repeats until all atoms are flagged. In Coleman and Sharp (Reference Coleman and Sharp2006), the method calculates the surface and the convex hull, which is defined as the smallest convex polyhedron containing all the surface points. For surface points, it calculates ‘travel depth’, as the minimal distance from a point to the convex hull, and determines pockets as points with higher ‘travel depth’. In Bock et al. (Reference Bock, Garutti, Guerra, Markstein and Xu2007), the method generates a protein surface, and surface points calculates ‘spin-images’ from classical computer vision algorithms, from which the largest spheres that can be placed on a particular surface point without intersection with other surfaces are defined. Then the method clusters large spheres and outputs them as predicted pockets. MSPocket (Zhu and Pisabarro, Reference Zhu and Pisabarro2011) generates a surface and converts it to a graph, where two surface points are considered adjacent if moving these points along their normals makes them closer to each other. Then this graph is pruned, and surface points in the left subgraphs represent final pockets. CurPocket (Liu et al., Reference Liu2020b) generates a solvent-accessible surface for a protein, calculates curvature at each point, and then clusters points with high curvature. In Xie and Bourne (Reference Xie and Bourne2007), the method constructs a Delaunay tessellation of protein C $ \alpha $ atoms. Then it removes too long edges and determines edges for protein boundaries. And, finally, it calculates surface directions and geometric potentials, from which the binding site is predicted. Fpocket (Le Guilloux et al., Reference Le Guilloux, Schmidtke and Tuffery2009) is the most widely used geometric method for binding pocket detection. It operates via alpha spheres. An alpha sphere is a sphere that contacts with four atoms and does not contain atoms inside. Intuitively, small spheres should lie inside the protein, large spheres are outside, and cavities should correspond to spheres of intermediate radii. So, the algorithm consists of the following steps: (i) detection of alpha spheres via Voronoi tessellation; (ii) filtering out too small and too large spheres; (iii) clustering alpha spheres; and (iv) calculation of additional descriptors and pocket re-ranking. It is worth mentioning, that there are other geometric methods that rely on the previously mentioned assumption, that residues in protein functional sites are more conserved. These methods map conservation scores of residues onto surface points of respective residues, and cluster the most conserved points in space to get binding sites (Glaser et al., Reference Glaser2006; Pupko et al., Reference Pupko2002; Armon et al., Reference Armon, Graur and Ben-Tal2001; Nimrod et al., Reference Nimrod2008; Panchenko et al., Reference Panchenko, Kondrashov and Bryant2004; Capra et al., Reference Capra2009).

Figure 3. Schematic overview of geometric methods for binding site detection. (a) Generation of occupancy grid and calculation of the fraction of directions enclosed by the target macromolecule for each empty grid point (used, for example, in POCKET (Levitt and Banaszak, Reference Levitt and Banaszak1992), LIGSITE (Hendlich et al., Reference Hendlich, Rippmann and Barnickel1997), PocketPocker (Weisel et al., Reference Weisel, Proschak and Schneider2007), SiteMap (Halgren, Reference Halgren2009), CAVIAR (Marchand et al., Reference Marchand2021)). (b) Rolling of spheres with two different radii around the target macromolecule. The spheres with a larger radius remove the smaller ones. The remaining small spheres are clustered to get final predictions (used, for example, in APROPOS (Peters et al., Reference Peters, Fauck and Frömmel1996), PHECOM (Kawabata and Go, Reference Kawabata and Go2007), (Masuya and Doi, Reference Masuya and Doi1995), GHECOM (Kawabata, Reference Kawabata2010), and POCASA (Yu et al., Reference Yu2010)). (c) The addition-removal algorithm, is used in Delaney (Reference Delaney1992), Kleywegt and Jones (Reference Kleywegt and Jones1994), and Brady and Stouten (Reference Brady and Stouten2000). Each step consists of adding and removing the surface-exposed points until the convergence. The target macromolecule is represented with a lilac surface, and grid points and probe spheres are shown with circles.

Geometry-based methods are usually faster than other methods, but they often have lower accuracy due to the lack of information about the physicochemical and energetic properties of a protein structure.

Energetic

Most energetic methods operate with atom probes placed in a 3D grid around the protein and determine low-energy clusters (see Figure 4). In Goodford (Reference Goodford1985), the authors proposed the first probe-based method, searching for energetically favorable positions on 3D maps for three types of probes: water probe, amino group $ {\mathrm{NH}}_3^{+} $ , and methyl group $ {\mathrm{CH}}_3 $ . For this, they used three-term energy functions including Lennard-Jones, electrostatic, and hydrogen-bond potentials. In Ruppert et al. (Reference Ruppert, Welch and Jain1997), the authors used another three types of probes (hydrogen atom for hydrophobic interaction, NH for hydrogen bond donor, and C=O for hydrogen bond acceptor probe) to obtain clusters of the lowest-energy points on the protein surface. DrugSite (An et al., Reference An, Totrov and Abagyan2004), Q-SiteFinder (Laurie and Jackson, Reference Laurie and Jackson2005), and PocketFinder (An et al., Reference An, Totrov and Abagyan2005) identify binding pockets via calculation of potential energy maps with aliphatic carbon probe using Lennard-Jones potential with parameters from ECEPP/3 (Nemethy et al., Reference Nemethy1992) or GRID (Wade et al., Reference Wade, Clark and Goodford1993) force fields. SiteHound (Ghersi and Sanchez, Reference Ghersi and Sanchez2009; Hernandez et al., Reference Hernandez, Ghersi and Sanchez2009) creates maps of potential energies for six probes (methyl, phosphate oxygen, hydroxyl oxygen, peptide nitrogen, water, and carbon), where potential energy is calculated as a sum of van der Waals and electrostatic interactions with parameters from GROMOS (Van Der Spoel et al., Reference Van Der Spoel2005) force field. FTSite (Ngan et al., Reference Ngan2012; Brenke et al., Reference Brenke2009) places 16 small molecular probes (ethanol, isopropanol, isobutanol, acetone, acetaldehyde, dimethyl ether, cyclohexane, ethane, acetonitrile, urea, methylamine, phenol, benzaldehyde, benzene, acetamide, and N,N-dimethylformamide) on a dense grid around the protein, optimizes their positions with extended energy expression using CHARMM force field (Brooks et al., Reference Brooks1983) and obtains low-energy clusters of probes. AutoSite (Ravindranath and Sanner, Reference Ravindranath and Sanner2016) calculates affinity maps for hydrophobic (carbon) and hydrophilic (oxygen, hydrogen) probes using the van der Waals interaction term from AutoDock energy function (Morris et al., Reference Morris2009; Huey et al., Reference Huey2007). SuperStar (Verdonk et al., Reference Verdonk2001) uses a slightly modified approach. This method places four different probes, $ {\mathrm{NH}}_3^{+} $ nitrogen atom, carbonyl oxygen atom, hydroxyl oxygen atom, and methyl carbon atom, into the grid, and converts these maps according to distributions of densities observed in a database of crystallographic structures. In Tsujikawa et al. (Reference Tsujikawa2016), the atom probe approach in addition takes into account the conservation of amino acid residues. The method places a carbon atom probe, calculates van der Waals energies for them, and, then, weights interaction energy by conservation scores of nearby amino acid residues. Another class of energy-based methods runs MD simulations and retrieves information about binders from trajectory analysis. OMD (Bhinge et al., Reference Bhinge2004) runs short MD simulations of protein in water and then determines binding pockets as volumes, where the RMSD of solvent molecules within the trajectory is low. SILCS (Faller et al., Reference Faller and Klon2015) runs an MD simulation of protein with water solvent and multiple small molecule fragments. It defines what protein regions are more likely to be occupied by which small molecule types. PlayMolecule CrypticScout (Martinez-Rosell et al., Reference Martinez-Rosell2020) runs mixed-solvent MD with benzene molecules and defines binding hotspots as regions with high occupancy of benzene molecules or regions with low RMSD for these molecules within the trajectory. Another approach is utilized in MDPA (Gu et al., Reference Gu, Li and Ming2022). It is based on the assumption, that ligand binding occurs in regions with higher conformational dynamics (Ming and Wall, Reference Ming and Wall2006). To calculate the external interaction of proteins with test points, the method treats proteins as elastic network structures and simulates interactions using connected springs.

Figure 4. Schematic presentation of the energy probe-based methods. (a) Different probes (shown as red, blue, and green circles) are placed on a 3D grid around the target macromolecule (shown as a lilac surface) and their interaction energies with the target’s atoms are calculated. (b) The probes corresponding to the high-energy values are filtered out. (c) The remaining probes are clustered. (d) The filtering procedure is applied to remove non-relevant clusters.

Generally, energetic methods have higher accuracy than geometric methods and high interpretability. However, they are computationally expensive and may miss some interactions not covered by the existing probe types.

Machine learning-based

Most machine learning-based methods can be described in the following way. Firstly, they calculate feature vectors for amino acid residues in the input protein; then, the method feeds the feature vectors into an ML classifier, which outputs the probability of a residue being in the binding site. Then, the method spatially clusters high-scoring residues to get the binding site composition (Figure 5). Feature vectors can contain sequential (conservation), physicochemical (electrostatics, hydrogen bonds, solvation energy, hydrophobicity, atom types), geometrical, or structural (solvent accessibility, secondary structure, local geometry) descriptors. In Gutteridge et al. (Reference Gutteridge, Bartlett and Thornton2003), the method identifies catalytic residues in enzymes. It calculates multiple descriptors for each residue: conservation score, relative solvent accessibility, secondary structure, and closeness to a cleft identified by Surfnet (Laskowski, Reference Laskowski1995), and residue depth. This feature vector is used as input for a single-layer NN. Petrova and Wu (Reference Petrova and Wu2006) calculates sequential and structural properties of residues (conservation, flexibility, solvent accessibility, position on the protein surface, hydrogen bonds, secondary structure) and classifies them using an SVM. Tong et al. (Reference Tong2009) calculate electrostatic features, geometric properties, and sequence-based conservation for each residue and classify them using a maximum-likelihood algorithm. Qiu and Wang (Reference Qiu and Wang2011) calculates eight structural properties (solvent accessible surface area, solvation energy, hydrophobicity, depth index, protrusion index, preference, theoretical b-factor) for residues, and uses a Random Forest classifier. ISMBLab-LIG (Jian et al., Reference Jian2016) first calculates 3D probability density maps that describe interacting atom types around the protein surface using pre-calculated distributions from a database. Then, for each surface atom, the method collects a feature vector of surface local geometry combined with properties retrieved from the described density maps. The method uses a NN model as a classifier. GRaSP (Santana et al., Reference Santana2020) retrieves a set of physicochemical properties (solvent relative accessibility, atom types, interaction level) from an atomic graph, and then uses an extremely randomized tree to classify residues as binding/non-binding.

Figure 5. Schematic presentation of the machine learning-based methods. On the top, the target structure is represented as a surface, and feature vectors are calculated for the surface points. On the bottom, feature vectors are calculated for the target’s residues or atoms. Then, an ML classifier predicts the binding scores for the points, residues, or atoms, based on the input feature vectors. Finally, the output predictions are filtered by a score threshold and clustered.

Another approach is to classify points or patches on the protein surface instead of residues. Bradford and Westhead (Reference Bradford and Westhead2005) classifies surface points, where for each point a feature vector consists of seven properties: shape index, curvedness, conservation, electrostatic potential, hydrophobicity, residue interface propensity, and solvent accessible surface area. P2Rank (Krivák and Hoksza, Reference Krivák and Hoksza2018) generates a protein surface, and projects features calculated for protein atoms to surface points. Afterward, it predicts the ligandability of each point using an RF classifier and clusters points with high scores. Similarly, SiteFerret (Gagliardi and Rocchia, Reference Gagliardi and Rocchia2023) first calculates a set of features for surface points, incorporating information about cavities, and then classifies them using the Isolation Forest method. It is also possible to directly classify cavities detected by a geometric method. SCREEN (Nayal and Honig, Reference Nayal and Honig2006) first identifies all possible cavities on the protein surface. After that, it calculates a large set of cavity descriptors of different types, such as cavity size, electrostatics, hydrogen bonding, hydrophobicity, polarity, amino acid composition, rigidity, secondary structure, and cavity shape. An RF model classifies cavities as drug-binding/non-drug-binding and selects a smaller fraction of relevant descriptors. FEATURE (Bagley and Altman, Reference Bagley and Altman1995, Reference Bagley and Altman1996; Wei and Altman, Reference Wei and Altman1998, Reference Wei and Altman2003) represents a set of tools for the prediction of binding sites of different types, such as calcium binding (Wei and Altman, Reference Wei and Altman1998), ATP binding (Wei and Altman, Reference Wei and Altman2003), serine protease active sites (Bagley and Altman, Reference Bagley and Altman1996), and others (Liang et al., Reference Liang2003). It represents microenvironments around a protein as concentric shells with centers placed on a grid and calculates physicochemical properties within these shells. These properties are compared with a set of features for known binding sites and non-binding sites. Then, the Bayesian classifier (Friedman et al., Reference Friedman, Geiger and Goldszmidt1997) is used to distinguish binding sites from non-binding ones. There are also consensus-based methods that retrieve predictions from multiple other methods, including geometric, energy-based, or template-based approaches and combine them into final predictions using an ML-based re-scoring. For example, MetaPocket2.0 (Zhang et al., Reference Zhang2011) aggregates results from eight different methods: $ {\mathrm{LIGSITE}}^{csc} $ (Huang and Schroeder, Reference Huang and Schroeder2006), PASS (Brady and Stouten, Reference Brady and Stouten2000), Q-SiteFinder (Laurie and Jackson, Reference Laurie and Jackson2005), Surfnet (Laskowski, Reference Laskowski1995), Fpocket (Le Guilloux et al., Reference Le Guilloux, Schmidtke and Tuffery2009), GHECOM (Kawabata, Reference Kawabata2010), ConCavity (Capra et al., Reference Capra2009), and POCASA (Yu et al., Reference Yu2010). Another example is COACH (Yang et al., Reference Yang, Roy and Zhang2013), which combines prediction results from TM-SITE (Yang et al., Reference Yang, Roy and Zhang2013), S-SITE (Yang et al., Reference Yang, Roy and Zhang2013), COFACTOR (Roy et al., Reference Roy, Yang and Zhang2012), FINDSITE (Brylinski and Skolnick, Reference Brylinski and Skolnick2011), and ConCavity (Capra et al., Reference Capra2009).

ML-based methods strongly rely on the dataset construction and calculated feature vectors, and can produce false positive predictions – that is, identification of ‘undruggable’ regions (Broomhead and Soliman, Reference Broomhead and Soliman2017). Moreover, even extensive feature engineering does not guarantee capturing all the information relevant to the binding site prediction.

Deep learning-based

The accumulation of large amounts of structural data and advancements in deep learning methods in other fields have led to the development of top-performing methods in structural bioinformatics problems, such as protein structure prediction (Jumper et al., Reference Jumper2021; Baek et al., Reference Baek2021) DL-based approaches may not require hand-crafted feature engineering and may be capable of capturing relevant structural context by construction. To begin with, these methods typically operate on protein structures represented as a point cloud, a graph, or a 3D density grid. Although a graph is typically a 2D representation of a structure, the 3D information can be encoded as the feature vectors of graph nodes or edges. Therefore, most of the DL-based methods can be classified based on the structure representation, and further split into two classes: (i) where a DL model performs segmentation using the entire graph or grid; and (ii) where a DL model samples sub-graphs or sub-grids and classifies whether their centers correspond to a binding site or not. Figure 6 demonstrates a schematic representation of this idea, and we provide more details about methods in each class below.

Figure 6. Schematic presentation of the DL-based methods. Most of the methods utilize graph-based or voxel grid representations of the target macromolecular structure. Then, they sample either sub-graphs or sub-grids around the structure and classify their centers as belonging to the binding site or not. Alternatively, they use segmentation models to operate with the full graph or grid.

We start with methods that sample small 3D voxel grids around a protein structure and classify whether the grid center corresponds to a binding site. DeepSite (Jiménez et al., Reference Jiménez2017) was one of the first methods developed for predicting ligand binding sites. It represents a protein structure as a 3D voxel grid with 1Å voxels in size, where each voxel contains eight channels for atoms of different types: hydrophobic, aromatic, hydrogen bond acceptor, hydrogen bond donor, positive ionizable, negative ionizable, metal, and excluded volume. Each channel of a voxel stores the occupancy value of nearby atoms of the respective type, where occupancy is calculated as $ n(r)=1-\mathit{\exp}\left(-{\left({r}_{vdw}/r\right)}^{12}\right) $ . Then, from the generated 3D voxel grid of a protein, subgrid cubes of size $ 16\times 16\times 16 $ voxels are sampled through a sliding window. These cubes are provided as input into a 3D CNN, which outputs a probability score for the cube center being closer than 4Å to the geometric center of a binding site. In Jiang et al. (Reference Jiang2019a), the authors utilized a similar approach but parameterized the input voxel grid differently. They calculate four pseudo-energy channels instead of using an occupancy-based grid: (i) a shape channel retrieved as output from the LIGSITE (Hendlich et al., Reference Hendlich, Rippmann and Barnickel1997) method; (ii) a van der Waals potential energy channel of an -CH3 probe; (iii) a hydrogen bond potential channel using an -OH probe; and (iv) an electric potential energy channel. DeepPocket (Aggarwal et al., Reference Aggarwal2021) re-scores pockets predicted by Fpocket (Le Guilloux et al., Reference Le Guilloux, Schmidtke and Tuffery2009) using a similar 3D CNN model. For this, it uses libmolgrid (Sunseri and Koes, Reference Sunseri and Koes2020) to obtain a cubic 3D voxel grid of size 23.5Å with a voxel size of 0.5Å around a pocket and passes it into a 3D CNN model that classifies the pocket as ligandable or non-ligandable. After that, the method also generates a segmented representation of ligandable pockets by passing a larger (32Å) cubic grid into another U-Net-like (Ronneberger et al., Reference Ronneberger, Fischer and Brox2015) 3D CNN. DeepSurf (Mylonas et al., Reference Mylonas, Axenopoulos and Daras2021), CAT-Site (Petrovski et al., Reference Petrovski, Hribar-Lee and Bosnić2022), and SAPocket (Wang et al., Reference Wang, He and Zhu2023c) instead of using a sliding window, sample points on the protein surface, calculate voxelized representations for cubic grids centered on these points and pass these grids as input into a 3D CNN classification model. FRSite (Jiang et al., Reference Jiang2019b) utilizes the faster R-CNN approach (Ren et al., Reference Ren2016): it first passes a voxelized representation of a protein into a 3D Region Proposal Network, and further feeds proposals into a 3D CNN for classification. BiteNet (Kozlovskii and Popov Reference Kozlovskii and Popov2020) utilizes the YOLO approach for real-time object detection in videos (Redmon et al., Reference Redmon2016). It first obtains a voxelized representation of an input protein and applies a 3D CNN model to it. The model splits the input grid into cells, where each cell contains predicted values for the probabilities of a binding site center being within the cell and the center of a binding site with respect to the cell. Kalasanty (Stepniewska-Dziubinska et al., Reference Stepniewska-Dziubinska, Zielenkiewicz and Siedlecki2020) was one of the first methods to perform 3D segmentation of binding sites in a single pass. The method represents a protein as a 3D grid of a constant size of 70Å along each direction with a 2Å voxel size and feeds it into a 3D U-Net (Ronneberger et al., Reference Ronneberger, Fischer and Brox2015). There are multiple follow-up methods of the Kalasanty approach with adjusted 3D CNN model (Kandel et al., Reference Kandel, Tayara and Chong2021; Li et al., Reference Li2022; Li et al., Reference Li2023b; Nazem et al., Reference Nazem2021; Liu et al., Reference Liu2023). PUResNet (Kandel et al., Reference Kandel, Tayara and Chong2021) modified the encoder in the U-Net model to ResNet. RefinePocket (Liu et al., Reference Liu2023) used an attention-enhanced encoder and a mask-guided decoder inside the U-Net. RecurPocket (Li et al., Reference Li2022) and GLPocket (Li et al., Reference Li2023b) used a recurrent LMSER (Least Mean Square Error Reconstruction) network with gated recurrent refinement. DUNet (Wang et al., Reference Wang2022b) added a DenseNet (Huang et al., Reference Huang2017) encoder into the U-Net model. InDeep (Mallet et al., Reference Mallet2022) used a 3D U-Net for the prediction and segmentation of small molecule binding sites occurring on protein–protein interfaces. PointSite (Yan et al., Reference Yan2022) represents a point cloud-based segmentation approach: it constructs a point cloud from all protein atoms, converts it into a 3D sparse grid, and applies a segmentation model with a U-Net architecture based on submanifold sparse convolutions.

A different approach is to consider the 3D structure of a protein as a graph. Some methods sample points on the protein surface or around the protein on a grid, and analyze the graph constructed from these points, that is predicting whether a point corresponds to a binding site or not. MaSIF (Gainza et al., Reference Gainza2020) pre-calculates physicochemical properties for the whole protein surface. Then, from this surface, the method samples patches represented as a graph with surface points as nodes and with surface point feature vectors as node feature vectors, containing geometric (shape index, distance-dependent curvature) and chemical (hydropathy, continuum electrostatics, free electrons/protons) descriptors with additional values for geodesic coordinates. Then, a GNN with geodesic convolutions is applied to the patch graph with multiple orientations to obtain an embedding vector for the input patch. The authors used this approach for different tasks: prediction of protein–ligand binding sites, prediction of protein–protein binding sites, and fast scanning of protein surfaces for identification of protein–protein binder partners. In dMaSIF (Sverrisson et al., Reference Sverrisson2021), the authors further improved this approach and made the method fully differentiable without the need for memory- and computation-demanding pre-calculation of surface descriptors. SiteRadar (Evteev et al., Reference Evteev, Ereshchenko and Ivanenkov2023) selects points on a 3D grid outside of the protein, then generates a graph with protein atoms around the selected points as the nodes, and uses a GNN to analyze the graphs, predicting whether the point center belongs to a binding site or not. PocketAnchor (Li et al., Reference Li2023c) samples a set of ‘anchor’ points around the protein, representing potentially ligandable positions. For each ‘anchor’ point, it further generates a graph with protein atoms and surface points within 6Å from the ‘anchor’ center. Atom and surface point nodes contain a different number of geometric and chemical features. The graphs are processed with MPNNs, outputting binding site scores for each ‘anchor’ point. Of note, GraphSite (Shi et al., Reference Shi2022) uses a graph representation of local protein regions and utilizes GNNs to classify ligand binding sites into 14 classes. A higher-level approach is to provide the whole protein graph as input into a graph neural network model for segmentation. For example, GraphBind (Xia et al., Reference Xia2021) uses residue centroids as node centers and structural and sequential features of residues as node features and passes the graph into a Hierarchical Graph Neural Network, which predicts a score for each residue indicating whether it is on the binding interface with nucleic acid. Similarly, GraphPLBR (Wang et al., Reference Wang2023d) operates on residues as nodes. FABind (Pei et al., Reference Pei2023) uses two separate GNNs, one for working with the protein residue graph and another operating on the ligand atomic graph. Embeddings from the two models are combined to make a ligand-specific prediction of binding residues. GrASP (Smith et al., Reference Smith2023) builds a graph using all protein atoms within 5Å from the surface and uses a GNN with graph attention to classify atoms as binding or non-binding. Similarly, GU-Net (Nazem et al., Reference Nazem2023) uses all protein atoms for the graph and predicts atom scores using a U-Net-like Graph Convolutional Network (Gao and Ji, Reference Gao and Ji2019). EquiPocket (Zhang et al., Reference Zhang2023) builds graphs using both protein atoms and surface points and applies GNNs with E(3)-equivariant convolutions (Satorras et al., Reference Satorras, Hoogeboom and Welling2021) to identify binding atoms. LigBind (Xia et al., Reference Xia, Pan and Shen2023) demonstrated another approach: it first pre-trains GNNs for a ligand-general binding residue predictor and a feature extractor for ligand-residue pair embeddings, and then fine-tunes ligand-specific binding residue predictors for more than 1000 ligand types from the BioLip (Yang et al., Reference Yang, Roy and Zhang2012) database.

Recently, models for protein structure prediction have made significant advances. A breakthrough occurred in the 14th Critical Assessment of Protein Structure Prediction (CASP14) challenge (Kryshtafovych et al., Reference Kryshtafovych2021) when AlphaFold2 (Jumper et al., Reference Jumper2021) achieved almost experimental accuracy in the prediction of full-atom protein structures. AlphaFold2 takes an MSA and structural templates as inputs and utilizes a complicated deep-learning model with the newly introduced Evolutionary Transformer. Some approaches for binding site prediction use these methods to generate protein structure models from sequences alone, and retrieve structural features along with sequential ones for each residue (Littmann et al., Reference Littmann2021; Ho et al., Reference Ho2021; Seo et al., Reference Seo2024). Furthermore, some models extract a structural graph from a generated protein model, where residues correspond to the graph nodes and the their features correspond to the node features. Then, the composed graph is used as the input for a graph neural network (GNN) (Yuan et al., Reference Yuan2022a; Zhang and Xie, Reference Zhang and Xie2023). Previously, it was shown that AlphaFold2 can successfully predict protein–peptide structures (Chang and Perez, Reference Chang and Perez2023; Tsaban et al., Reference Tsaban2022), but it was not clear whether AlphaFold2 could be used for the analysis of interactions between proteins and small molecules, as the latter were absent in the training objective for modeling. However, in some cases, AlphaFold2 predicts rotamers as if they were interacting with small molecules, suggesting that it can be used to train a binding site detection model. Moreover, as AlphaFold2 can predict protein–peptide complexes, it can be reasoned that this model can also be useful for the identification of interactions with small molecules, as they or their fragments can resemble amino-acid side chains (Polizzi and DeGrado, Reference Polizzi and DeGrado2020). AF2BIND (Gazizov et al., Reference Gazizov2023) is constructed as follows: as input to the AlphaFold2 model, it provides a sequence of a target protein, the protein backbone structure as a template, and 20 ‘bait’ amino acids as individual chains, appending them to the sequence with large offsets. The method further uses AlphaFold2 output pairwise representations between target residues and each of the twenty ‘bait’ amino acids as input into a logistic regression model, predicting whether a target residue is ligand binding or not. The authors demonstrated a correlation between the chemical properties of the small molecule ligands and the 20 ‘bait’ amino acids.

Finally, there are methods combining multiple representations of the protein. PocketMiner (Meller et al., Reference Meller2023) uses a geometric vector perceptron GNN (GVP-GNN) (Jing et al., Reference Jing2020) and a 3D CNN for the prediction of putative cryptic pockets. For this, the authors generated 40-ns simulations for 37 proteins and trained the models to predict the positions in each structure where a pocket would open during a short simulation. The authors showed that both GVP-GNN and 3D CNN work equally well.

Benchmarks

Most of the newest ML- and DL-based methods rely on scPDB (Desaphy et al., Reference Desaphy2015), PDBbind (Liu et al., Reference Liu2015), or BioLip (Yang et al., Reference Yang, Roy and Zhang2012) databases for training and validation. scPDB (Desaphy et al., Reference Desaphy2015) is a large database containing $ \sim $ 16,000 complexes, where each entry is annotated with calculated properties for the ligand, cavity, and interactions. PDBbind (Liu et al., Reference Liu2015) is a curated database of protein–ligand complexes ( $ \sim $ 23,000), with experimentally determined binding affinity. BioLip (Yang et al., Reference Yang, Roy and Zhang2012; Zhang et al., Reference Zhang2024) contains $ \sim $ 460,000 structures of proteins or nucleic acids, with a total of approximately $ \sim $ 890,000 ligands. BioLip includes a wide range of classes of macromolecules and ligands, allowing researchers to construct various training and validation sets. In addition, it incorporates a comprehensive procedure to select relevant ligands and includes cross-references with many other databases (PDBbind (Liu et al., Reference Liu2015), BindingDB (Gilson et al., Reference Gilson2016), SIFTS (Dana et al., Reference Dana2019), UniProt (Consortium, Reference Consortium2019), and DrugBank (Knox et al., Reference Knox2024), etc.). Note that the provided numbers are for 2024; these are likely to increase in future versions of the databases. Finally, other approaches rely on training datasets compiled from PDB, followed by structure refinement, clustering, and filtering of redundant structures.

There are two benchmark sets, COACH420 (Krivák and Hoksza, Reference Krivák and Hoksza2018) and HOLO4K (Schmidtke et al., Reference Schmidtke2010), which are widely used for the comparison of binding site detection methods. COACH420 (Krivák and Hoksza, Reference Krivák and Hoksza2018) is a dataset of 420 single-chain proteins containing natural compounds and drug-like ligands. It was first created for the evaluation of the P2Rank method (Krivák and Hoksza, Reference Krivák and Hoksza2018) as a subset of a test set from (Roy et al., Reference Roy, Yang and Zhang2012; Yang et al., Reference Yang, Roy and Zhang2013), without proteins from the training set of P2Rank. HOLO4K (Schmidtke et al., Reference Schmidtke2010), in turn, is a large set of protein–ligand complexes. It was initially composed for the validation of the PocketFinder (An et al., Reference An, Totrov and Abagyan2005) method and later used for a comprehensive large-scale comparison of binding site prediction methods (Schmidtke et al., Reference Schmidtke2010). Interestingly, originally, it comprised apo complexes; but after the work of (Krivák and Hoksza, Reference Krivák and Hoksza2018), a subset of holo complexes is mainly used. It is important to note that, although the COACH420 and HOLO4K benchmarks are used by many methods, most of them perform additional filtering (e.g., removing irrelevant ligands or addressing data leakage between the training and test sets), resulting in slightly different subsets of COACH420 and HOLO4K. Therefore, a direct comparison of methods based on these benchmarks may not be as straightforward as it may seem. Nonetheless, one may see the performance metrics of different methods in Supplementary Tables S2S10.

The other benchmarks include: CHEN11 (Chen et al., Reference Chen2011), B48/U48 (Huang and Schroeder, Reference Huang and Schroeder2006), B210 (Huang and Schroeder, Reference Huang and Schroeder2006), DT198 (Zhang et al., Reference Zhang2011), ASTEX (Hartshorn et al., Reference Hartshorn2007), and CASP (Lopez et al., Reference Lopez, Ezkurdia and Tress2009; Schmidt et al., Reference Schmidt2011; Gallo Cassarino et al., Reference Gallo Cassarino, Bordoli and Schwede2014). CHEN11 (Chen et al., Reference Chen2011) is a non-redundant dataset of 251 proteins, where each structure is the most representative structure of a family, with a ligand superimposed from the closest homolog in cases where a ligand is absent in the original structure. B48/U48 (Huang and Schroeder, Reference Huang and Schroeder2006) is a small dataset of pairs of apo and holo-structures of the same protein. The Astex Diverse set (Hartshorn et al., Reference Hartshorn2007) is a small benchmark for docking methods, which was used as the binding site detection benchmark (Le Guilloux et al., Reference Le Guilloux, Schmidtke and Tuffery2009; Yan et al., Reference Yan2022). Some of the older classical methods used CASP8 (Lopez et al., Reference Lopez, Ezkurdia and Tress2009), CASP9 (Schmidt et al., Reference Schmidt2011), and CASP10 (Gallo Cassarino et al., Reference Gallo Cassarino, Bordoli and Schwede2014) benchmarks to evaluate their performance for the prediction of ligand-binding residues. However, these benchmarks are much smaller compared to the other ones described above. Finally, we would like to note that, while for older methods, the most used metrics correspond to binary classification metrics derived from the residue scores, newer methods include metrics based on the distances between the predicted binding site center and the true binding site, as well as the overlap of the predicted and true binding site cavity in the case of binding site segmentation. Supplementary Section Metrics provides more details on commonly used performance metrics for binding site prediction methods.

Protein–peptide binding sites

Protein–protein interactions (PPIs) regulate numerous essential biological pathways, making them a key class of pharmacological targets (Ruffner et al., Reference Ruffner, Bauer and Bouwmeester2007). There is an increasing need to develop inhibitors of intracellular PPIs to modulate critical biological processes. However, PPIs have long been considered difficult to target (Tsomaia, Reference Tsomaia2015). On the one hand, large biologics, which are effective in targeting extracellular PPIs, cannot penetrate cell membranes to reach intracellular PPIs. On the other hand, traditional small molecule scaffolds can cross membranes but are often unsuitable for the large, shallow surfaces typical for PPI interfaces (Tsomaia, Reference Tsomaia2015). PPI interfaces exhibit distinct characteristics, such as larger contact areas ( $ \sim 1500-3000{\mathring{A}}^2 $ for PPI compared to $ \sim 300-1000{\mathring{A}}^2 $ for protein–small molecule interactions (Smith and Gestwicki, Reference Smith and Gestwicki2012)) and the absence of deep binding pockets usually found in small molecule interactions ( $ \sim 270{\mathring{A}}^3 $ in volume (Buchwald, Reference Buchwald2010)). Notably, PPI interfaces often contain smaller binding pockets ( $ \sim 100{\mathring{A}}^3 $ (Fuller et al., Reference Fuller, Burgoyne and Jackson2009)) that play a crucial role in binding affinity (Clackson and Wells, Reference Clackson and Wells1995). Peptides and peptide-based molecules occupy a unique position between small molecules (with a molecular weight $ <0.5 kDa $ ) and biologics ( $ >150 kDa $ ). They offer a promising therapeutic approach for targeting intracellular PPIs, as they can potentially combine the benefits of biologics, such as low toxicity, high specificity, and strong affinity, with the membrane permeability of small molecules (Tsomaia, Reference Tsomaia2015). The successful design of therapeutic peptides requires detailed knowledge of the binding sites on their protein targets. Identifying new protein–peptide binding sites could broaden the range of druggable targets, opening up new opportunities for drug discovery. Many methods for protein–peptide binding site prediction utilize approaches similar to the ones described in the previous section, but we still cover these methods here to highlight some specific characteristics. See Table 2 for an extensive list of methods for prediction of protein-peptide binding sites.

Table 2. List of methods for prediction of protein–peptide binding sites

Machine learning-based

Multiple sequence-based methods calculate features for each residue (e.g., PSSM, predicted ASA, SS, physicochemical properties, and intrinsic disorder) in an input protein sequence and pass these features as input into a classical ML model (Taherzadeh et al., Reference Taherzadeh2016; Zhao et al., Reference Zhao, Peng and Yang2018; Iqbal and Hoque, Reference Iqbal and Hoque2018; Shafiee et al., Reference Shafiee, Fathi and Taherzadeh2022). More advanced approaches tend to rely on additional information. SPRINT-Str (Taherzadeh et al., Reference Taherzadeh2018) and Multi-VORFFIP (Segura et al., Reference Segura, Jones and Fernandez-Fuentes2012) calculate structural and physicochemical descriptors for each residue in a target protein and use RF for the binary classification of residues as binding/non-binding. PINUP (Liang et al., Reference Liang2006) calculates structural and physicochemical descriptors for interface residues, then selects surface patches by choosing a central surface residue and 19 residues nearest to it, and then classifies the patch based on a set of features for the 20 patch residues. P2Rank-Pept (Krivák et al., Reference Krivák, Jendele and Hoksza2018) calculates geometrical and physicochemical descriptors for protein surface points and classifies these points using RF. PepSite (Trabuco et al., Reference Trabuco2012) uses spatial PSSMs for the identification of peptide-binding hot spots on the protein surface. For this, the method estimates the densities of protein atoms around each amino acid type in the peptide and encodes them into a 3D grid. Then, PepSite screens the target protein with these S-PSSM grids and identifies appropriate hot spots. PepBind (Zhao et al., Reference Zhao, Peng and Yang2018) is a consensus method combining predictions from SVMpep, S-SITE, and TM-SITE.

Deep learning-based

The sequence-based approaches, such as VisualP (Wardah et al., Reference Wardah2020) encode a window around a residue into a 2D image and apply a CNN. MTDSite (Sun et al., Reference Sun2021) uses a BiLSTM to predict binding residues for DNA, RNA, carbohydrates, and peptides. PepBCL (Wang et al., Reference Wang2022a) and PepNN-Seq (Abdin et al., Reference Abdin2022) retrieve protein sequence embeddings from the language model ProtTrans (Elnaggar et al., Reference Elnaggar2021). Similarly to the machine learning-based methods, recent deep learning-based approaches tend to incorporate different types of information into the model. PepCNN (Chandra et al., Reference Chandra2023) represents residues using sequential and structural descriptors, along with embeddings from the ProtT5 (Elnaggar et al., Reference Elnaggar2021) model, and passes them into a 1D CNN model. PepNN-Struct (Abdin et al., Reference Abdin2022) uses a GNN with attention to extract embeddings from a graph of protein residues and uses multi-head attention to encode a peptide sequence for predicting of binding residues. The authors also demonstrated that pre-training on protein–protein complexes significantly increases the model accuracy in predicting peptide-binding residues. GraphPPepIS (Li et al., Reference Li2023a) represents both protein and peptide structures as graphs and passes them into a GCN, extracting binding residues on both the protein and peptide sides. GAPS (Zhu et al., Reference Zhu2023) encodes a protein into a point cloud of atoms and uses a geometric attention-based network to classify atoms as binding or non-binding. BiteNet $ {}_{Pp} $ (Kozlovskii and Popov, Reference Kozlovskii and Popov2021a) represents peptide binding sites as a set of hotspots and utilizes an approach similar to BiteNet (Kozlovskii and Popov, Reference Kozlovskii and Popov2020): it encodes an input protein into a 3D voxel grid and feeds it into a 3D CNN, which splits the grid into cells containing probabilities of a peptide binding site hotspot being in the cell and hotspot center coordinates. DeepProSite (Fang et al., Reference Fang2023) builds a model using ESMFold (Rives et al., Reference Rives2021), retrieves embeddings using the ProtTrans (Elnaggar et al., Reference Elnaggar2021) model, and feeds the graph into a Graph Transformer network (Ingraham et al., Reference Ingraham2019) afterward to predict protein–protein and protein–peptide binding sites.

Template- and energy-based methods

There are a few template-based and energy-based approaches. For example, SPOT-peptide (Litfin et al., Reference Litfin, Yang and Zhou2019) and InterPep (Johansson-Åkhe et al., Reference Johansson-Åkhe, Mirabello and Wallner2019) screen a query protein against a database of known protein–peptide complexes. Energy-based methods sample small molecule probes around a protein and cluster low-energy conformations to get final predictions. PeptiMap (Lavi et al., Reference Lavi2013) adapts the FTmap (Brenke et al., Reference Brenke2009) method for protein–small molecule binding site prediction with additional post-processing for filtering out irrelevant sites. ACCLUSTER (Yan and Zou, Reference Yan and Zou2014) scans a protein surface with 20 amino acid probes. In Verschueren et al. (Reference Verschueren2013), the method uses polypeptide fragments from the BriX (Vanhee et al., Reference Vanhee2011) database mapped around the target protein and generates ensembles of energetically favorable protein–peptide complexes.

Benchmarks

For protein–peptide binding sites, the most widely used benchmark is TS125, which is a test set from SPRINT-Seq (Taherzadeh et al., Reference Taherzadeh2016), constructed as a non-redundant subset of 1,279 protein–peptide complexes from the BioLip database (Yang et al., Reference Yang, Roy and Zhang2012). Other benchmarks include TS092, TS251, and TS639. TS092 is a test benchmark from PepNN (Abdin et al., Reference Abdin2022), designed as a subset of protein–peptide complexes from the PDB, submitted after a specific date and having a sequence identity lower than $ 30\% $ with all protein targets in the training set. The TS251 benchmark from InterPep (Johansson-Åkhe et al., Reference Johansson-Åkhe, Mirabello and Wallner2019) was constructed such that the TM-score (Zhang and Skolnick, Reference Zhang and Skolnick2005) of the protein structures is lower than $ 0.5 $ with all the structures in the template database. Finally, TS639 from PepBind (Zhao et al., Reference Zhao, Peng and Yang2018) is a different subset of T1279, used for training and validation of SPRINT-Seq (Taherzadeh et al., Reference Taherzadeh2016), described above. Table 3 lists performance metrics (AUC and MCC, see also Supplementary Section Metrics) for the protein–peptide binding site prediction methods. As one can see, the top methods are ML- or DL-based, with $ {\mathrm{BiteNet}}_{Pp} $ (Kozlovskii and Popov, Reference Kozlovskii and Popov2021a) being the top-performing one.

Table 3. Performance of protein–peptide binding site detection methods on test benchmarks retrieved from Kozlovskii and Popov (Reference Kozlovskii and Popov2021a), Abdin et al. (Reference Abdin2022), and Fang et al. (Reference Fang2023)

Nucleic acid–small molecule binding sites

RNA molecules are emerging as a significant class of pharmacological targets (Warner et al., Reference Warner, Hajdin and Weeks2018). Efforts in RNA-targeting drug discovery span various approaches, such as designing stabilizers for DNA G-quadruplexes (Ortiz de Luzuriaga et al., Reference Ortiz de Luzuriaga, Lopez and Gil2021), developing antibiotics that target riboswitches (Panchal and Brenk, Reference Panchal and Brenk2021), using antisense RNA (McClorey and Wood, Reference McClorey and Wood2015), and creating RNA-targeting antivirals. RNA targets that expand the druggable genome, including those associated with ‘undruggable’ proteins or non-coding microRNAs, hold particular promise (Matsui and Corey, Reference Matsui and Corey2017). However, the development of RNA-targeted drugs faces significant challenges, such as limited chemical diversity and the dynamic nature of RNA structures (Falese et al., Reference Falese, Donlic and Hargrove2021). To advance RNA-targeting drug discovery, efficient tools for detecting structure-specific RNA-small molecule binding sites are needed.

There are many approaches targeting binding sites on proteins; however, there is a limited number of methods for nucleic acids. Table 4 provides a list of methods for prediction of nucleic acid-small molecule binding sites.

Table 4. List of methods for prediction of nucleic acid–small molecule binding sites

Knowledge-based

Firstly, there are several knowledge-based methods. Rsite (Zeng et al., Reference Zeng2015) and Rsite2 (Zeng and Cui, Reference Zeng and Cui2016) calculate distances between nucleotides based on tertiary and secondary structures, respectively, and determine nucleotides that are the most distant from others as the binding nucleotides. Similarly, RBind (Wang, Jian, et al., Reference Wang2018b) calculates the degree and closeness of nodes in a nucleotide network and determines binding nucleotides as those with values exceeding a specified threshold. RNetsite (Liu et al., Reference Liu2024) represents an RNA molecule as a graph and calculates local (degree, neighborhood connectivity) and global (betweenness centrality, closeness, and eccentricity) properties for each node of the graph. Then, each node is classified as binding or non-binding based on the property statistics computed from a reference set of RNA molecules.

Energetic

To the best of our knowledge, only two methods use an energy-based approach. SILCS-RNA (Kognole et al., Reference Kognole, Hazel and MacKerell2022) runs simulations of a target macromolecule in a mixed solvent with eight different probes. From these simulations, the method calculates a 3D grid with energy maps, which can be used for binding site identification, docking, and binding affinity evaluation tasks. SHAMAN (Panei et al., Reference Panei, Gkeka and Bonomi2024) is also a probe-based approach, but adds a metadynamics enhanced-sampling technique to explore wider conformational changes of the input RNA molecule.

Machine learning-based

Machine learning-based methods for binding site detection in nucleic acids have emerged very recently. RNAsite (Su et al., Reference Su, Peng and Yang2021) calculates sequential features (e.g., conservation from MSA) and structural features (e.g., topological properties, solvent accessibility, and Laplacian norm) for each nucleotide and passes them into an RF classifier to distinguish between binding and non-binding nucleotides. Similarly, DrugPred_RNA (Rekand and Brenk, Reference Rekand and Brenk2021) calculates a set of simple structural descriptors such as size, shape, and polarity for a pocket and uses an XGBoost model (Chen and Guestrin, Reference Chen and Guestrin2016) to classify it as druggable or non-druggable. As descriptors are constructed in a macromolecule type-agnostic way, the model is first pre-trained on a protein dataset and then fine-tuned for binding sites in RNAs.

Deep learning-based

RLBind (Wang, Zhou, et al., Reference Wang2023b) calculates local and global sequential features (e.g., nucleotide types and evolutionary conservation) and structural features (e.g., network topological properties, biochemical properties, and ASAs), retrieves a window of 11 nucleotides for each position, and feeds it into a 1D CNN that classifies the position as binding/non-binding. RNet (Möller et al., Reference Möller2022) utilizes an approach similar to DeepSite for predicting binding sites in proteins (Jiménez et al., Reference Jiménez2017). It represents a macromolecule structure as an $ 80\times 80\times 80{\mathring{A}}^3 $ 3D voxel grid with eight channels representing different atom types: carbon, nitrogen, oxygen, phosphorus, sulfur, fluorine, bromine, and iodine. The method passes this grid as input into a 3D CNN model, predicting ligandability scores for voxels of size $ 4\times 4\times 4{\mathring{A}}^3 $ . Binding sites are retrieved by clustering predicted ligandable voxels. The authors pre-trained the model on protein binding sites and fine-tuned it to RNAs. MultiModRLBP (Wang et al., Reference Wang2024) uses a relational GCN to obtain features from a nucleotide structure graph and a pre-trained language model (RNABert (Kalicki and Haritaoglu, Reference Kalicki and Haritaoglu2022)) to get embeddings from an RNA sequence. The model concatenates these structural and sequential features and feeds the resulting vector into a small neural network of fully connected layers to obtain a prediction for each nucleotide. BiteNet N (Kozlovskii and Popov, Reference Kozlovskii and Popov2021b) predicts binding site centers on both RNA and DNA macromolecules. To train the model, the authors composed the largest dataset of $ \sim $ 2000 nucleic acid–small molecule structures. First, the method converts an input nucleic acid macromolecule structure into a voxel-based representation. Then, a 3D CNN model takes this grid as input and produces a set of binding site centers and coordinates, along with a binding score for each nucleotide.

Benchmarks

One of the most widely used benchmarks for RNA-ligand binding site detection methods is the TE18 test set from RNAsite (Su et al., Reference Su, Peng and Yang2021). Another benchmark is RB19 from RBind (Wang, Jian, et al., Reference Wang2018b). Note that, typically, methods use only a subset of these test sets to avoid sharing similar complexes with the training sets. Most recently, the authors of SHAMAN (Panei et al., Reference Panei, Gkeka and Bonomi2024) created a test set based on seven RNA complexes: riboswitches (FMN, THF, TPP, and dG) and viral RNAs (HIV-1 TAR, HCV-IRES-IIa, and IAV). They also introduced different strategies to evaluate the methods’ performance based on the holo or apo structures of these complexes. Supplementary Tables S11S16 list the performance of RNA-small molecule binding site detection methods.

Protein–ion binding site prediction

Ions are crucial for various physiological processes, such as enzymatic function, signal transduction, and muscle contraction, through their interactions with proteins (Al-Fartusie and Mohssan, Reference Al-Fartusie and Mohssan2017). Ions can bind to protein-active sites (Andreini et al., Reference Andreini2008), stabilize or trigger conformational changes in protein structures (Dudev and Lim, Reference Dudev and Lim2014; Jernigan et al., Reference Jernigan, Raghunathan and Bahar1994), regulate the activity of DNA/RNA polymerases (De Baaij et al., Reference De Baaij, Hoenderop and Bindels2015), or affect the concentration-dependent aggregation rate of proteins (Poulson et al., Reference Poulson2020). In addition, ions can act as allosteric modulators. For instance, sodium ions modulate G protein-coupled receptors (Katritch et al., Reference Katritch2014), while in calcium-sensing receptors (CaSR), $ {\mathrm{Ca}}^{2+} $ , and $ {\mathrm{Mg}}^{2+} $ serve as activators, $ {\mathrm{Cl}}^{-} $ acts as a positive allosteric modulator, and $ {\mathrm{SO}}_4^{2-} $ / $ {\mathrm{PO}}_4^{3-} $ act as negative modulators (Liu et al., Reference Liu2020a). Chloride ions ( $ {\mathrm{Cl}}^{-} $ ) also modulate mGluRs (metabotropic glutamate receptors) (Tora et al., Reference Tora2015), and calcium ions ( $ {\mathrm{Ca}}^{2+} $ ) influence nAChRs (nicotinic acetylcholine receptors) (Changeux, Reference Changeux2018). Therefore, understanding protein–ion interactions, particularly ion binding sites, is critical to deciphering protein function. Ion binding sites differ from protein–ligand and protein–peptide binding sites in several ways. First, the size of ion binding sites is generally smaller, as small molecules or peptides typically interact with more residues on the protein surface. Furthermore, ion-binding sites are often more adaptable than those of ligands (Chakrabarti, Reference Chakrabarti1993). Another distinction is that many ions require specific coordination geometries with protein atoms. For example, $ {\mathrm{Zn}}^{2+} $ binding sites are typically formed by residues such as Cys, His, Asp, or Glu, and are coordinated by four or five atoms, adopting a distorted-tetrahedral or trigonal-bipyramidal geometry (Auld, Reference Auld and Maret2001). Various computational approaches have been proposed to identify ion binding sites, as summarized in Table 5.

Table 5. List of methods for prediction of protein–ion binding sites

Sequence-based

Similarly to sequence-based methods that predict protein–small molecule binding sites, almost all of the sequence-based methods for the identification of binding sites for ions utilize a machine learning-based approach (Chen et al., Reference Chen2013; Shu et al., Reference Shu, Zhou and Hovmöller2008; Lippi et al., Reference Lippi2008; Passerini et al., Reference Passerini, Lippi and Frasconi2011; Ferrè and Clote, Reference Ferrè and Clote2006; Passerini et al., Reference Passerini2007; Haberal and Oğul, Reference Haberal and Oğul2017, Reference Haberal and Oğul2019; Qiao and Xie, Reference Qiao and Xie2019; Yu et al., Reference Yu2013, Reference Yu2015; Li et al., Reference Li, Pi and Chen2019a; Li et al., Reference Li2019b; Yan et al., Reference Yan2019; Jiang et al., Reference Jiang2016; Ding et al., Reference Ding, Tang and Guo2017; Srivastava and Kumar, Reference Srivastava and Kumar2018; Zhao et al., Reference Zhao, Xu and Zhao2019; Essien et al., Reference Essien, Wang and Xu2019; Sun et al., Reference Sun2022). First, they move a sliding window along the input sequence and calculate sequential features for each position. Features can include: evolutionary information such as position-specific scoring matrix (PSSM) or conservation score, predicted secondary structure, and predicted solvent accessibility of residues. Then, these features are fed into an SVM, RF, AdaBoost, or a simple NN. ZincExplorer (Chen et al., Reference Chen2013) combines a machine learning approach with a templates-based search of known binders to identify Zn-binding sites. IBayes_Zinc (Li et al., Reference Li, Pi and Chen2019a) uses previously described sequence descriptors and predictions from other methods (ZincExplorer (Chen et al., Reference Chen2013), ZincFinder (Passerini et al., Reference Passerini2007), and ZincPred (Shu et al., Reference Shu, Zhou and Hovmöller2008)) as input into a Bayesian algorithm to predict Zn sites. MetalPredator (Valasatava et al., Reference Valasatava2016) searches through a database of Pfam domains for Fe-S clustering binding and metal binding fragments from MetalPDB (Andreini et al., Reference Andreini2012). ZINCCLUSTER (Ajitha et al., Reference Ajitha2018) first creates a database of all monopeptides, dipeptides, and tripeptides and assigns a Z-score for each of them to be Zn-binding based on a dataset. Then, it screens an input sequence with pentapeptides and retrieves a Z-score from the database for two central dipeptides and three tripeptides. The method considers this fragment to be Zn-binding if the average Z-score of dipeptides and tripeptides is higher than zero. With advancements in deep learning, transformer-based models have been developed for ion binding site prediction. IonPred (Essien et al., Reference Essien2023) employs a transformer architecture to predict ion binding sites directly from protein sequences. M-Ionic (Shenoy et al., Reference Shenoy2024) leverages residue embeddings generated by the pre-trained protein language model ESM-2 (Lin et al., Reference Lin2023) to identify binding sites for various ions. Similarly, LMetalSite (Yuan et al., Reference Yuan2022b) utilizes residue embeddings from ProtTrans (Elnaggar et al., Reference Elnaggar2021) for the prediction of binding sites specific to $ {\mathrm{Zn}}^{2+} $ , $ {\mathrm{Ca}}^{2+} $ , $ {\mathrm{Mg}}^{2+} $ , and $ {\mathrm{Mn}}^{2+} $ .

Template-based

Many methods aim to find fragments of an input structure that are present in the template database of known ion-binding sites. MIB (Lin et al., Reference Lin2016) and (Lu et al., Reference Lu2012) use a fragment transformation method to search for parts of an input protein that are present in a database of binding residue templates for multiple ion types. For this, they split residues in the input structure and the template into residue triplets, measured triplet pair similarity and performed clustering of triplets similar to binding ones to get the final predictions. FindSite-metal (Brylinski and Skolnick, Reference Brylinski and Skolnick2011) utilizes TM-align (Zhang and Skolnick, Reference Zhang and Skolnick2005; Pandit and Skolnick, Reference Pandit and Skolnick2008) to align template fragments onto the input structure, clusters the obtained alignments, and outputs residue binding scores as the fraction of templates including corresponding positions. TEMSP (Zhao et al., Reference Zhao2011) creates a database of Zn-binding templates from all pairs of residues interacting with this ion. Then, for an input protein, it screens all residue pairs and detects the ones present in the template library. After that, matched pairs are combined into ‘pairs-of-pairs’, which are further filtered using predefined geometrical thresholds to get the final predictions. In Garg and Pal (Reference Garg and Pal2021), the authors used a geometric hashing technique to match query structures with templates of binding sites for different ion types. In Schymkowitz et al. (Reference Schymkowitz2005b), the method creates a database of canonical positions of water molecules or ions with respect to protein atom triads. Then, it screens the surface of the protein with these triads, clusters favorable points, and performs optimization of positions using the empirical force field. GASS-Metal (Paiva et al., Reference Paiva2022) uses a genetic algorithm for the effective search of structural patterns similar to ion binding sites from a curated database of templates.

Machine learning-based

Apart from the sequence-based machine learning approaches, most of the structure-based machine learning methods (Sodhi et al., Reference Sodhi2004; Bordner, Reference Bordner2008; Zheng et al., Reference Zheng2012; Ireland and Martin, Reference Ireland and Martin2021; Song and Jiang, Reference Song and Jiang2023) calculate sequential and structural features for each residue and feed them into an SVM, RF, or neural network classifier. For example, FEATURE (Ebert and Altman, Reference Ebert and Altman2008) constructs concentric radial shells for the atomic environments, calculates physicochemical features inside each of them, and uses Bayesian learning to differentiate whether an environment corresponds to a Zn-binding site or not. IonCom (Hu et al., Reference Hu2016) combines predictions from the sequence-based approach IonSeq and other tools for binding site prediction: COFACTOR (Roy et al., Reference Roy, Yang and Zhang2012), TM-SITE, S-SITE, and COACH (Yang et al., Reference Yang, Roy and Zhang2013), and trains a classifier on top of them. PinMyMetal (Zheng et al., Reference Zheng2024) uses geometrical features, including residue properties, interatomic distances, bond angles, and atomic types as input into the ensemble ML model predicting $ {\mathrm{Zn}}^{2+} $ binding sites.

Deep learning-based

GraphBind (Xia et al., Reference Xia2021) is a GNN-based model that predicts binding sites for $ {\mathrm{Ca}}^{2+} $ , $ {\mathrm{Mn}}^{2+} $ , and $ {\mathrm{Mg}}^{2+} $ . DeepProSite (Fang et al., Reference Fang2023), as mentioned in Section Deep learning-based, uses ESMFold (Rives et al., Reference Rives2021), ProtTrans (Elnaggar et al., Reference Elnaggar2021), and a GNN for the prediction of different types of binding sites, including those for $ {\mathrm{Ca}}^{2+} $ , $ {\mathrm{Mn}}^{2+} $ , and $ {\mathrm{Mg}}^{2+} $ . In Gamouh et al. (Reference Gamouh, Hoksza and Novotny2023), the authors used embeddings from ProtTrans (Elnaggar et al., Reference Elnaggar2021) as features for the graph nodes and used GNN to predict binding sites for nucleotides and ions: $ {\mathrm{Ca}}^{2+} $ , $ {\mathrm{Mg}}^{2+} $ , $ {\mathrm{Mn}}^{2+} $ , $ {\mathrm{Fe}}^{3+} $ , and $ {\mathrm{Zn}}^{2+} $ . DELIA (Xia et al., Reference Xia, Pan and Shen2020) first constructs a feature vector as the combination of outputs from two other models: (i) BiLSTM which takes as the input the sequence-based features; and (ii) 2D CNN ResNet model, which takes as the input the distance matrix. The constructed feature vectors are used as the input to the next fully connected layer, which outputs the probability of each residue being binding or non-binding. Metal3D (Dürr et al., Reference Dürr, Levy and Rothlisberger2023) employs a 3D CNN to predict the probability density of $ {\mathrm{Zn}}^{2+} $ binding across the protein structure. MoM (Laveglia et al., Reference Laveglia2023) utilizes a GNN to classify local protein environments composed of Cys, His, Asp, and Glu residues, determining whether these environments are likely to bind $ {\mathrm{Zn}}^{2+} $ . BindWeb (Xia et al., Reference Xia2022) is a consensus method combining predictions from GraphBind (Xia et al., Reference Xia2021) and DELIA (Xia et al., Reference Xia, Pan and Shen2020) models.

Other

There are several geometric methods that search positions with surrounding atoms whose geometry resembles the ion coordination shell. GRE4Zn (Liu et al., Reference Liu2014) utilizes the fact that most known Zn-binding sites comprise sets of four or three residues with distinctly specific geometries. GaudiMM Metals (Sciortino et al., Reference Sciortino2019) retrieves information about acceptable coordination shell geometries for a set of ions and implements them as an additional objective for optimization with ion presence in the GaudiMM platform (Rodrı́guez-Guerra Pedregal et al., Reference Rodrı́guez-Guerra Pedregal2017). BioMetAll (Sánchez-Aparicio et al., Reference Sánchez-Aparicio2020) constructs a grid of metal probes around a protein and checks each grid position to see if the amino acid environment matches geometric constraints determined from statistics in a dataset of protein structures. The method obtains final predictions from the clustering of relevant points. Also, there are methods that utilize an energetic approach. For example, BION (Shashikala et al., Reference Shashikala2021) calculates electrostatic potential maps with a gaussian-smooth dielectric function term to predict the positions of non-specifically surface-bound ions.

It is important to note that many of the ion binding site identification methods consider only Cys, His, Glu, and Asp residues (Chen et al., Reference Chen2013; Shu et al., Reference Shu, Zhou and Hovmöller2008; Passerini et al., Reference Passerini2007), as these four amino acids are involved in the coordination shell of a bound ion in many cases. Moreover, MetalDetector (Lippi et al., Reference Lippi2008; Passerini et al., Reference Passerini, Lippi and Frasconi2011) and DeepMBS (Haberal and Oğul, Reference Haberal and Oğul2017, Reference Haberal and Oğul2019) operate only with His and Cys, and DiANNA (Ferrè and Clote, Reference Ferrè and Clote2006) work solely with Cys residues. On the other hand, many methods have been developed to predict binding regions for specific ions. For example, multiple approaches aim to predict Zn-binding regions (Chen et al., Reference Chen2013; Shu et al., Reference Shu, Zhou and Hovmöller2008; Passerini et al., Reference Passerini2007; Haberal and Oğul, Reference Haberal and Oğul2017, Reference Haberal and Oğul2019; Li et al., Reference Li, Pi and Chen2019a; Li et al., Reference Li2019b; Yan et al., Reference Yan2019; Ajitha et al., Reference Ajitha2018).

It is worth noting that for some ions (e.g., $ {\mathrm{Ca}}^{2+} $ , $ {\mathrm{Mg}}^{2+} $ , $ {\mathrm{Na}}^{+} $ , $ \mathrm{and}\hskip0.5em {\mathrm{K}}^{+} $ ), the performance metrics are much lower compared to others, as can be seen in Supplementary Table S19. As pointed out in Lu et al. (Reference Lu2012) and Qiao and Xie (Reference Qiao and Xie2019), this can be caused by the higher variability of these binding sites in terms of amino acid composition and structure. Indeed, in Qiao and Xie (Reference Qiao and Xie2019), the authors calculate the frequency difference index, defined as the average difference in the ratio of binding and non-binding residues of each type among the 20 amino acid types, and observed that the index values are much lower for $ {\mathrm{Ca}}^{2+} $ , $ {\mathrm{Mg}}^{2+} $ , $ {\mathrm{Na}}^{+} $ , and $ {\mathrm{K}}^{+} $ compared to other ions.

Benchmarks

We observed that different methods use various benchmarks for evaluation; here, we list benchmarks that were used by several methods. The Passerini dataset (Passerini et al., Reference Passerini2006) is a dataset containing 2,727 sequences with 687 protein chains bound to a metal atom. There are four methods that used it as a training or validation set for the prediction of Zn-binding sites (ZincFinder (Passerini et al., Reference Passerini2007), ZincPred (Shu et al., Reference Shu, Zhou and Hovmöller2008), ZincExplorer (Chen et al., Reference Chen2013), DeepMBS (Haberal and Oğul, Reference Haberal and Oğul2017)). However, note that these methods calculated different metrics on different sets of residues (e.g., Cys and His or Cys, His, Glu, and Asp). The Zhao dataset (Zhao et al., Reference Zhao2011) is a dataset used for training and validation of a template-based method TEMSP, consisting of $ \sim $ 600 protein targets with bound Zn ions. Although many methods use this dataset as an independent test set, some methods retrieved only a subset from it. SSWPNN (Li et al., Reference Li2019b) provides the most complete comparison of methods on this dataset (see Supplementary Table S21). Furthermore, for the validation of SSWPNN, the authors also collected a second independent test set from PDB consisting of 213 protein chains with 1,017 Zn-binding sites, and compared SSWPNN with five other approaches for the prediction of Zn binding sites on the Zhao and SSWPNN datasets (see Supplementary Tables S21 and S22). ZincBindDB (Ireland and Martin, Reference Ireland and Martin2019) is the largest database of Zn binding sites (about 35,000 binding sites from about 16,000 structures), that was used for training and validation of the ZincBindPredict method (Ireland and Martin, Reference Ireland and Martin2021) (however, it was not used by the other methods). Note that a newer version of the database contains more samples ( $ \sim $ 40,000 binding sites from $ \sim $ 16,000 structures); so one may expect improved performance for newer methods. The BION dataset (Petukh et al., Reference Petukh, Kimmet and Alexov2013) contains binding sites for $ {\mathrm{Ca}}^{2+} $ , $ {\mathrm{Zn}}^{2+} $ , $ {\mathrm{Cl}}^{-} $ , and $ {\mathrm{Mg}}^{2+} $ ions from 446 protein structures. In Shashikala et al. (Reference Shashikala2021), the authors used this dataset to compare the performances of BION (Petukh et al., Reference Petukh, Kimmet and Alexov2013) and BION-2 (Shashikala et al., Reference Shashikala2021) methods with forcefield-based tools from VMD (Humphrey et al., Reference Humphrey, Dalke and Schulten1996) and Fold-X (Schymkowitz et al., Reference Schymkowitz2005a). In Hu et al. (Reference Hu2016), the authors created a large dataset of 2,100 protein structures in complex with 3,075 ions ( $ {\mathrm{Zn}}^{2+} $ , $ {\mathrm{Cu}}^{2+} $ , $ {\mathrm{Fe}}^{2+} $ , $ {\mathrm{Fe}}^{3+} $ , $ {\mathrm{Ca}}^{2+} $ , $ {\mathrm{Mg}}^{2+} $ , $ {\mathrm{Mn}}^{2+} $ , $ {\mathrm{Na}}^{+} $ , $ {\mathrm{K}}^{+} $ , $ {\mathrm{CO}}_3^{2-} $ , $ {\mathrm{NO}}_2^{-} $ , $ {\mathrm{SO}}_4^{2-} $ , $ \mathrm{and}\hskip0.5em {\mathrm{PO}}_4^{3-} $ ) retrieved from the BioLip database (Yang et al., Reference Yang, Roy and Zhang2012). The authors used it for 5-fold cross-validation of IonSeq and IonCom methods, and there are available scores for several methods on this benchmark (see Supplementary Table S18). In MIonSite (Qiao and Xie, Reference Qiao and Xie2019), the authors created a large dataset of 7,676 sequences for training and 274 sequences for an independent test set. These sets include ions of multiple types: $ {\mathrm{Zn}}^{2+} $ , $ {\mathrm{Ca}}^{2+} $ , $ {\mathrm{Mg}}^{2+} $ , $ {\mathrm{Mn}}^{2+} $ , $ {\mathrm{Fe}}^{3+} $ , $ {\mathrm{Cu}}^{2+} $ , $ {\mathrm{Fe}}^{2+} $ , $ {\mathrm{Co}}^{2+} $ , $ {\mathrm{Na}}^{+} $ , $ {\mathrm{K}}^{+} $ , $ {\mathrm{Cd}}^{2+} $ , $ \mathrm{and}\hskip0.5em {\mathrm{Ni}}^{2+} $ . MIonSite was compared with other methods on their test set (see Supplementary Table S19). The authors also created a small dataset (BTD) of 10 proteins with metal ion-binding sites and 10 proteins without metal ion-binding sites for additional comparison with other methods (see Supplementary Table S20). TargetS (Yu et al., Reference Yu2013) used the BioLip database (Yang et al., Reference Yang, Roy and Zhang2012) to assemble training and validation sets with metal ion binding sites ( $ {\mathrm{Ca}}^{2+} $ , $ {\mathrm{Zn}}^{2+} $ , $ {\mathrm{Mg}}^{2+} $ , $ {\mathrm{Mn}}^{2+} $ , $ \mathrm{and}\hskip0.5em {\mathrm{Fe}}^{3+} $ ) and nucleotides with 3,779 and 642 ion-bound protein sequences, respectively. The authors used an independent test set to compare TargetS with ligand-specific predictors and an alignment-based predictor (see Supplementary Table S23). Garg and Pal (Garg and Pal, Reference Garg and Pal2021) assembled datasets for five metal ions ( $ {\mathrm{Cu}}^{2+} $ , $ {\mathrm{Fe}}^{3+} $ , $ {\mathrm{Ca}}^{2+} $ , $ {\mathrm{Mg}}^{2+} $ , $ \mathrm{and}\hskip0.5em {\mathrm{Zn}}^{2+} $ ) and split them into training and testing sets with 1,079 and 268 structures in total, respectively. The authors compared their method with IonCom (Hu et al., Reference Hu2016) and MIB (Lin et al., Reference Lin2016) (see Supplementary Table S24). In Yan et al. (Reference Yan2019), the authors prepared the zn1436 dataset of proteins with bound zinc ions for a comparison of the ZnMachine method with ZincExplorer (Chen et al., Reference Chen2013) (see Supplementary Table S25). In Yuan et al. (Reference Yuan2022b), the authors compiled a test set from the BioLip database (Yang, Reference Yang, Roy and Zhang2012), consisting of 211, 183, 235, and 57 protein chains bound to $ {\mathrm{Zn}}^{2+} $ , $ {\mathrm{Ca}}^{2+} $ , $ {\mathrm{Mg}}^{2+} $ , and $ {\mathrm{Mn}}^{2+} $ ions, respectively (see Supplementary Table S27). Similarly, the authors of M-Ionic (Shenoy et al., Reference Shenoy2024) constructed an independent test set from BioLip, but reported performance metrics only for LMetalSite (Yuan et al., Reference Yuan2022b) and M-Ionic (Shenoy et al., Reference Shenoy2024) on this dataset (see Supplementary Table S28). Among the benchmarks used by a single method, one may notice the BioMetAll test set (Sánchez-Aparicio et al., Reference Sánchez-Aparicio2020), which consists of 53 crystallographic structures containing the two-histidine one-carboxylate motif (FTM). This is an interesting benchmark since its structures vary in size, and this motif may bind multiple types of metal ions ( $ {\mathrm{Cd}}^{2+} $ , $ {\mathrm{Co}}^{2+} $ , $ {\mathrm{Cu}}^{2+} $ , $ {\mathrm{Fe}}^{3+} $ , $ {\mathrm{Hg}}^{2+} $ , $ {\mathrm{Mg}}^{2+} $ , $ {\mathrm{Mn}}^{2+} $ , $ {\mathrm{Ni}}^{2+} $ , $ {\mathrm{Ru}}^{3+} $ , $ \mathrm{and}\hskip0.5em {\mathrm{Zn}}^{2+} $ ). Note that almost all methods operate with residue-based scores as their performance metric, and only GaudiMM Metals (Sciortino et al., Reference Sciortino2019) and BION (Shashikala et al., Reference Shashikala2021) used distance-based scores, which are likely more representative of the ion binding site prediction problem.

Despite the large variety of methods and benchmarks (see Table 5), one can see from Supplementary Tables S18 and S19 that MIonSite (Qiao and Xie, Reference Qiao and Xie2019) and IonCom (Hu et al., Reference Hu2016) demonstrate better performance for different ions. Interestingly, MIonSite is a sequence-based method, and IonCom is a structure-based method; therefore, it would be interesting to see if a combined approach shows even better results. As for the Zn-specific predictors, Supplementary Table S22 shows that SSWPNN (Li et al., Reference Li2019b) outperforms other methods.

Challenges

The composition of high-quality and diverse labeled datasets and benchmarks is one of the biggest challenges in the binding site detection problem. As one can see from the Benchmarks subsections, typically, there are no unified training-validation sets and test benchmarks to perform a rigorous comparison of the developed methods. Moreover, the existing training and test splits often contain data leakage (let alone structural artifacts), resulting in likely overly optimistic performance metrics, that should not be compared between the different methods. Unfortunately, despite the exponential growth of available experimental structures in PDB, in some cases, the test benchmarks are too small to make statistically solid conclusions. For example, the commonly used RNA-small molecule binding site benchmark, TE18, contains just 18 structures (see Section Nucleic acid–small molecule binding sites Benchmarks). Related to this, another challenge to consider, especially with respect to the ML and DL methods, is overfitting. Overfitting is a common problem in machine learning, where models perform well on training data but fail to generalize to the unseen cases. Deficient training sets or incorrect training-validation-test splitting can also result in models showing artificially high score values (Kapoor and Narayanan, Reference Kapoor and Narayanan2023). Thus, to overcome these challenges, not only high-quality datasets are required but also unified training-validation-test splits. This will also help to make the comparison of different methods more rigorous. However, other hidden biases may remain. For example, using pre-trained language models (LMs) may lead to data leakage, as the LM model itself might have seen data similar to the training set.

One rather technical but important thing that also prevents a rigorous comparison of binding site detection methods is the use of different performance metrics for their evaluation. First of all, many papers present accuracy, precision, recall, or ROC AUC metrics, which may be misleading. Indeed, precision or recall metrics are high when the model outputs either the lowest or highest number of positive predictions, respectively, and accuracy or ROC AUC tends to be 1 when there is a high imbalance in binary labels. MCC is a much more suitable metric for binding site detection; however, it still depends on the choice of the threshold for determining binary labels. On the other hand, AP or AURPC, which is the area under the precision-recall curve, is much more convenient for the classification of binding residues, as it takes into account the ranking of predictions and does not depend on class imbalance. Also, instead of calculating precision on the top-N predictions for each structure (see DCC and DCA metrics in Supplementary Section Metrics), one can use AP for assessing model performance using a distance-based criterion as well (Kozlovskii and Popov, Reference Kozlovskii and Popov2020), following this idea from object detection in computer vision (Everingham et al., Reference Everingham2010). More reliable performance metrics can be roughly divided into three categories: distance-, volume-, and residue-based. Distance-based metrics define a prediction as successful if the distance from the predicted binding site center to the true center or any atom of the binding site or ligand is lower than a threshold value, which is usually set to 4Å. Volume-based metrics calculate the overlap between the predicted and true binding site cavities. Finally, residue-based metrics rely on binary classification metrics calculated from residue scores. There is no one-size-fits-all solution, however. For example, residue-based scores may be misleading for protein–ion binding sites. Indeed, the number of interacting residues is small; thus, the impact of a single residue on the metric is high. Furthermore, the definition of interacting residues varies too, resulting in different metric values for the same predictions but with a slightly different set of true labels. Similarly, a residue-based metric may be misleading for nucleic acid–small molecule binding site prediction, though with the opposite reasoning. In this case, the number of nucleotides in the structure is typically small; thus, the binding site covers a larger portion of nucleotide residues. As a result, residue-based metrics may become insensitive to very different predictions. In contrast, distance- and volume-based metrics have been shown to be informative enough for protein–small molecule binding site predictions. Therefore, distance-based metrics would be more robust for protein–ion and nucleic acid–small molecule binding site detection problems. However, as one can see from Sections Nucleic acid–small molecule binding sites Benchmarks and Protein–ion binding site prediction Benchmarks, most of the methods rely on residue-based metrics. On the other hand, residue-based metrics can be more suitable for protein–peptide binding site detection methods compared to distance- and volume-based metrics, because of the large size of protein–peptide binding site interfaces.

Another challenge is the interpretability of ML and DL-based binding site detection models. While these methods could achieve superior accuracy compared to classical approaches, their predictions often lack clear mechanistic explanations, making it difficult to extract meaningful insights about the underlying molecular interactions (Murdoch et al., Reference Murdoch2019; Vecchietti et al., Reference Vecchietti2024). This issue can be particularly relevant when understanding why a model identifies a particular region as a potential binding site is essential for hypothesis-driven drug design. Unlike physics-based or first-principle methods, which rely on well-understood physical and chemical principles, DL models operate as complex, non-linear transformations of input data, obscuring the contributions of individual molecular features. This ambiguity also hampers debugging models, detecting biases in datasets, and ensuring reliable generalization across diverse molecular structures. Recent advances in explainable AI (XAI) (Jiménez-Luna et al., Reference Jiménez-Luna, Grisoni and Schneider2020; Bhatt et al., Reference Bhatt, Koes and Durrant2024), such as feature attribution techniques (e.g., SHAP (Lundberg and Lee, Reference Lundberg, Lee, Guyon, Von Luxburg, Bengio, Wallach, Fergus, Vishwanathan and Garnett2017), LRP (Montavon et al., Reference Montavon2019), Grad-CAM (Selvaraju et al., Reference Selvaraju2017)) and attention mechanisms in transformer-based models (Wiegreffe and Pinter, Reference Wiegreffe and Pinter2019), have been proposed to increase interpretability, but their application to binding site prediction remains limited. Incorporating interpretable ML techniques into binding site detection could improve trust in DL-based predictions and enhance their practical usability in drug discovery pipelines.

In drug discovery, one of the challenges is to assess the ‘druggability’ or ‘ligandability’ of the detected binding sites. Currently, there are no strict criteria for the ‘druggable’ binding sites. Similarly to the characterization of drug-like molecules using the rule of 5, Ghose filter, or other heuristics, one can compose such knowledge-based criteria for the binding sites based on their properties. While an exhaustive review of binding site characterization methods is beyond the scope of this article, it is noteworthy that several computational tools have been developed to analyze specific properties of binding sites. These tools assess attributes such as volume, surface area, and flexibility, and often identify sub-pockets within larger pockets (Durrant et al., Reference Durrant2014; Guerra et al., Reference Guerra2021), typically, employing approaches similar to those discussed in Section Protein–small molecule binding sites Geometric. Moreover, certain geometric binding site prediction methods inherently provide volume estimations of binding sites (Kawabata and Go, Reference Kawabata and Go2007; Capra et al., Reference Capra2009; Le Guilloux et al., Reference Le Guilloux, Schmidtke and Tuffery2009; Zhu and Pisabarro, Reference Zhu and Pisabarro2011), and some methods analyze pockets throughout molecular dynamics (MD) trajectories, offering dynamic insights into binding site properties (Craig et al., Reference Craig2011; Schmidtke et al., Reference Schmidtke2011; Paramo et al., Reference Paramo2014; Laurent et al., Reference Laurent2015; Wagner et al., Reference Wagner2017; Chen et al., Reference Chen2019; Lv and Cao, Reference Lv and Cao2024). Other tools, like MOLE (Pravda et al., Reference Pravda2018), CAVER (Stourac et al., Reference Stourac2019), and others (Yaffe et al., Reference Yaffe2008; Lee and Helms, Reference Lee and Helms2012) aim at the characterization of protein tunnels, channels, and pores.

Last but not least challenge is the prospective validation of the developed methods. Given the aforementioned challenges that can lead to over-optimistic performance on the retrospective benchmarks, the real-world application is of crucial importance. However, such case studies are quite rare (Popov et al., Reference Popov2024; Naz et al., Reference Naz2015) and absent for most of the developed methods. In this regard, community-driven challenges, such as CASP (https://predictioncenter.org) and CACHE (https://cache-challenge.org ), may comprise targets with previously unpublished binding sites and, thus, provide an opportunity to demonstrate the predictive power of the developed methods.

Trends and future directions

It is no wonder that machine learning-based approaches are gradually displacing the first-principle methods, and while older research focuses more on searching for the most powerful features, newer research is more focused on exploring various neural network architectures. Moreover, with the advances in large language models, it has become common to utilize embeddings produced by, for example, protein language models, as the feature vectors for the downstream task of the binding site detection. While this idea seems promising, extensive exploration of this research direction is computationally expensive, requiring significant hardware resources and time, which can limit accessibility for some research groups.

When applying or testing binding site detection methods, an important question to address is the flexibility of the target. Naturally, one expects that a method should detect a binding site in the holo conformation of the target. But in practice, one needs to discover novel binding sites given the unbound conformation. At what point in the imaginary trajectory between the unbound and bound conformations should a method detect the binding site? The answer to this question likely depends on the considered set of ligands for a particular target, such that the method should detect a binding site in the structure, if the corresponding conformation is within a certain vicinity of the bound conformation for at least one of the ligands. The vicinity can be simply defined as all conformations within a given RMSD threshold relative to the bound conformations. Constructing such a benchmark of conformational ensembles would be a valuable step forward for the development of robust binding site detection approaches. Currently, one typically addresses the flexibility issue by generating multiple conformations of the target using molecular dynamics or another method and applying binding site detection to them (Kozlovskii and Popov, Reference Kozlovskii and Popov2020; Martinez-Rosell et al., Reference Martinez-Rosell2020; Meller et al., Reference Meller2023; Panei et al., Reference Panei, Gkeka and Bonomi2024). The development of spatiotemporal methods to analyze target binding sites and their dynamics is a valuable direction for future research.

There are other types of binding sites besides those covered in this review. For example, specific models have been developed for protein-nucleotide binding sites (Chauhan et al., Reference Chauhan, Mishra and Raghava2009; Chen et al., Reference Chen, Mizianty and Kurgan2012; Kusuma et al., Reference Kusuma2019), carbohydrate binding sites (Canner et al., Reference Canner, Shanker and Gray2023), vitamin binding sites (Panwar et al., Reference Panwar, Gupta and Raghava2013), catalytic sites (Dou et al., Reference Dou2012), as well as water positions (Zaucha et al., Reference Zaucha2020; Park and Seok, Reference Park and Seok2022). As for RNA targets, there are methods to predict RNA-ion binding sites, including MetalionRNA (Philips et al., Reference Philips2012), MgNet (Zhou and Chen, Reference Zhou and Chen2022), Metal3DRNA (Zhao et al., Reference Zhao2023), as well as machine learning methods to predict RNA-protein binding sites using only sequence data (Choi and Han, Reference Choi and Han2013; Panwar and Raghava, Reference Panwar and Raghava2015; Tuvshinjargal et al., Reference Tuvshinjargal2016; Choi et al., Reference Choi2017; Zhan et al., Reference Zhan2018; Pan et al., Reference Pan2020; Tahir et al., Reference Tahir2021; Zhao et al., Reference Zhao, Zhang and Du2022), sequential and secondary structure (Uhl et al., Reference Uhl2019), or sequential and tertiary data (Luo et al., Reference Luo2017). Some of the sequence-based methods rely on large databases with experimental data on RNA-protein binding, such as RNAcompete (Ray et al., Reference Ray2009), CLIP-seq, and RIP-seq (Ray et al., Reference Ray2013). Notably, there are approaches to predict DNA binding sites (e.g., DeepBind (Alipanahi et al., Reference Alipanahi2015) and DeepSTF (Ding et al., Reference Ding2023)), trained on datasets from protein binding microarrays (PBMs) (Mukherjee et al., Reference Mukherjee2004), ChIP-seq (Kharchenko et al., Reference Kharchenko, Tolstorukov and Park2008), or HT-SELEX (Jolma et al., Reference Jolma2010). We would like to separately note that protein-covalent ligand binding sites constitute a special case of protein–small molecule binding sites. Covalent ligands may be useful as a therapeutic modality in various diseases; hence, the prediction of this type of binding site is relevant in covalent drug discovery (Boike et al., Reference Boike, Henning and Nomura2022). A ligand can form a covalent bond with target residues (commonly Cys, Ser, and Lys, but there are other cases as well) upon binding, which imposes strict constraints on the binding site detection problem. There are several databases of covalent agents, including CovalentInDB (Du et al., Reference Du2021) and CovPDB (Gao et al., Reference Gao2022), and there are methods for predicting the ability of cysteines to form a covalent bond with ligands ( Zhang et al., Reference Zhang2016, Reference Zhang, Pei and Lai2017; Zhao et al., Reference Zhao2017; Du et al., Reference Du2022; Gao and Günther, Reference Gao and Günther2023). Other examples include methods to predict macromolecular binding sites, such as protein-nucleic acids ( Hendrix et al., Reference Hendrix2021; Wei et al., Reference Wei2022; Yuan et al., Reference Yuan2022a; Liu and Tian, Reference Liu and Tian2023; Roche et al., Reference Roche2023; Song et al., Reference Song2023; Zhu and Yu, Reference Zhu and Yu2023) or protein–protein (Fout et al., Reference Fout2017; Gainza et al., Reference Gainza2020; Dai and Bailey-Kellogg, Reference Dai and Bailey-Kellogg2021; Renaud et al., Reference Renaud2021; Sverrisson et al., Reference Sverrisson2021; Tubiana et al., Reference Tubiana, Schneidman-Duhovny and Wolfson2022; Gao et al., Reference Gao2023; Jha et al., Reference Jha, Karmakar and Saha2023; Krapp et al., Reference Krapp2023). Most of these methods are based on approaches similar to the ones described here. Given the large variety of binding site types on one hand and the advances in multi-modal and multi-task machine learning approaches on the other hand, we expect that the next-generation binding site prediction methods will operate across different types of macromolecular structures as well as their binding counterparts. Although there is currently no strong evidence that this will improve model accuracy, one may expect that a single model trained on comprehensive datasets could have stronger generalization ability and robustness. We observed that some methods implicitly explore this idea already; for example, RNet (Liu et al., Reference Liu2024) and $ {\mathrm{BiteNet}}_{Pp} $ (Kozlovskii and Popov, Reference Kozlovskii and Popov2021a) started with protein–small molecule binding site detection models and fine-tuned them for RNA-small molecule and protein–peptide binding site models, respectively.

Finally, the discovery of novel binding sites may come from an orthogonal direction. For example, global molecular docking approaches search for the optimal configuration of two binding partners without prior knowledge of the corresponding binding site. Molecular docking methods have been developed for different types of macromolecules and ligands, including those described here in the context of the binding site detection problem. Needless to say, the molecular docking field and virtual ligand screening, in general, are also affected by the machine-learning era (Fadahunsi et al., Reference Fadahunsi2024). Moreover, breakthroughs in protein structure prediction by DeepMind (https://deepmind.google/technologies/alphafold/) have also opened new opportunities to solve binding site detection problems. Particularly, one promising direction is the development of co-folding approaches that aim to predict the 3D structure of the molecular complex starting from a 1D (sequence) or 2D (graph) representation of its subunits. The most recent examples of such approaches include AlphaFold3 (Abramson et al., Reference Abramson2024), RoseTTaFold All-Atom (Krishna et al., Reference Krishna2024), or NeuralPLexer (Qiao et al., Reference Qiao2024). Although their predictive power has yet to be assessed on independent test benchmarks to date, one may expect the rise of end-to-end approaches to solving structure prediction, binding site detection, and molecular docking problems simultaneously in the future.

List of abbreviations

ML

machine learning

DL

deep learning

RF

random forest

SVM

support vector machine

MSA

multiple sequence alignment

NN

neural network

CNN

convolutional neural network

GNN

graph neural network

GCN

graph convolutional network

MPNN

message passing neural network

GRU

gated recurrent unit

LSTM

long short-term memory

RMSD

root-mean-square deviation

MD

molecular dynamics

SS

secondary structure

ASA

accessible surface area

SASA

solvent accessible surface area

RSASA

relative solvent accessible surface area

PSSM

position specific scoring matrix

AF2

AlphaFold2

ATP

adenosine triphosphate

RNA

ribonucleic acid

DNA

deoxyribonucleic acid

PDB

Protein Data Bank

NMR

nuclear magnetic resonance

pLM

protein language model

Supplementary material

The supplementary material for this article can be found at http://doi.org/10.1017/S003358352500006X.

Acknowledgements

This study was done within the PROXIDRUGS consortium. PROXIDRUGS is as part of the initiative ‘Clusters4Future’ funded by the Federal Ministry of Education and Research BMBF (03ZU2109JE; 03ZU2109ID).

Competing interests

The authors declare no competing interests.

References

Abdin, O, et al. (2022) PepNN: a deep attention model for the identification of peptide binding sites. Communications Biology 5, 503.Google Scholar
Abramson, J, et al. (2024) Accurate structure prediction of biomolecular interactions with AlphaFold 3. Nature, 630(8016), 493500.Google Scholar
Adamczak, R, Porollo, A and Meller, J (2004) Accurate prediction of solvent accessibility using neural networks–based regression. Proteins: Structure, Function, and Bioinformatics 56, 753767.Google Scholar
Aggarwal, R, et al. (2021) DeepPocket: ligand binding site detection and segmentation using 3D convolutional neural networks. Journal of Chemical Information and Modeling 62, 50695079.Google Scholar
Ahmad, S, Gromiha, MM and Sarai, A (2003) Real value prediction of solvent accessibility from amino acid sequence. Proteins: Structure, Function, and Bioinformatics 50, 629635.Google Scholar
Ajitha, M, et al. (2018) Development of metal-active site and Zinccluster tool to predict active site pockets. Proteins: Structure, Function, and Bioinformatics 86, 322331.Google Scholar
Alberts, B (2017) Molecular Biology of the Cell. New York: Garland Science.Google Scholar
Al-Fartusie, FS and Mohssan, SN (2017) Essential trace elements and their vital roles in the human body. Indian Journal of Advanced Chemical Sciences 5, 127136.Google Scholar
Alipanahi, B, et al. (2015) Predicting the sequence specificities of DNA- and RNA-binding proteins by deep learning. Nature Biotechnology 33, 831838.Google Scholar
An, J, Totrov, M and Abagyan, R (2004) Comprehensive identification of ‘druggable’ protein ligand binding sites. Genome Informatics 15, 3141.Google Scholar
An, J, Totrov, M and Abagyan, R (2005) Pocketome via comprehensive identification and classification of ligand binding envelopes. Molecular & Cellular Proteomics 4, 752761.Google Scholar
Andreini, C, et al. (2008) Metal ions in biological catalysis: from enzyme databases to general principles. JBIC Journal of Biological Inorganic Chemistry 13, 12051218.Google Scholar
Andreini, C, et al. (2012) MetalPDB: a database of metal sites in biological macromolecular structures. Nucleic Acids Research 41, D312D319.Google Scholar
Armon, A, Graur, D and Ben-Tal, N (2001) ConSurf: an algorithmic tool for the identification of functional regions in proteins by surface mapping of phylogenetic information. Journal of Molecular Biology 307, 447463.Google Scholar
Auld, DS (2001) Zinc coordination sphere in biochemical zinc sites. In Maret, W. Zinc Biochemistry, Physiology, and Homeostasis, Dordrecht: Springer pp. 85127.Google Scholar
Baek, M, et al. (2021) Accurate prediction of protein structures and interactions using a three-track neural network. Science 373, 871876.Google Scholar
Bagley, SC and Altman, RB (1995) Characterizing the microenvironment surrounding protein sites. Protein Science 4, 622635.Google Scholar
Bagley, SC and Altman, RB (1996) Conserved features in the active site of nonhomologous serine proteases. Folding and Design 1, 371379.Google Scholar
Barker, JA and Thornton, JM (2003) An algorithm for constraint-based structural template matching: application to 3D templates with statistical analysis. Bioinformatics 19, 16441649.Google Scholar
Békés, M, Langley, DR and Crews, CM (2022) PROTAC targeted protein degraders: the past is prologue. Nature Reviews Drug Discovery 21, 181200.Google Scholar
Benet, LZ, et al. (2016) BDDCS, the Rule of 5 and drugability. Advanced Drug Delivery Reviews 101, 8998.Google Scholar
Bhatt, R, Koes, DR and Durrant, JD (2024) CENsible: interpretable insights into small-molecule binding with context explanation networks. Journal of Chemical Information and Modeling 64, 46514660.Google Scholar
Bhinge, A, et al. (2004) Accurate detection of protein: ligand binding sites using molecular dynamics simulations. Structure 12, 19891999.Google Scholar
Binkowski, TA, Naghibzadeh, S and Liang, J (2003) CASTp: computed atlas of surface topography of proteins. Nucleic Acids Research 31, 33523355.Google Scholar
Bock, ME, Garutti, C and Guerra, C (2007) Effective labeling of molecular surface points for cavity detection and location of putative binding sites. In Markstein, Peter and Xu, Ying Computational Systems Bioinformatics : (Volume 6), pp. 263274. Singapore: World Scientific.Google Scholar
Boike, L, Henning, NJ and Nomura, DK (2022) Advances in covalent drug discovery. Nature Reviews Drug Discovery 21, 881898.Google Scholar
Bondugula, R and Xu, D (2007) MUPRED: a tool for bridging the gap between template-based methods and sequence profile-based methods for protein secondary structure prediction. Proteins: Structure, Function, and Bioinformatics 66, 664670.Google Scholar
Bordner, AJ (2008) Predicting small ligand binding sites in proteins using backbone structure. Bioinformatics 24, 28652871.Google Scholar
Bradford, JR and Westhead, DR (2005) Improved prediction of protein–protein binding sites using a support vector machines approach. Bioinformatics 21, 14871494.Google Scholar
Brady, GP and Stouten, PFW (2000) Fast prediction and visualization of protein binding pockets with PASS. Journal of Computer-Aided Molecular Design 14, 383401.Google Scholar
Brandes, N, et al. (2022) ProteinBERT: a universal deep-learning model of protein sequence and function. Bioinformatics 38, 21022110.Google Scholar
Brenke, R, et al. (2009) Fragment-based identification of druggable ‘hot spots’ of proteins using Fourier domain correlation techniques. Bioinformatics 25, 621627.Google Scholar
Bron, C and Kerbosch, J (1973) Algorithm 457: finding all cliques of an undirected graph. Communications of the ACM 16, 575577.Google Scholar
Brooks, BR, et al. (1983) CHARMM: a program for macromolecular energy, minimization, and dynamics calculations. Journal of Computational Chemistry 4, 187217.Google Scholar
Broomhead, NK and Soliman, ME (2017) Can we rely on computational predictions to correctly identify ligand binding sites on novel protein drug targets? Assessment of binding site prediction methods and a protocol for validation of predicted binding sites. Cell Biochemistry and Biophysics 75, 1523.Google Scholar
Brylinski, M and Skolnick, J (2008) A threading-based method (FINDSITE) for ligand-binding site prediction and functional annotation. Proceedings of the National Academy of Sciences 105, 129134.Google Scholar
Brylinski, M and Skolnick, J (2011) FINDSITE-Metal: integrating evolutionary information and machine learning for structure-based metal-binding site prediction at the proteome level. Proteins: Structure, Function, and Bioinformatics 79, 735751.Google Scholar
Buchwald, P (2010) Small-molecule protein–protein interaction inhibitors: therapeutic potential in light of molecular size, chemical space, and ligand binding efficiency considerations. IUBMB Life 62, 724731.Google Scholar
Canner, SW, Shanker, S and Gray, JJ (2023) Structure-based neural network protein–carbohydrate interaction predictions at the residue level. Frontiers in Bioinformatics 3, 1186531.Google Scholar
Capra, JA, et al. (2009) Predicting protein ligand binding sites by combining evolutionary sequence conservation and 3D structure. PLoS Computational Biology 5, e1000585.Google Scholar
Chakrabarti, P (1993) Anion binding sites in protein structures. Journal of Molecular Biology 234, 463482.Google Scholar
Chalmers, MJ, et al. (2006) Probing protein ligand interactions by automated hydrogen/deuterium exchange mass spectrometry. Analytical Chemistry 78, 10051014.Google Scholar
Chandra, A, et al. (2023) PepCNN deep learning tool for predicting peptide binding residues in proteins using sequence, structural, and language model features. Scientific Reports 13, 20882.Google Scholar
Chang, L and Perez, A (2023) Ranking peptide binders by affinity with AlphaFold. Angewandte Chemie 135, e202213362.Google Scholar
Changeux, JP (2013) The concept of allosteric modulation: an overview. Drug Discovery Today: Technologies 10, e223e228.Google Scholar
Changeux, JP (2018) The nicotinic acetylcholine receptor: a typical ‘allosteric machine’. Philosophical Transactions of the Royal Society, B: Biological Sciences 373, 20170174.Google Scholar
Changeux, JP and Christopoulos, A (2016) Allosteric modulation as a unifying mechanism for receptor function and regulation. Cell 166, 10841102.Google Scholar
Chauhan, JS, Mishra, NK and Raghava, GPS (2009) Identification of ATP binding residues of a protein from its primary sequence. BMC Bioinformatics 10, 19.Google Scholar
Chen, K, et al. (2011) A critical comparative assessment of predictions of protein-binding sites for biologically relevant organic compounds. Structure 19, 613621.Google Scholar
Chen, K, Mizianty, MJ and Kurgan, L (2012) Prediction and analysis of nucleotide-binding residues using sequence and sequence-derived structural descriptors. Bioinformatics 28, 331341.Google Scholar
Chen, P, et al. (2015) A sequence-based dynamic ensemble learning system for protein ligand-binding site prediction. IEEE/ACM Transactions on Computational Biology and Bioinformatics 13, 901912.Google Scholar
Chen, P, Huang, JZ and Gao, X (2014) LigandRFs: random forest ensemble to identify ligand-binding residues from sequence information alone. In BMC Bioinformatics 15, 112.Google Scholar
Chen, S, et al. (2024) RNA-binding small molecules in drug discovery and delivery: an overview from fundamentals. Journal of Medicinal Chemistry 67, 1600216017.Google Scholar
Chen, T and Guestrin, C (2016) XGBoost: a scalable tree boosting system. In Proceedings of the 22nd ACM SIGKDD International Conference on Knowledge Discovery and Data Mining, 785794.Google Scholar
Chen, Z, et al. (2019) D3Pockets: a method and web server for systematic analysis of protein pocket dynamics. Journal of Chemical Information and Modeling 59, 33533358.Google Scholar
Chen, Z, et al. (2013) ZincExplorer: an accurate hybrid method to improve the prediction of zinc-binding sites from protein sequences. Molecular BioSystems 9, 22132222.Google Scholar
Cheng, H, Sen, TZ, Jernigan, RL and Kloczkowski, A (2007) Consensus data mining (CDM) protein secondary structure prediction server: combining GOR V and fragment database mining (FDM). Bioinformatics 23, 26282630.Google Scholar
Choi, D, et al. (2017) Predicting protein-binding regions in RNA using nucleotide profiles and compositions. BMC Systems Biology 11, 112.Google Scholar
Choi, S and Han, K (2013) Predicting protein-binding RNA nucleotides using the feature-based removal of data redundancy and the interaction propensity of nucleotide triplets. Computers in Biology and Medicine 43, 16871697.Google Scholar
Clackson, T and Wells, JA (1995) A hot spot of binding energy in a hormone-receptor interface. Science 267, 383386.Google Scholar
Coleman, RG and Sharp, KA (2006) Travel depth, a new shape descriptor for macromolecules: application to ligand binding. Journal of Molecular Biology 362, 441458.Google Scholar
Consortium, U (2019) UniProt: a worldwide hub of protein knowledge. Nucleic Acids Research 47, D506D515.Google Scholar
Cortes, C and Vapnik, VN (1995) Support-vector networks. Machine Learning 20, 273297.Google Scholar
Craig, IR, et al. (2011) Pocket-space maps to identify novel binding-site conformations in proteins. Journal of Chemical Information and Modeling 51, 26662679.Google Scholar
Cui, Y, et al. (2019) Predicting protein-ligand binding residues with deep convolutional neural networks. BMC Bioinformatics 20, 112.Google Scholar
Dai, B and Bailey-Kellogg, C (2021) Protein interaction interface region prediction by geometric deep learning. Bioinformatics 37, 25802588.Google Scholar
Dana, JM, et al. (2019) SIFTS: updated structure integration with function, taxonomy and sequences resource allows 40-fold increase in coverage of structure-based annotations for proteins. Nucleic Acids Research 47, D482D489.Google Scholar
De Baaij, JHF, Hoenderop, GJ and Bindels, RJM (2015) Magnesium in man: implications for health and disease. Physiological Reviews 95(1) 146Google Scholar
Delaney, JS (1992) Finding and filling protein cavities using cellular logic operations. Journal of Molecular Graphics 10, 174177.Google Scholar
Del Carpio, CA, Takahashi, Y and Sasaki, SI (1993) A new approach to the automatic identification of candidates for ligand receptor sites in proteins: (I) search for pocket regions. Journal of Molecular Graphics 11, 2329.Google Scholar
Denessiouk, KA, Rantanen, VV and Johnson, MS (2001) Adenine recognition: a motif present in ATP-, CoA-, NAD-, NADP-, and FAD-dependent proteins. Proteins: Structure, Function, and Bioinformatics 44, 282291.Google Scholar
Desaphy, J, et al. (2015) sc-PDB: a 3D-database of ligandable binding sites—10 years on. Nucleic Acids Research 43, D399D404.Google Scholar
Devlin, J, et al. (2018) BERT: pre-training of deep bidirectional transformers for language understanding. arXiv preprint. arXiv:1810.04805.Google Scholar
Devos, D and Valencia, A (2000) Practical limits of function prediction. Proteins: Structure, Function, and Bioinformatics 41, 98107.Google Scholar
Dewey, JA, et al. (2023) Molecular glue discovery: current and future approaches. Journal of Medicinal Chemistry 66, 92789296.Google Scholar
Ding, P, et al. (2023) DeepSTF: predicting transcription factor binding sites by interpretable deep neural networks combining sequence and shape. Briefings in Bioinformatics, 24(4) bbad231.Google Scholar
Ding, Y, Tang, J and Guo, F (2017) Identification of protein–ligand binding sites by sequence information and ensemble classifier. Journal of Chemical Information and Modeling 57, 31493161.Google Scholar
Di Pietro, O, et al. (2017) Unveiling a novel transient druggable pocket in BACE-1 through molecular simulations: conformational analysis and binding mode of multisite inhibitors. PLoS One 12, e0177683.Google Scholar
Dor, O and Zhou, Y (2007) Real-SPINE: an integrated system of neural networks for real-value prediction of protein structural properties. Proteins: Structure, Function, and Bioinformatics 68, 7681.Google Scholar
Dou, Y, et al. (2012) L1pred: a sequence-based prediction tool for catalytic residues in enzymes with the L1-Logreg classifier. PLoS One 7, e35666.Google Scholar
Doyle, SK, et al. (2016) Advances in discovering small molecules to probe protein function in a systems context. Current Opinion in Chemical Biology 30, 2836.Google Scholar
Du, H, et al. (2021) CovalentInDB: a comprehensive database facilitating the discovery of covalent inhibitors. Nucleic Acids Research 49, D1122D1129.Google Scholar
Du, H, et al. (2022) Proteome-wide profiling of the covalent-druggable cysteines with a structure-based deep graph learning network. Research 2022, 9873564.Google Scholar
Dudev, T and Lim, C (2014) Competition among metal ions for protein binding sites: determinants of metal ion selectivity in proteins. Chemical Reviews 114, 538556.Google Scholar
Dürr, SL, Levy, A and Rothlisberger, U (2023) Metal3D: a general deep learning framework for accurate metal ion location prediction in proteins. Nature Communications 14, 2713.Google Scholar
Durrant, JD, et al. (2014) POVME 2.0: an enhanced tool for determining pocket shape and volume characteristics. Journal of Chemical Theory and Computation 10, 50475056.Google Scholar
Ebert, JC and Altman, RB (2008) Robust recognition of zinc binding sites in proteins. Protein Science 17, 5465.Google Scholar
Eguida, M and Rognan, D (2022) Estimating the similarity between protein pockets. International Journal of Molecular Sciences 23, 12462.Google Scholar
Elnaggar, A, et al. (2021) ProtTrans: toward understanding the language of life through self-supervised learning. IEEE Transactions on Pattern Analysis and Machine Intelligence 44, 71127127.Google Scholar
Essien, C, et al. (2023) Prediction of protein ion–ligand binding sites with ELECTRA. Molecules 28, 6793.Google Scholar
Essien, C, Wang, D and Xu, D (2019) Capsule network for predicting zinc binding sites in metalloproteins. In 2019 IEEE International Conference on Bioinformatics and Biomedicine (BIBM), 23372341.Google Scholar
Everingham, M, et al. (2010) The Pascal Visual Object Classes (VOC) challenge. International Journal of Computer Vision 88, 303338.Google Scholar
Evteev, SA, Ereshchenko, AV and Ivanenkov, YA (2023) SiteRadar: utilizing graph machine learning for precise mapping of protein–ligand-binding sites. Journal of Chemical Information and Modeling 63, 11241132.Google Scholar
Fadahunsi, AA, et al. (2024) Revolutionizing drug discovery: an AI-powered transformation of molecular docking. Medicinal Chemistry Research, 33(12) 21872203.Google Scholar
Falese, JP, Donlic, A and Hargrove, AE (2021) Targeting RNA with small molecules: from fundamental principles towards the clinic. Chemical Society Reviews 50, 22242243.Google Scholar
Faller, CE, et al. (2015) Site identification by ligand competitive saturation (SILCS) simulations for fragment-based drug design. In Klon, Anthony E. Fragment-Based Methods in Drug Discovery, New York, NY: Springer New York 7587.Google Scholar
Fang, Y, et al. (2023) DeepProSite: structure-aware protein binding site prediction using ESMFold and pretrained language model. Bioinformatics 39, btad718.Google Scholar
Faraggi, E, et al. (2012) SPINE X: improving protein secondary structure prediction by multistep learning coupled with prediction of solvent accessible surface area and backbone torsion angles. Journal of Computational Chemistry 33, 259267.Google Scholar
Ferrè, F and Clote, P (2006) DiANNA 1.1: an extension of the DiANNA web server for ternary cysteine classification. Nucleic Acids Research 34, W182W185.Google Scholar
Ferré, S, et al. (2014) G protein–coupled receptor oligomerization revisited: functional and pharmacological perspectives. Pharmacological Reviews 66, 413434.Google Scholar
Fout, A, et al. (2017) Protein interface prediction using graph convolutional networks. Advances in Neural Information Processing Systems 30, 65306539.Google Scholar
Friedman, N, Geiger, D and Goldszmidt, M (1997) Bayesian network classifiers. Machine Learning 29, 131163.Google Scholar
Fuller, JC, Burgoyne, NJ and Jackson, RM (2009) Predicting druggable binding sites at the protein–protein interface. Drug Discovery Today 14, 155161.Google Scholar
Gagliardi, L and Rocchia, W (2023) SiteFerret: beyond simple pocket identification in proteins. Journal of Chemical Theory and Computation 19, 52425259.Google Scholar
Gainza, P, et al. (2020) Deciphering interaction fingerprints from protein molecular surfaces using geometric deep learning. Nature Methods 17, 184192.Google Scholar
Gallo Cassarino, T, Bordoli, L and Schwede, T (2014) Assessment of ligand binding site predictions in CASP10. Proteins: Structure, Function, and Bioinformatics 82, 154163.Google Scholar
Gamouh, H, Hoksza, D and Novotny, M (2023) Hybrid protein-ligand binding residue prediction with protein language models: does the structure matter? bioRxiv, 2023–08.Google Scholar
Gao, H and Ji, S (2019) Graph U-Nets. In International Conference on Machine Learning, 20832092. PMLR.Google Scholar
Gao, J, et al. (2016) BSiteFinder, an improved protein-binding sites prediction server based on structural alignment: more accurate and less time-consuming. Journal of Cheminformatics 8, 110.Google Scholar
Gao, M and Günther, S (2023) HyperCys: a structure- and sequence-based predictor of hyper-reactive druggable cysteines. International Journal of Molecular Sciences 24, 5960.Google Scholar
Gao, M, et al. (2022) CovPDB: a high-resolution coverage of the covalent protein–ligand interactome. Nucleic Acids Research 50, D445D450.Google Scholar
Gao, Z, et al. (2023) Hierarchical graph learning for protein–protein interaction. Nature Communications 14, 1093.Google Scholar
Garg, A, Kaur, H and Raghava, GPS (2005) Real value prediction of solvent accessibility in proteins using multiple sequence alignment and secondary structure. Proteins: Structure, Function, and Bioinformatics 61, 318324.Google Scholar
Garg, A and Pal, D (2021) Inferring metal binding sites in flexible regions of proteins. Proteins: Structure, Function, and Bioinformatics 89(9) 11251133.Google Scholar
Gazizov, A, et al. (2023) AF2BIND: predicting ligand-binding sites using the pair representation of AlphaFold2. bioRxiv, 2023–10.Google Scholar
Gelis, L, et al. (2012) Prediction of a ligand-binding niche within a human olfactory receptor by combining site-directed mutagenesis with dynamic homology modeling. Angewandte Chemie International Edition 51, 12741278.Google Scholar
Ghersi, D and Sanchez, R (2009) EasyMIFS and SiteHound: a toolkit for the identification of ligand-binding sites in protein structures. Bioinformatics 25, 31853186.Google Scholar
Gilson, MK, et al. (2016) BindingDB in 2015: a public database for medicinal chemistry, computational chemistry and systems pharmacology. Nucleic Acids Research 44, D1045D1053.Google Scholar
Glaser, F, et al. (2006) A method for localizing ligand binding pockets in protein structures. Proteins: Structure, Function, and Bioinformatics 62, 479488.Google Scholar
Gold, ND and Jackson, RM (2006) SitesBase: a database for structure-based protein–ligand binding site comparisons. Nucleic Acids Research 34, D231D234.Google Scholar
Goodford, PJ (1985) A computational procedure for determining energetically favorable binding sites on biologically important macromolecules. Journal of Medicinal Chemistry 28, 849857.Google Scholar
Gu, L, Li, B and Ming, D (2022) A multilayer dynamic perturbation analysis method for predicting ligand–protein interactions. BMC Bioinformatics 23, 456.Google Scholar
Guerra, JVdS, et al. (2021) pyKVFinder: an efficient and integrable Python package for biomolecular cavity detection and characterization in data science. BMC Bioinformatics 22, 607.Google Scholar
Gutteridge, A, Bartlett, GJ and Thornton, JM (2003) Using a neural network and spatial clustering to predict the location of active sites in enzymes. Journal of Molecular Biology 330, 719734.Google Scholar
Haberal, I and Oğul, H (2019) Prediction of protein metal binding sites using deep neural networks. Molecular Informatics 38, 1800169.Google Scholar
Haberal, I and Oğul, H (2017) DeepMBS: prediction of protein metal binding-site using deep learning networks. In 2017 Fourth International Conference on Mathematics and Computers in Sciences and Industry (MCSI), 2125.Google Scholar
Halgren, TA (2009) Identifying and characterizing binding sites and assessing druggability. Journal of Chemical Information and Modeling 49, 377389.Google Scholar
Hardy, JA and Wells, JA (2004) Searching for new allosteric sites in enzymes. Current Opinion in Structural Biology 14, 706715.Google Scholar
Hartshorn, MJ, et al. (2007) Diverse, high-quality test set for the validation of protein–ligand docking performance. Journal of Medicinal Chemistry 50, 726741.Google Scholar
Heffernan, R, et al. (2015) Improving prediction of secondary structure, local backbone angles and solvent accessible surface area of proteins by iterative deep learning. Scientific Reports 5, 11476.Google Scholar
Hendlich, M, Rippmann, F and Barnickel, G (1997) LIGSITE: automatic and efficient detection of potential small molecule-binding sites in proteins. Journal of Molecular Graphics and Modelling 15, 359363.Google Scholar
Hendrix, SG, et al. (2021) DeepDISE: DNA binding site prediction using a deep learning method. International Journal of Molecular Sciences 22, 5510.Google Scholar
Henrich, S, et al. (2010) Computational approaches to identifying and characterizing protein binding sites for ligand design. Journal of Molecular Recognition: An Interdisciplinary Journal 23, 209219.Google Scholar
Hernandez, M, Ghersi, D and Sanchez, R (2009) SITEHOUND-Web: a server for ligand binding site identification in protein structures. Nucleic Acids Research 37, W413W416.Google Scholar
Ho, CMW and Marshall, GR (1990) Cavity search: an algorithm for the isolation and display of cavity-like binding regions. Journal of Computer-Aided Molecular Design 4, 337354.Google Scholar
Ho, QT, et al. (2021) FAD-BERT: improved prediction of FAD binding sites using pre-training of deep bidirectional transformers. Computers in Biology and Medicine 131, 104258.Google Scholar
Ho, TK (1995) Random decision forests. In Proceedings of 3rd International Conference on Document Analysis and Recognition 1, 278282. IEEE.Google Scholar
Hopkins, AL and Groom, CR (2002) The druggable genome. Nature Reviews Drug Discovery 1, 727.Google Scholar
Hu, X, et al. (2016) Recognizing metal and acid radical ion-binding sites by integrating ab initio modeling with template-based transferals. Bioinformatics 32, 32603269.Google Scholar
Huang, B and Schroeder, M (2006) LIGSITE CSC: predicting ligand binding sites using the Connolly surface and degree of conservation. BMC Structural Biology 6, 111.Google Scholar
Huang, G, et al. (2017) Densely connected convolutional networks. In Proceedings of the IEEE Conference on Computer Vision and Pattern Recognition, 47004708.Google Scholar
Huey, R, et al. (2007) A semiempirical free energy force field with charge-based desolvation. Journal of Computational Chemistry 28, 11451152.Google Scholar
Humphrey, W, Dalke, A and Schulten, K (1996) VMD: visual molecular dynamics. Journal of Molecular Graphics 14, 3338.Google Scholar
Ingraham, J, et al. (2019) Generative models for graph-based protein design. Advances in Neural Information Processing Systems 32, 1582015831.Google Scholar
Iqbal, S and Hoque, MT (2018) PBRpredict-suite: a suite of models to predict peptide-recognition domain residues from protein sequence. Bioinformatics 34, 32893299.Google Scholar
Ireland, SM and Martin, ACR (2019) ZincBind—the database of zinc binding sites. Database 2019, baz006.Google Scholar
Ireland, SM and Martin, ACR (2021) ZincBindPredict—prediction of zinc binding sites in proteins. Molecules 26, 966.Google Scholar
Jernigan, RL, Raghunathan, G and Bahar, I (1994) Characterization of interactions and metal ion binding sites in proteins. Current Opinion in Structural Biology 4, 256263.Google Scholar
Jha, K, Karmakar, S and Saha, S (2023) Graph-BERT and language model-based framework for protein–protein interaction identification. Scientific Reports 13, 5663.Google Scholar
Jian, JW, et al. (2016) Predicting ligand binding sites on protein surfaces by 3-dimensional probability density distributions of interacting atoms. PLoS One 11, e0160315.Google Scholar
Jiang, M, et al. (2019a) A novel protein descriptor for the prediction of drug binding sites. BMC Bioinformatics 20, 113.Google Scholar
Jiang, M, et al. (2019b) FrSite: protein drug binding site prediction based on Faster R–CNN. Journal of Molecular Graphics and Modelling 93, 107454.Google Scholar
Jiang, Z, et al. (2016) Identification of Ca2+-binding residues of a protein from its primary sequence. Genetics and Molecular Research 15(2), gmr.15027618.Google Scholar
Jiménez, J, et al. (2017) DeepSite: protein-binding site predictor using 3D-convolutional neural networks. Bioinformatics 33, 30363042.Google Scholar
Jiménez-Luna, J, Grisoni, F and Schneider, G (2020) Drug discovery with explainable artificial intelligence. Nature Machine Intelligence 2, 573584.Google Scholar
Jing, B, et al. (2020) Learning from protein structure with geometric vector perceptrons. arXiv preprint. arXiv:2009.01411.Google Scholar
Johansson-Åkhe, I, Mirabello, C and Wallner, B (2019) Predicting protein-peptide interaction sites using distant protein complexes as structural templates. Scientific Reports 9, 113.Google Scholar
Jolma, A, et al. (2010) Multiplexed massively parallel SELEX for characterization of human transcription factor binding specificities. Genome Research 20, 861873.Google Scholar
Jumper, J, et al. (2021) Highly accurate protein structure prediction with AlphaFold. Nature 596, 583589.Google Scholar
Kalicki, CH and Haritaoglu, ED (2022) RNABERT: RNA family classification and secondary structure prediction with BERT pretrained on RNA sequences.Google Scholar
Kandel, J, Tayara, H and Chong, KT (2021) PUResNet: prediction of protein-ligand binding sites using deep residual neural network. Journal of Cheminformatics 13, 114.Google Scholar
Kapoor, S and Narayanan, A (2023) Leakage and the reproducibility crisis in machine-learning-based science. Patterns 4(9), 100804.Google Scholar
Katritch, V, et al. (2014) Allosteric sodium in class A GPCR signaling. Trends in Biochemical Sciences 39, 233244.Google Scholar
Kauffman, C and Karypis, G (2009) LIBRUS: combined machine learning and homology information for sequence-based ligand-binding residue prediction. Bioinformatics 25, 30993107.Google Scholar
Kawabata, T (2010) Detection of multiscale pockets on protein surfaces using mathematical morphology. Proteins: Structure, Function, and Bioinformatics 78, 11951211.Google Scholar
Kawabata, T and Go, N (2007) Detection of pockets on protein surfaces using small and large probe spheres to find putative ligand binding sites. Proteins: Structure, Function, and Bioinformatics 68, 516529.Google Scholar
Kawashima, S, et al. (2007) AAindex: amino acid index database, progress report 2008. Nucleic Acids Research 36, D202D205.Google Scholar
Kharchenko, PV, Tolstorukov, MY and Park, PJ (2008) Design and analysis of ChIP-Seq experiments for DNA-binding proteins. Nature Biotechnology 26, 13511359.Google Scholar
Kim, D, et al. (2008) Pocket extraction on proteins via the Voronoi diagram of spheres. Journal of Molecular Graphics and Modelling 26, 11041112.Google Scholar
Kinoshita, K and Nakamura, H (2005) Identification of the ligand binding sites on the molecular surface of proteins. Protein Science 14, 711718.Google Scholar
Kleywegt, GJ and Jones, TA (1994) Detection, delineation, measurement and display of cavities in macromolecular structures. Acta Crystallographica Section D: Biological Crystallography 50, 178185.Google Scholar
Knox, C, et al. (2024) DrugBank 6.0: the DrugBank knowledgebase for 2024. Nucleic Acids Research 52, D1265D1275.Google Scholar
Kognole, AA, Hazel, A and MacKerell, AD Jr (2022) SILCS-RNA: toward a structure-based drug design approach for targeting RNAs with small molecules. Journal of Chemical Theory and Computation 18(9), 56725691.Google Scholar
Konc, J and Janežič, D (2010) ProBiS algorithm for detection of structurally similar protein binding sites by local structural alignment. Bioinformatics 26, 11601168.Google Scholar
Kozlovskii, I and Popov, P (2020) Spatiotemporal identification of druggable binding sites using deep learning. Communications Biology 3, 112.Google Scholar
Kozlovskii, I and Popov, P (2021a) Protein–peptide binding site detection using 3D convolutional neural networks. Journal of Chemical Information and Modeling 61, 38143823.Google Scholar
Kozlovskii, I and Popov, P (2021b) Structure-based deep learning for binding site detection in nucleic acid macromolecules. NAR Genomics and Bioinformatics 3, lqab111.Google Scholar
Krapp, LF, et al. (2023) PeSTo: parameter-free geometric deep learning for accurate prediction of protein binding interfaces. Nature Communications 14, 2175.Google Scholar
Krishna, R, et al. (2024) Generalized biomolecular modeling and design with RosettaFold all-atom. Science 384, eadl2528.Google Scholar
Krivák, R and Hoksza, D (2018) P2Rank: machine learning based tool for rapid and accurate prediction of ligand binding sites from protein structure. Journal of Cheminformatics 10, 112.Google Scholar
Krivák, R, Jendele, L and Hoksza, D (2018) Peptide-binding site prediction from protein structure via points on the solvent accessible surface. In Proceedings of the 2018 ACM International Conference on Bioinformatics, Computational Biology, and Health Informatics, 645650.Google Scholar
Kryshtafovych, A, et al. (2021) Critical assessment of methods of protein structure prediction (CASP)—round XIV. Proteins: Structure, Function, and Bioinformatics 89, 16071617.Google Scholar
Kusuma, RMI, et al. (2019) Prediction of ATP-binding sites in membrane proteins using a two-dimensional convolutional neural network. Journal of Molecular Graphics and Modelling 92, 8693.Google Scholar
Laskowski, RA (1995) SURFNET: a program for visualizing molecular surfaces, cavities, and intermolecular interactions. Journal of Molecular Graphics 13, 323330.Google Scholar
Laskowski, RA, Gerick, F and Thornton, JM (2009) The structural basis of allosteric regulation in proteins. FEBS Letters 583, 16921698.Google Scholar
Laurent, B, et al. (2015) Epock: rapid analysis of protein pocket dynamics. Bioinformatics 31, 14781480.Google Scholar
Laurie, ATR and Jackson, RM (2005) Q-SiteFinder: an energy-based method for the prediction of protein–ligand binding sites. Bioinformatics 21, 19081916.Google Scholar
Laurie, AT and Jackson, RM (2006) Methods for the prediction of protein-ligand binding sites for structure-based drug design and virtual ligand screening. Current Protein and Peptide Science 7, 395406.Google Scholar
Laveglia, V, et al. (2023) Hunting down zinc(II)-binding sites in proteins with distance matrices. Bioinformatics 39, btad653.Google Scholar
Lavi, A, et al. (2013) Detection of peptide-binding sites on protein surfaces: the first step toward the modeling and targeting of peptide-mediated interactions. Proteins: Structure, Function, and Bioinformatics 81, 20962105.Google Scholar
Lawson, ADG (2012) Antibody-enabled small-molecule drug discovery. Nature Reviews Drug Discovery 11, 519.Google Scholar
Lee, HS and Im, W (2013) Ligand binding site detection by local structure alignment and its performance complementarity. Journal of Chemical Information and Modeling 53, 24622470.Google Scholar
Lee, I and Nam, H (2022) Sequence-based prediction of protein binding regions and drug–target interactions. Journal of Cheminformatics 14, 115.Google Scholar
Lee, PH and Helms, V (2012) Identifying continuous pores in protein structures with PROPORES by computational repositioning of gating residues. Proteins: Structure, Function, and Bioinformatics 80, 421432.Google Scholar
Le Guilloux, V, Schmidtke, P and Tuffery, P (2009) Fpocket: an open-source platform for ligand pocket detection. BMC Bioinformatics 10, 111.Google Scholar
Leis, S, Schneider, S and Zacharias, M (2010) In silico prediction of binding sites on proteins. Current Medicinal Chemistry 17, 15501562.Google Scholar
Levitt, DG and Banaszak, LJ (1992) POCKET: a computer graphics method for identifying and displaying protein cavities and their surrounding amino acids. Journal of Molecular Graphics 10, 229234.Google Scholar
Li, H, Pi, D and Chen, C (2019a) An improved prediction model for zinc-binding sites in proteins based on Bayesian method. Tehnički Vjesnik 26, 14221426.Google Scholar
Li, H, et al. (2019b) A novel prediction method for zinc-binding sites in proteins by an ensemble of SVM and sample-weighted probabilistic neural network. IEEE Access 7, 186147186157.Google Scholar
Li, K, et al. (2023a) Simultaneous prediction of interaction sites on the protein and peptide sides of complexes through multilayer graph convolutional networks. Journal of Chemical Information and Modeling 63, 22512262.Google Scholar
Li, P, et al. (2022) RecurPocket: recurrent LMSER network with gating mechanism for protein binding site detection. In 2022 IEEE International Conference on Bioinformatics and Biomedicine (BIBM), 334339. IEEE.Google Scholar
Li, P, et al. (2023b) GLPocket: a multi-scale representation learning approach for protein binding site prediction. In Proceedings of the Thirty-Second International Joint Conference on Artificial Intelligence, 48214828.Google Scholar
Li, S, et al. (2023c) PocketAnchor: learning structure-based pocket representations for protein-ligand interaction prediction. Cell Systems 14, 692705.Google Scholar
Li, Y, et al. (2023d) MsPBRsP: multi-scale protein binding residues prediction using language model. bioRxiv, 2023–02.Google Scholar
Liang, J, Woodward, C and Edelsbrunner, H (1998) Anatomy of protein pockets and cavities: measurement of binding site geometry and implications for ligand design. Protein Science 7, 18841897.Google Scholar
Liang, MP, et al. (2003) WebFEATURE: an interactive web tool for identifying and visualizing functional sites on macromolecular structures. Nucleic Acids Research 31, 33243327.Google Scholar
Liang, S, et al. (2006) Protein binding site prediction using an empirical scoring function. Nucleic Acids Research 34, 36983707.Google Scholar
Liao, J, et al. (2022) In silico methods for identification of potential active sites of therapeutic targets. Molecules 27, 7103.Google Scholar
Lin, HN, et al. (2005) HYPROSP II-A knowledge-based hybrid method for protein secondary structure prediction based on local prediction confidence. Bioinformatics 21, 32273233.Google Scholar
Lin, YF, et al. (2016) MIB: metal ion-binding site prediction and docking server. Journal of Chemical Information and Modeling 56, 22872291.Google Scholar
Lin, Z, et al. (2023) Evolutionary-scale prediction of atomic-level protein structure with a language model. Science 379, 11231130.Google Scholar
Lippi, M, et al. (2008) MetalDetector: a web server for predicting metal-binding sites and disulfide bridges in proteins from sequence. Bioinformatics 24, 20942095.Google Scholar
Litfin, T, Yang, Y and Zhou, Y (2019) Spot-Peptide: template-based prediction of peptide-binding proteins and peptide-binding sites. Journal of Chemical Information and Modeling 59, 924930.Google Scholar
Littmann, M, et al. (2021) Protein embeddings and deep learning predict binding residues for various ligand classes. Scientific Reports 11, 23916.Google Scholar
Liu, H, et al. (2024) RNet: a network strategy to predict RNA binding preferences. Briefings in Bioinformatics 25, bbad482.Google Scholar
Liu, H, et al. (2020a) Illuminating the allosteric modulation of the calcium-sensing receptor. Proceedings of the National Academy of Sciences 117, 2171121722.Google Scholar
Liu, X and Ciulli, A (2023) Proximity-based modalities for biology and medicine. ACS Central Science 9, 12691284.Google Scholar
Liu, Y, et al. (2020b) CB-Dock: a web server for cavity detection-guided protein–ligand blind docking. Acta Pharmacologica Sinica 41, 138144.Google Scholar
Liu, Y, et al. (2023) RefinePocket: an attention-enhanced and mask-guided deep learning approach for protein binding site prediction. IEEE/ACM Transactions on Computational Biology and Bioinformatics 20(5) 33143321.Google Scholar
Liu, Y and Tian, B (2023) Protein-DNA binding sites prediction based on pre-trained protein language model and contrastive learning. arXiv preprint. arXiv:2306.15912.Google Scholar
Liu, Z, et al. (2014) Computationally characterizing and comprehensive analysis of zinc-binding sites in proteins. Biochimica et Biophysica Acta (BBA)-Proteins and Proteomics 1844, 171180.Google Scholar
Liu, Z, et al. (2015) PDB-wide collection of binding data: current status of the PDBbind database. Bioinformatics 31, 405412.Google Scholar
Lopez, G, Ezkurdia, I and Tress, ML (2009) Assessment of ligand binding residue predictions in CASP8. Proteins: Structure, Function, and Bioinformatics 77, 138146.Google Scholar
López, G, Valencia, A and Tress, ML (2007) Firestar—prediction of functionally important residues using structural templates and alignment reliability. Nucleic Acids Research 35, W573W577.Google Scholar
Lu, C, et al. (2019) MPLs-Pred: predicting membrane protein-ligand binding sites using hybrid sequence-based features and ligand-specific models. International Journal of Molecular Sciences 20, 3120.Google Scholar
Lu, CH, et al. (2012) Prediction of metal ion–binding sites in proteins using the fragment transformation method. PLoS One 7, e39252.Google Scholar
Lu, S, et al. (2018) Discovery of hidden allosteric sites as novel targets for allosteric drug design. Drug Discovery Today 23, 359365.Google Scholar
Ludlow, RF, et al. (2015) Detection of secondary binding sites in proteins using fragment screening. Proceedings of the National Academy of Sciences 112, 1591015915.Google Scholar
Lundberg, SM and Lee, SI (2017) A unified approach to interpreting model predictions. In Guyon, I, Von Luxburg, U, Bengio, S, Wallach, H, Fergus, R, Vishwanathan, S and Garnett, R (eds.), Advances in Neural Information Processing Systems, Vol. 30. Red Hook, NY, USA: Curran Associates, Inc.Google Scholar
Luo, J, et al. (2017) RPI-Bind: a structure-based method for accurate identification of RNA-protein binding sites. Scientific Reports 7, 614.Google Scholar
Lv, N and Cao, Z (2024) Subpocket-based analysis approach for the protein pocket dynamics. Journal of Chemical Theory and Computation 20, 49094920.Google Scholar
Lyons, J, et al. (2014) Predicting backbone Cα angles and dihedrals from protein sequences by stacked sparse auto-encoder deep neural network. Journal of Computational Chemistry 35, 20402046.Google Scholar
Mallet, V, et al. (2022) InDeep: 3D fully convolutional neural networks to assist in silico drug design on protein–protein interactions. Bioinformatics 38, 12611268.Google Scholar
Marchand, JR, et al. (2021) CAVIAR: a method for automatic cavity detection, description and decomposition into subcavities. Journal of Computer-Aided Molecular Design 35, 737750.Google Scholar
Martinez-Rosell, G, et al. (2020) PlayMolecule CrypticScout: predicting protein cryptic sites using mixed-solvent molecular simulations. Journal of Chemical Information and Modeling 60, 23142324.Google Scholar
Masuya, M and Doi, J (1995) Detection and geometric modeling of molecular surfaces and cavities using digital mathematical morphological operations. Journal of Molecular Graphics 13, 331336.Google Scholar
Matsui, M and Corey, DR (2017) Non-coding RNAs as drug targets. Nature Reviews Drug Discovery 16, 167179.Google Scholar
McClorey, G and Wood, MJ (2015) An overview of the clinical application of antisense oligonucleotides for RNA-targeting therapies. Current Opinion in Pharmacology 24, 5258.Google Scholar
McGreig, JE, et al. (2022) 3DLigandSite: structure-based prediction of protein–ligand binding sites. Nucleic Acids Research 50, W13W20.Google Scholar
Meller, A, et al. (2023) Predicting the locations of cryptic pockets from single protein structures using the PocketMiner graph neural network. Biophysical Journal 122, 445a.Google Scholar
Ming, D and Wall, ME (2006) Interactions in native binding sites cause a large change in protein dynamics. Journal of Molecular Biology 358, 213223.Google Scholar
Möller, L, et al. (2022) Translating from proteins to ribonucleic acids for ligand-binding site detection. Molecular Informatics 41, 2200059.Google Scholar
Montavon, G, et al. (2019) Layer-wise relevance propagation: an overview. In Explainable AI: Interpreting, Explaining and Visualizing Deep Learning, Cham, Switzerland: Springer International Publishing 193209.Google Scholar
Moodie, SL, Mitchell, JBO and Thornton, JM (1996) Protein recognition of adenylate: an example of a fuzzy recognition template. Journal of Molecular Biology 263, 486500.Google Scholar
Morris, GM, et al. (2009) AutoDock4 and AutoDockTools4: automated docking with selective receptor flexibility. Journal of Computational Chemistry 30, 27852791.Google Scholar
Mukherjee, S, et al. (2004) Rapid analysis of the DNA-binding specificities of transcription factors with DNA microarrays. Nature Genetics 36, 13311339.Google Scholar
Murdoch, WJ, et al. (2019) Definitions, methods, and applications in interpretable machine learning. Proceedings of the National Academy of Sciences 116, 2207122080.Google Scholar
Mylonas, SK, Axenopoulos, A and Daras, P (2021) DeepSurf: a surface-based deep learning approach for the prediction of ligand binding sites on proteins. Bioinformatics 37, 16811690.Google Scholar
Nayal, M and Honig, B (2006) On the nature of cavities on protein surfaces: application to the identification of drug-binding sites. Proteins: Structure, Function, and Bioinformatics 63, 892906.Google Scholar
Naz, F, et al. (2015) Designing new kinase inhibitor derivatives as therapeutics against common complex diseases: structural basis of microtubule affinity-regulating kinase 4 (MARK4) inhibition. OMICS: A Journal of Integrative Biology 19, 700711.Google Scholar
Nazem, F, et al. (2021) 3D U-Net: a voxel-based method in binding site prediction of protein structure. Journal of Bioinformatics and Computational Biology 19, 2150006.Google Scholar
Nazem, F, et al. (2023) A GU-Net-based architecture predicting ligand–protein-binding atoms. Journal of Medical Signals and Sensors 13, 1.Google Scholar
Nemethy, G, et al. (1992) Energy parameters in polypeptides. 10. Improved geometrical parameters and nonbonded interactions for use in the ECEPP/3 algorithm, with application to proline-containing peptides. The Journal of Physical Chemistry 96, 64726484.Google Scholar
Ngan, CH, et al. (2012) FTSite: high accuracy detection of ligand binding sites on unbound protein structures. Bioinformatics 28, 286287.Google Scholar
Nimrod, G, et al. (2008) Detection of functionally important regions in ‘hypothetical proteins’ of known structure. Structure 16, 17551763.Google Scholar
Ortiz de Luzuriaga, I, Lopez, X and Gil, A (2021) Learning to model G-quadruplexes: current methods and perspectives. Annual Review of Biophysics 50, 209243.Google Scholar
Paiva, VA, et al. (2022) GASS-Metal: identifying metal-binding sites on protein structures using genetic algorithms. Briefings in Bioinformatics 23, bbac178.Google Scholar
Pan, X, et al. (2020) RBPsuite: RNA-protein binding sites prediction suite based on deep learning. BMC Genomics 21, 18.Google Scholar
Panchal, V and Brenk, R (2021) Riboswitches as drug targets for antibiotics. Antibiotics 10, 45.Google Scholar
Panchenko, AR, Kondrashov, F and Bryant, S (2004) Prediction of functional sites by analysis of sequence and structure conservation. Protein Science 13, 884892.Google Scholar
Pandit, SB and Skolnick, J (2008) Fr-TM-Align: a new protein structural alignment method based on fragment alignments and the TM-score. BMC Bioinformatics 9, 111.Google Scholar
Panei, FP, Gkeka, P and Bonomi, M (2024) Identifying small-molecules binding sites in RNA conformational ensembles with Shaman. Nature Communications 15, 5725.Google Scholar
Panwar, B, Gupta, S and Raghava, GPS (2013) Prediction of vitamin interacting residues in a vitamin binding protein using evolutionary information. BMC Bioinformatics 14, 114.Google Scholar
Panwar, B and Raghava, GPS (2015) Identification of protein-interacting nucleotides in an RNA sequence using composition profile of tri-nucleotides. Genomics 105, 197203.Google Scholar
Paramo, T, et al. (2014) Efficient characterization of protein cavities within molecular simulation trajectories: Trj_cavity. Journal of Chemical Theory and Computation 10, 21512164.Google Scholar
Park, S and Seok, C (2022) GalaxyWater-CNN: prediction of water positions on the protein structure by a 3D-convolutional neural network. Journal of Chemical Information and Modeling 62, 31573168.Google Scholar
Passerini, A, et al. (2007) Predicting zinc binding at the proteome level. BMC Bioinformatics 8, 113.Google Scholar
Passerini, A, Lippi, M and Frasconi, P (2011) MetalDetector V2.0: predicting the geometry of metal binding sites from protein sequence. Nucleic Acids Research 39, W288W292.Google Scholar
Passerini, A, et al. (2006) Identifying cysteines and histidines in transition-metal-binding sites using support vector machines and neural networks. Proteins: Structure, Function, and Bioinformatics 65, 305316.Google Scholar
Pei, J and Grishin, NV (2004) Combining evolutionary and structural information for local protein structure prediction. Proteins: Structure, Function, and Bioinformatics 56, 782794.Google Scholar
Pei, Q, et al. (2023) FABind: fast and accurate protein-ligand binding. arXiv preprint. arXiv:2310.06763.Google Scholar
Peters, KP, Fauck, J and Frömmel, C (1996) The automatic search for ligand binding sites in proteins of known three-dimensional structure using only geometric criteria. Journal of Molecular Biology 256, 201213.Google Scholar
Petrova, NV and Wu, CH (2006) Prediction of catalytic residues using support vector machine with selected protein sequence and structural properties. BMC Bioinformatics 7, 112.Google Scholar
Petrovski, ŽH, Hribar-Lee, B and Bosnić, Z (2022) CAT-Site: predicting protein binding sites using a convolutional neural network. Pharmaceutics 15, 119.Google Scholar
Petukh, M, Kimmet, T and Alexov, E (2013) BION web server: predicting non-specifically bound surface ions. Bioinformatics 29, 805806.Google Scholar
Philips, A, et al. (2012) MetalionRNA: computational predictor of metal-binding sites in RNA structures. Bioinformatics 28, 198205.Google Scholar
Polizzi, NF and DeGrado, WF (2020) A defined structural unit enables de novo design of small-molecule–binding proteins. Science 369, 12271233.Google Scholar
Popov, P, et al. (2024) Unraveling viral drug targets: a deep learning-based approach for the identification of potential binding sites. Briefings in Bioinformatics 25, bbad459.Google Scholar
Poulson, BG, et al. (2020) Aggregation of biologically important peptides and proteins: inhibition or acceleration depending on protein and metal ion concentrations. RSC Advances 10, 215227.Google Scholar
Pravda, L, et al. (2018) MOLEonline: a web-based tool for analyzing channels, tunnels and pores (2018 update). Nucleic Acids Research 46, W368W373.Google Scholar
Pupko, T, et al. (2002) Rate4Site: an algorithmic tool for the identification of functional regions in proteins by surface mapping of evolutionary determinants within their homologues. Bioinformatics 18, S71S77.Google Scholar
Qiao, L and Xie, D (2019) MIonSite: ligand-specific prediction of metal ion-binding sites via enhanced Adaboost algorithm with protein sequence information. Analytical Biochemistry 566, 7588.Google Scholar
Qiao, Z, et al. (2024) State-specific protein–ligand complex structure prediction with a multiscale deep generative model. Nature Machine Intelligence 6, 195208.Google Scholar
Qiu, Z and Wang, X (2011) Improved prediction of protein ligand-binding sites using random forests. Protein and Peptide Letters 18, 12121218.Google Scholar
Ravindranath, PA and Sanner, MF (2016) AutoSite: an automated approach for pseudo-ligands prediction—from ligand-binding sites identification to predicting key ligand atoms. Bioinformatics 32, 31423149.Google Scholar
Ray, D, et al. (2009) Rapid and systematic analysis of the RNA recognition specificities of RNA-binding proteins. Nature Biotechnology 27, 667670.Google Scholar
Ray, D, et al. (2013) A compendium of RNA-binding motifs for decoding gene regulation. Nature 499, 172177.Google Scholar
Redmon, J, et al. (2016) You only look once: unified, real-time object detection. In Proceedings of the IEEE Conference on Computer Vision and Pattern Recognition, 779788.Google Scholar
Rekand, IH and Brenk, R (2021) DrugPred_RNA—a tool for structure-based druggability predictions for RNA binding sites. Journal of Chemical Information and Modeling 61(8), 40684081.Google Scholar
Ren, S, et al. (2016) Faster R-CNN: towards real-time object detection with region proposal networks. IEEE Transactions on Pattern Analysis and Machine Intelligence 39, 11371149.Google Scholar
Renaud, N, et al. (2021) DeepRank: a deep learning framework for data mining 3D protein-protein interfaces. Nature Communications 12, 7068.Google Scholar
Rinaldis, M de, et al. (1998) Three-dimensional profiles: a new tool to identify protein surface similarities. Journal of Molecular Biology 284, 12111221.Google Scholar
Rives, A, et al. (2021) Biological structure and function emerge from scaling unsupervised learning to 250 million protein sequences. Proceedings of the National Academy of Sciences 118, e2016239118.Google Scholar
Roche, DB, Brackenridge, DA and McGuffin, LJ (2015) Proteins and their interacting partners: an introduction to protein–ligand binding site prediction methods. International Journal of Molecular Sciences 16, 2982929842.Google Scholar
Roche, R, et al. (2023) EquiPNAS: improved protein-nucleic acid binding site prediction using protein-language-model-informed equivariant deep graph neural networks. bioRxiv 2023, 2023–09.Google Scholar
Rodrı́guez-Guerra Pedregal, J, et al. (2017) GaudiMM: a modular multi-objective platform for molecular modeling. Wiley Online Library 38(24), 21182126.Google Scholar
Rodwell, VW, Bender, D and Botham, KM (2018) Harper’s Illustrated Biochemistry. New York, NY, USA: McGraw-Hill.Google Scholar
Ronneberger, O, Fischer, P and Brox, T (2015) U-Net: convolutional networks for biomedical image segmentation. In Medical Image Computing and Computer-Assisted Intervention–MICCAI 2015: 18th International Conference, Munich, Germany, October 5-9, 2015, Proceedings, Part III 18, 234241.Google Scholar
Rosen, M, et al. (1998) Molecular shape comparisons in searches for active sites and functional similarity. Protein Engineering 11, 263277.Google Scholar
Roy, A, Yang, J and Zhang, Y (2012) COFACTOR: an accurate comparative algorithm for structure-based protein function annotation. Nucleic Acids Research 40, W471W477.Google Scholar
Ruffner, H, Bauer, A and Bouwmeester, T (2007) Human protein–protein interaction networks and the value for drug discovery. Drug Discovery Today 12, 709716.Google Scholar
Ruppert, J, Welch, W and Jain, AN (1997) Automatic identification and representation of protein binding sites for molecular docking. Protein Science 6, 524533.Google Scholar
Sánchez-Aparicio, JE, et al. (2020) BioMetAll: identifying metal-binding sites in proteins from backbone preorganization. Journal of Chemical Information and Modeling 61, 311323.Google Scholar
Santana, CA, et al. (2020) GRaSP: a graph-based residue neighborhood strategy to predict binding sites. Bioinformatics 36, i726i734.Google Scholar
Satorras, VG, Hoogeboom, E and Welling, M (2021) E(N) equivariant graph neural networks. In International Conference on Machine Learning, 93239332. PMLR.Google Scholar
Schmidt, T, et al. (2011) Assessment of ligand-binding residue predictions in CASP9. Proteins: Structure, Function, and Bioinformatics 79, 126136.Google Scholar
Schmidtke, P, et al. (2011) MDpocket: open-source cavity detection and characterization on molecular dynamics trajectories. Bioinformatics 27, 32763285.Google Scholar
Schmidtke, P, et al. (2010) Large-scale comparison of four binding site detection algorithms. Journal of Chemical Information and Modeling 50, 21912200.Google Scholar
Schmitt, S, Kuhn, D and Klebe, G (2002) A new method to detect related function among proteins independent of sequence and fold homology. Journal of Molecular Biology 323, 387406.Google Scholar
Schymkowitz, J, et al. (2005a) The FoldX web server: an online force field. Nucleic Acids Research 33, W382W388.Google Scholar
Schymkowitz, JWH, et al. (2005b) Prediction of water and metal binding sites and their affinities by using the Fold-X force field. Proceedings of the National Academy of Sciences 102, 1014710152.Google Scholar
Sciortino, G, et al. (2019) Simple coordination geometry descriptors allow to accurately predict metal-binding sites in proteins. ACS Omega 4, 37263731.Google Scholar
Segura, J, Jones, PF and Fernandez-Fuentes, N (2012) A holistic in silico approach to predict functional sites in protein structures. Bioinformatics 28, 18451850.Google Scholar
Selvaraju, RR, et al. (2017) Grad-CAM: visual explanations from deep networks via gradient-based localization. In Proceedings of the IEEE International Conference on Computer Vision, 618626.Google Scholar
Seo, S, et al. (2024) Pseq2Sites: enhancing protein sequence-based ligand binding-site prediction accuracy via the deep convolutional network and attention mechanism. Engineering Applications of Artificial Intelligence 127, 107257.Google Scholar
Shafiee, S, Fathi, A and Taherzadeh, G (2022) SPPPred: sequence-based protein-peptide binding residue prediction using genetic programming and ensemble learning. IEEE/ACM Transactions on Computational Biology and Bioinformatics 20(3), 20292040.Google Scholar
Shashikala, HB, et al. (2021) BION-2: predicting positions of non-specifically bound ions on protein surface by a Gaussian-based treatment of electrostatics. International Journal of Molecular Sciences 22, 272.Google Scholar
Shenoy, A, et al. (2024) M-Ionic: prediction of metal-ion-binding sites from sequence using residue embeddings. Bioinformatics 40, btad782.Google Scholar
Shi, W, et al. (2022) GraphSite: ligand binding site classification with deep graph learning. Biomolecules 12, 1053.Google Scholar
Shu, N, Zhou, T and Hovmöller, S (2008) Prediction of zinc-binding sites in proteins from sequence. Bioinformatics 24, 775782.Google Scholar
Shulman-Peleg, A, Nussinov, R and Wolfson, HJ (2004) Recognition of functional sites in protein structures. Journal of Molecular Biology 339, 607633.Google Scholar
Simões, T, et al. (2017) Geometric detection algorithms for cavities on protein surfaces in molecular graphics: a survey. Computer Graphics Forum 36, 643683.Google Scholar
Smith, MC and Gestwicki, JE (2012) Features of protein-protein interactions that translate into potent inhibitors: topology, surface area and affinity. Expert Review of Molecular Medicine 14, e16.Google Scholar
Smith, Z, et al. (2023) Graph attention site prediction (Grasp): identifying druggable binding sites using graph neural networks with attention. bioRxiv.Google Scholar
Sodhi, JS, et al. (2004) Predicting metal-binding site residues in low-resolution structural models. Journal of Molecular Biology 342, 307320.Google Scholar
Song, C and Jiang, J (2023) A novel prediction method for metal-ion binding sites in protein sequence based on ensemble learning. In Proceedings of the 2022 5th International Conference on Algorithms, Computing and Artificial Intelligence. New York, NY, USA: Association for Computing Machinery, 17.Google Scholar
Song, Y, et al. (2023) Accurately identifying nucleic-acid-binding sites through geometric graph learning on language model predicted structures. Briefings in Bioinformatics 24, bbad360.Google Scholar
Southey, MWY and Brunavs, M (2023) Introduction to small molecule drug discovery and preclinical development. Frontiers in Drug Discovery 3, 1314077.Google Scholar
Spriggs, RV, Artymiuk, PJ and Willett, P (2003) Searching for patterns of amino acids in 3D protein structures. Journal of Chemical Information and Computer Sciences 43, 412421.Google Scholar
Srivastava, A and Kumar, M (2018) Prediction of zinc binding sites in proteins using sequence derived information. Journal of Biomolecular Structure and Dynamics 36(16), 44134423.Google Scholar
Stark, A, Sunyaev, S and Russell, RB (2003) A model for statistical significance of local similarities in structure. Journal of Molecular Biology 326, 13071316.Google Scholar
Stepniewska-Dziubinska, MM, Zielenkiewicz, P and Siedlecki, P (2020) Improving detection of protein-ligand binding sites with 3D segmentation. Scientific Reports 10, 5035.Google Scholar
Stourac, J, et al. (2019) Caver Web 1.0: identification of tunnels and channels in proteins and analysis of ligand transport. Nucleic Acids Research 47, W414W422.Google Scholar
Su, H, Peng, Z and Yang, J (2021) Recognition of small molecule-RNA binding sites using RNA sequence and structure. Bioinformatics 37(1), 3642.Google Scholar
Sun, K, et al. (2022) Predicting Ca2+ and Mg2+ ligand binding sites by deep neural network algorithm. BMC Bioinformatics 22, 324.Google Scholar
Sun, Z, et al. (2021) To improve prediction of binding residues with DNA, RNA, carbohydrate, and peptide via multi-task deep neural networks. IEEE/ACM Transactions on Computational Biology and Bioinformatics 19, 37353743.Google Scholar
Sun, Z, et al. (2020) Structure-based analysis of cryptic-site opening. Structure 28, 223235.Google Scholar
Sunseri, J and Koes, DR (2020) Libmolgrid: graphics processing unit accelerated molecular gridding for deep learning applications. Journal of Chemical Information and Modeling 60, 10791084.Google Scholar
Sverrisson, F, et al. (2021) Fast end-to-end learning on protein surfaces. In Proceedings of the IEEE/CVF Conference on Computer Vision and Pattern Recognition, 1527215281.Google Scholar
Taherzadeh, G, et al. (2016) Sequence-based prediction of protein–peptide binding sites using support vector machine. Journal of Computational Chemistry 37, 12231229.Google Scholar
Taherzadeh, G, et al. (2018) Structure-based prediction of protein–peptide binding regions using random forest. Bioinformatics 34, 477484.Google Scholar
Tahir, M, et al. (2021) KDeepBind: prediction of RNA-proteins binding sites using convolution neural network and k-gram features. Chemometrics and Intelligent Laboratory Systems 208, 104217.Google Scholar
Tan, X, et al. (2024) Molecular glue-mediated targeted protein degradation: a novel strategy in small-molecule drug development. iScience 27(9), 110712.Google Scholar
Todd, AE, Orengo, CA and Thornton, JM (2001) Evolution of function in protein superfamilies, from a structural perspective. Journal of Molecular Biology 307, 11131143.Google Scholar
Tong, W, et al. (2009) Partial order optimum likelihood (POOL): maximum likelihood prediction of protein active site residues using 3D structure and sequence properties. PLoS Computational Biology 5, e1000266.Google Scholar
Tong, Y, Childs-Disney, JL and Disney, MD (2024) Targeting RNA with small molecules, from RNA structures to precision medicines: IUPHAR Review: 40. British Journal of Pharmacology 181, 41524173.Google Scholar
Tora, AS, et al. (2015) Allosteric modulation of metabotropic glutamate receptors by chloride ions. The FASEB Journal 29, 41744188.Google Scholar
Trabuco, LG, et al. (2012) PepSite: prediction of peptide-binding sites from protein surfaces. Nucleic Acids Research 40, W423W427.Google Scholar
Tsaban, T, et al. (2022) Harnessing protein folding neural networks for peptide–protein docking. Nature Communications 13, 176.Google Scholar
Tsomaia, N (2015) Peptide therapeutics: targeting the undruggable space. European Journal of Medicinal Chemistry 94, 459470.Google Scholar
Tsujikawa, H, et al. (2016) Development of a protein–ligand-binding site prediction method based on interaction energy and sequence conservation. Journal of Structural and Functional Genomics 17, 3949.Google Scholar
Tubiana, J, Schneidman-Duhovny, D and Wolfson, HJ (2022) ScanNet: an interpretable geometric deep learning model for structure-based protein binding site prediction. Nature Methods 19, 730739.Google Scholar
Tuvshinjargal, N, et al. (2016) PRIdictor: protein–RNA interaction predictor. Biosystems 139, 1722.Google Scholar
Uhl, M, et al. (2019) GraphProt2: a graph neural network-based method for predicting binding sites of RNA-binding proteins. bioRxiv, 850024.Google Scholar
Ullmann, JR (1976) An algorithm for subgraph isomorphism. Journal of the ACM (JACM) 23, 3142.Google Scholar
Utgés, JS and Barton, GJ (2024) Comparative evaluation of methods for the prediction of protein–ligand binding sites. Journal of Cheminformatics 16, 126.Google Scholar
Valasatava, Y, et al. (2016) MetalPredator: a web server to predict iron–sulfur cluster binding proteomes. Bioinformatics 32, 28502852.Google Scholar
Valasatava, Y, et al. (2014) MetalS3, a database-mining tool for the identification of structurally similar metal sites. JBIC Journal of Biological Inorganic Chemistry 19, 937945.Google Scholar
Van Der Spoel, D, et al. (2005) GROMACS: fast, flexible, and free. Journal of Computational Chemistry 26, 17011718.Google Scholar
Vanhee, P, et al. (2011) BriX: a database of protein building blocks for structural analysis, modeling and design. Nucleic Acids Research 39, D435D442.Google Scholar
Vaswani, A, et al. (2017) Attention is all you need. Advances in Neural Information Processing Systems 30, 59986008.Google Scholar
Vecchietti, LF, et al. (2024) Recent advances in interpretable machine learning using structure-based protein representations. arXiv Preprint. arXiv:2409.17726.Google Scholar
Verdonk, ML, et al. (2001) SuperStar: improved knowledge-based interaction fields for protein binding sites. Journal of Molecular Biology 307, 841859.Google Scholar
Verschueren, E, et al. (2013) Protein-peptide complex prediction through fragment interaction patterns. Structure 21, 789797.Google Scholar
Viet Hung, L, et al. (2015) LIBRA: ligand binding site recognition application. Bioinformatics 31, 40204022.Google Scholar
Volkamer, A, et al. (2010) Analyzing the topology of active sites: on the prediction of pockets and subpockets. Journal of Chemical Information and Modeling 50, 20412052.Google Scholar
Wade, RC, Clark, KJ and Goodford, PJ (1993) Further development of hydrogen bond functions for use in determining energetically favorable binding sites on molecules of known structure. 1. Ligand probe groups with the ability to form two hydrogen bonds. Journal of Medicinal Chemistry 36, 140147.Google Scholar
Wagner, JR, et al. (2016) Emerging computational methods for the rational discovery of allosteric drugs. Chemical Reviews 116, 63706390.Google Scholar
Wagner, JR, et al. (2017) POVME 3.0: software for mapping binding pocket flexibility. Journal of Chemical Theory and Computation 13, 45844592.Google Scholar
Wang, J, et al. (2018a) Druggable negative allosteric site of P2X3 receptors. Proceedings of the National Academy of Sciences 115, 49394944.Google Scholar
Wang, J, et al. (2024) MultiModRLBP: a deep learning approach for multi-modal RNA-small molecule ligand binding sites prediction. IEEE Journal of Biomedical and Health Informatics 28(8), 49955006.Google Scholar
Wang, J, et al. (2023a) Computational modeling of binding site for RNA-ligand complex by learning multi-modal features. Authorea Preprints.Google Scholar
Wang, K, et al. (2018b) RBind: computational network method to predict RNA binding sites. Bioinformatics 34, 31313136.Google Scholar
Wang, K, et al. (2023b) RLBind: a deep learning method to predict RNA–ligand binding sites. Briefings in Bioinformatics 24, bbac486.Google Scholar
Wang, R, et al. (2022a) Predicting protein–peptide binding residues via interpretable deep learning. Bioinformatics 38, 33513360.Google Scholar
Wang, T, He, Y and Zhu, F (2023c) SAPocket: finding pockets on protein surfaces with a focus towards position and voxel channels. Expert Systems with Applications 227, 120235.Google Scholar
Wang, W, et al. (2023d) GraphPLBR: protein-ligand binding residue prediction with deep graph convolution network. IEEE/ACM Transactions on Computational Biology and Bioinformatics 20(3), 22232232.Google Scholar
Wang, X, et al. (2022b) DUnet: a deep learning guided protein-ligand binding pocket prediction. bioRxiv, 2022–08.Google Scholar
Wardah, W, et al. (2020) Predicting protein-peptide binding sites with a deep convolutional neural network. Journal of Theoretical Biology, 496, 110278.Google Scholar
Warner, KD, Hajdin, CE and Weeks, KM (2018) Principles for targeting RNA with drug-like small molecules. Nature Reviews Drug Discovery 17, 547558.Google Scholar
Wass, MN, Kelley, LA and Sternberg, MJE (2010) 3DLigandSite: predicting ligand-binding sites using similar structures. Nucleic Acids Research 38, W469W473.Google Scholar
Wei, J, et al. (2022) Protein–RNA interaction prediction with deep learning: structure matters. Briefings in Bioinformatics 23, bbab540.Google Scholar
Wei, L and Altman, RB (1998) Recognizing protein binding sites using statistical descriptions of their 3D environments. In Pacific Symposium on Biocomputing, pp. 497508.Google Scholar
Wei, L and Altman, RB (2003) Recognizing complex, asymmetric functional sites in protein structures using a Bayesian scoring function. Journal of Bioinformatics and Computational Biology 1, 119138.Google Scholar
Weisel, M, Proschak, E and Schneider, G (2007) PocketPicker: analysis of ligand binding-sites with shape descriptors. Chemistry Central Journal 1, 117.Google Scholar
Wiegreffe, S and Pinter, Y (2019) Attention is not not explanation. arXiv preprint. arXiv:1908.04626.Google Scholar
Wilson, CA, Kreychman, J and Gerstein, M (2000) Assessing annotation transfer for genomics: quantifying the relations between protein sequence, structure and function through traditional and probabilistic scores. Journal of Molecular Biology 297, 233249.Google Scholar
Wood, MJ and Hirst, JD (2005) Protein secondary structure prediction with dihedral angles. Proteins: Structure, Function, and Bioinformatics 59, 476481.Google Scholar
Xia, CQ, Pan, X and Shen, HB (2020) Protein–ligand binding residue prediction enhancement through hybrid deep heterogeneous learning of sequence and structure data. Bioinformatics 36, 30183027.Google Scholar
Xia, Y, Pan, X and Shen, HB (2023) LigBind: identifying binding residues for over 1000 ligands with relation-aware graph neural networks. Journal of Molecular Biology 435, 168091.Google Scholar
Xia, Y, et al. (2022) BindWeb: a web server for ligand binding residue and pocket prediction from protein structures. Protein Science 31, e4462.Google Scholar
Xia, Y, et al. (2021) GraphBind: protein structural context embedded rules learned by hierarchical graph neural networks for recognizing nucleic-acid-binding residues. Nucleic Acids Research 49, e51.Google Scholar
Xie, L and Bourne, PE (2007) A robust and efficient algorithm for the shape description of protein structures and its application in predicting ligand binding sites. BMC Bioinformatics 8, 113.Google Scholar
Xie, ZR and Hwang, MJ (2012) Ligand-binding site prediction using ligand-interacting and binding site-enriched protein triangles. Bioinformatics 28, 15791585.Google Scholar
Xue, B, et al. (2008) Real-value prediction of backbone torsion angles. Proteins: Structure, Function, and Bioinformatics 72, 427433.Google Scholar
Yaffe, E, et al. (2008) MolAxis: a server for identification of channels in macromolecules. Nucleic Acids Research 36, W210W215.Google Scholar
Yan, C and Zou, X (2014) Predicting peptide binding sites on protein surfaces by clustering chemical interactions. Journal of Computational Chemistry 36, 4961.Google Scholar
Yan, R, et al. (2019) Prediction of zinc-binding sites using multiple sequence profiles and machine learning methods. Molecular Omics 15, 205215.Google Scholar
Yan, X, et al. (2022) PointSite: a point cloud segmentation tool for identification of protein ligand binding atoms. Journal of Chemical Information and Modeling 62, 28352845.Google Scholar
Yang, J, Roy, A and Zhang, Y (2012) BioLiP: a semi-manually curated database for biologically relevant ligand–protein interactions. Nucleic Acids Research 41, D1096D1103.Google Scholar
Yang, J, Roy, A and Zhang, Y (2013) Protein–ligand binding site recognition using complementary binding-specific substructure comparison and sequence profile alignment. Bioinformatics 29, 25882595.Google Scholar
Yao, H, et al. (2003) An accurate, sensitive, and scalable method to identify functional sites in protein structures. Journal of Molecular Biology 326, 255261.Google Scholar
Yaseen, A and Li, Y (2014) Context-based features enhance protein secondary structure prediction accuracy. Journal of Chemical Information and Modeling 54, 9921002.Google Scholar
Yu, AM, Choi, YH and Tu, MJ (2020) RNA drugs and RNA targets for small molecules: principles, progress, and challenges. Pharmacological Reviews 72, 862898.Google Scholar
Yu, DJ, et al. (2015) Constructing query-driven dynamic machine learning model with application to protein-ligand binding sites prediction. IEEE Transactions on Nanobioscience 14, 4558.Google Scholar
Yu, DJ, et al. (2013) Designing template-free predictor for targeting protein-ligand binding sites with classifier ensemble and spatial clustering. IEEE/ACM Transactions on Computational Biology and Bioinformatics 10, 9941008.Google Scholar
Yu, J, et al. (2010) Roll: a new algorithm for the detection of protein pockets and cavities with a rolling probe sphere. Bioinformatics 26, 4652.Google Scholar
Yuan, Q, et al. (2022a) AlphaFold2-aware protein–DNA binding site prediction using graph transformer. Briefings in Bioinformatics 23, bbab564.Google Scholar
Yuan, Q, et al. (2022b) Alignment-free metal ion-binding site prediction from protein sequence through pretrained language model and multi-task learning. Briefings in Bioinformatics 23, bbac444.Google Scholar
Yuan, Z and Huang, B (2004) Prediction of protein accessible surface areas by support vector regression. Proteins: Structure, Function, and Bioinformatics 57, 558564.Google Scholar
Zaucha, J, et al. (2020) Deep learning model predicts water interaction sites on the surface of proteins using limited-resolution data. Chemical Communications 56, 1545415457.Google Scholar
Zeng, P and Cui, Q (2016) Rsite2: an efficient computational method to predict the functional sites of noncoding RNAs. Scientific Reports 6, 19016.Google Scholar
Zeng, P, et al. (2015) Rsite: a computational method to identify the functional sites of noncoding RNAs. Scientific Reports 5, 9179.Google Scholar
Zhan, ZH, et al. (2018) Accurate prediction of ncRNA-protein interactions from the integration of sequence and evolutionary information. Frontiers in Genetics 9, 458.Google Scholar
Zhang, C, et al. (2024) BioLiP2: an updated structure database for biologically relevant ligand–protein interactions. Nucleic Acids Research 52, D404D412.Google Scholar
Zhang, S and Xie, L (2023) Protein language model-powered 3D ligand binding site prediction from protein sequence. In NeurIPS 2023 AI for Science Workshop.Google Scholar
Zhang, W, Pei, J and Lai, L (2017) Statistical analysis and prediction of covalent ligand targeted cysteine residues. Journal of Chemical Information and Modeling 57, 14531460.Google Scholar
Zhang, Y, et al. (2023) EquiPocket: an E (3)-equivariant geometric graph neural network for ligand binding site prediction. arXiv Preprint. arXiv:2302.12177.Google Scholar
Zhang, Y and Skolnick, J (2005) TM-align: a protein structure alignment algorithm based on the TM-score. Nucleic Acids Research 33, 23022309.Google Scholar
Zhang, Y, et al. (2016) Identification of covalent binding sites targeting cysteines based on computational approaches. Molecular Pharmaceutics 13, 31063118.Google Scholar
Zhang, Z, et al. (2011) Identification of cavities on protein surface using multiple computational approaches for drug binding site prediction. Bioinformatics 27, 20832088.Google Scholar
Zhao, J, Cao, Y and Zhang, L (2020) Exploring the computational methods for protein-ligand binding site prediction. Computational and Structural Biotechnology Journal 18, 417426.Google Scholar
Zhao, W, et al. (2011) Structure-based de novo prediction of zinc-binding sites in proteins of unknown function. Bioinformatics 27, 12621268.Google Scholar
Zhao, X, Zhang, Y and Du, X (2022) DFpin: Deep learning–based protein-binding site prediction with feature-based non-redundancy from RNA level. Computers in Biology and Medicine 142, 105216.Google Scholar
Zhao, Y, et al. (2023) Identification of metal ion-binding sites in RNA structures using deep learning method. Briefings in Bioinformatics 24(2), bbad049.Google Scholar
Zhao, Z, et al. (2017) Determining cysteines available for covalent inhibition across the human kinome. Journal of Medicinal Chemistry 60, 28792889.Google Scholar
Zhao, Z, Peng, Z and Yang, J (2018) Improving sequence-based prediction of protein–peptide binding residues by introducing intrinsic disorder and a consensus method. Journal of Chemical Information and Modeling 58, 14591468.Google Scholar
Zhao, Z, Xu, Y and Zhao, Y (2019) Sxgbsite: Prediction of protein–ligand binding sites using sequence information and extreme gradient boosting. Genes 10, 965.Google Scholar
Zheng, C, et al. (2012) An integrative computational framework based on a two-step random forest algorithm improves prediction of zinc-binding sites in proteins. PLoS One 7, e49716.Google Scholar
Zheng, H, et al. (2024) PinMyMetal: A hybrid learning system to accurately model metal binding sites in macromolecules. Research Square, rs.3.rs–3908734.Google Scholar
Zhou, Y and Chen, SJ (2022) Graph deep learning locates magnesium ions in RNA. QRB Discovery 3, e20.Google Scholar
Zhu, C, et al. (2023) GAPS: Geometric attention-based networks for peptide binding sites identification by the transfer learning approach. bioRxiv, 2023–12.Google Scholar
Zhu, H and Pisabarro, MT (2011) MSPocket: An orientation-independent algorithm for the detection of ligand binding pockets. Bioinformatics 27(3) 351358.Google Scholar
Zhu, Y and Yu, DJ (2023) ULDNA: Integrating unsupervised multi-source language models with LSTM-attention network for protein-DNA binding site prediction. bioRxiv, 2023–05.Google Scholar
Figure 0

Table 1. List of methods for prediction of protein–small molecule binding sites

Figure 1

Figure 1. Schematic presentation of the sequence-based methods. The top part demonstrates the pipeline for a template-based approach: the target sequence is aligned against a database of template sequences with known binding residues, and the output binding residues are defined by the consensus score from the alignment. The bottom part demonstrates the pipeline for ML or DL methods. First, the feature vectors (e.g., sequence or physicochemical properties) or the embeddings (e.g., using language models) are calculated. Then, a method uses a moving window across the sequence and feeds feature vectors for each position into an ML or DL model outputting a binding score for each position, or utilizing a larger DL model to get binding scores for each position simultaneously.

Figure 2

Figure 2. Schematic presentation of the structure template-based methods. In the first stage, the target is screened against a database of template structures with known binding sites. In the second stage, the output prediction is obtained based on the most similar template structures with respect to the target.

Figure 3

Figure 3. Schematic overview of geometric methods for binding site detection. (a) Generation of occupancy grid and calculation of the fraction of directions enclosed by the target macromolecule for each empty grid point (used, for example, in POCKET (Levitt and Banaszak, 1992), LIGSITE (Hendlich et al., 1997), PocketPocker (Weisel et al., 2007), SiteMap (Halgren, 2009), CAVIAR (Marchand et al., 2021)). (b) Rolling of spheres with two different radii around the target macromolecule. The spheres with a larger radius remove the smaller ones. The remaining small spheres are clustered to get final predictions (used, for example, in APROPOS (Peters et al., 1996), PHECOM (Kawabata and Go, 2007), (Masuya and Doi, 1995), GHECOM (Kawabata, 2010), and POCASA (Yu et al., 2010)). (c) The addition-removal algorithm, is used in Delaney (1992), Kleywegt and Jones (1994), and Brady and Stouten (2000). Each step consists of adding and removing the surface-exposed points until the convergence. The target macromolecule is represented with a lilac surface, and grid points and probe spheres are shown with circles.

Figure 4

Figure 4. Schematic presentation of the energy probe-based methods. (a) Different probes (shown as red, blue, and green circles) are placed on a 3D grid around the target macromolecule (shown as a lilac surface) and their interaction energies with the target’s atoms are calculated. (b) The probes corresponding to the high-energy values are filtered out. (c) The remaining probes are clustered. (d) The filtering procedure is applied to remove non-relevant clusters.

Figure 5

Figure 5. Schematic presentation of the machine learning-based methods. On the top, the target structure is represented as a surface, and feature vectors are calculated for the surface points. On the bottom, feature vectors are calculated for the target’s residues or atoms. Then, an ML classifier predicts the binding scores for the points, residues, or atoms, based on the input feature vectors. Finally, the output predictions are filtered by a score threshold and clustered.

Figure 6

Figure 6. Schematic presentation of the DL-based methods. Most of the methods utilize graph-based or voxel grid representations of the target macromolecular structure. Then, they sample either sub-graphs or sub-grids around the structure and classify their centers as belonging to the binding site or not. Alternatively, they use segmentation models to operate with the full graph or grid.

Figure 7

Table 2. List of methods for prediction of protein–peptide binding sites

Figure 8

Table 3. Performance of protein–peptide binding site detection methods on test benchmarks retrieved from Kozlovskii and Popov (2021a), Abdin et al. (2022), and Fang et al. (2023)

Figure 9

Table 4. List of methods for prediction of nucleic acid–small molecule binding sites

Figure 10

Table 5. List of methods for prediction of protein–ion binding sites

Supplementary material: File

Kozlovskii and Popov supplementary material

Kozlovskii and Popov supplementary material
Download Kozlovskii and Popov supplementary material(File)
File 128.9 KB