Next Article in Journal
Microglial Autophagy and Mitophagy in Ischemic Stroke: From Dual Roles to Therapeutic Modulation
Previous Article in Journal
Hepatocyte-Specific Transcriptional Responses to Liver-Targeted Delivery of a Soluble Epoxide Hydrolase Inhibitor in a Mouse Model of Alcohol-Associated Liver Disease
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Role of AI-Driven De Novo Protein Design in the Exploration of the Protein Functional Universe

College of Science, National University of Defense Technology, Changsha 410073, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Biology 2025, 14(9), 1268; https://doi.org/10.3390/biology14091268
Submission received: 6 August 2025 / Revised: 2 September 2025 / Accepted: 3 September 2025 / Published: 15 September 2025

Simple Summary

Proteins are the molecular machines of life, essential for countless processes from building cellular structures to fighting disease. Nature has created a stunning array of these molecules, but this diversity is merely a glimpse of what is theoretically possible. This vast, untapped potential holds promise for solutions to some of our biggest challenges, such as cleaning up pollution, curing diseases, and creating new materials. Exploring these possibilities by experiments alone is impossibly slow and expensive. This review explains how artificial intelligence (AI) is changing the game. AI now allows us to design entirely new proteins de novo on a computer, predicting how they will fold and function. This powerful approach is yielding breakthroughs across biotechnology at an unprecedented pace. As AI continues to evolve, it promises to unlock a new era of biological engineering, providing custom-made protein tools for advances in medicine, agriculture, and green technology.

Abstract

The extraordinary diversity of protein sequences and structures gives rise to a vast protein functional universe with extensive biotechnological potential. Nevertheless, this universe remains largely unexplored, constrained by the limitations of natural evolution and conventional protein engineering. Substantial evidence further indicates that the known natural fold space is approaching saturation, with novel folds rarely emerging. AI-driven de novo protein design is overcoming these constraints by enabling the computational creation of proteins with customized folds and functions. This review systematically surveys the rapidly advancing field of AI-based de novo protein design, reviewing current methodologies and examining how cutting-edge computational frameworks accelerate discovery through three complementary vectors: (1) exploring novel folds and topologies; (2) designing functional sites de novo; (3) exploring sequence–structure–function landscapes. We highlight key applications across therapeutic, catalytic, and synthetic biology and discuss the persistent challenges. By fusing recent progress and the existing limitations, this review outlines how AI is not only accelerating the exploration of the protein functional universe but also fundamentally expanding the possibilities within protein engineering, paving the way for bespoke biomolecules with tailored functionalities.

1. Introduction

Artificial intelligence (AI) is causing a paradigm shift across numerous fields within biology and medicine [1,2]. In particular, AI-driven approaches are increasingly being harnessed to tackle one of the most fundamental challenges in protein engineering: the exploration and design of functional proteins [3]. Proteins are central to virtually all biological processes, yet the vast majority of possible protein sequences and structures remain unexplored, constrained by evolutionary history and by experimental throughput.
Conventional protein engineering, while yielding remarkable successes, is inherently limited by its dependence on existing biological templates. Current methods often fail to access novel functional regions of the protein universe that lie beyond natural evolutionary pathways. Moreover, they typically require experimental screening of large variant libraries, a process that is labor-intensive, costly, and ultimately confined to incremental improvements within well-explored neighborhoods of the sequence–structure space [4,5]. Consequently, systematic exploration of the uncharted territories within the protein functional universe demands a disruptive, more pioneering approach.
Recent advances in AI, however, have begun to transcend these limitations. By integrating generative models, structure prediction tools, and iterative experimental validation, AI-driven de novo protein design offers a powerful framework for systematically exploring and engineering proteins with customized functions. This approach leverages known statistical patterns from vast biological datasets to establish high-dimensional mappings between sequence, structure, and function, enabling the rapid generation of novel, stable, and functional proteins. This approach empowers researchers to directly explore regions of the functional landscape that natural evolution has not sampled, thereby accelerating the discovery of novel biomolecules and opening new avenues for addressing global challenges in health, sustainability, and biotechnology.

2. The Vast but Evolutionarily Constrained Protein Functional Universe

Proteins drive critical cellular processes, including enzymatic catalysis [6,7], signal transduction [8,9,10], molecular recognition [11,12], structural support [13,14,15], and immune defense [16,17], highlighting their extensive functional diversity. This breadth of activity constitutes the “protein functional universe”: a theoretical space encompassing all possible protein sequences and structures and the biological activities they can perform, governed by the complex mapping between sequence space and structure space. This conceptual universe includes not only the folds and functions observed in nature but also every other stable protein fold and corresponding activity that could in principle exist (Figure 1). Systematically probing the unexplored functional universe could reveal novel enzymes, binding activities, or molecular machines that do not exist in the natural world, opening up new solutions in biotechnology, medicine, and synthetic biology [18].
Researchers exploring this universe, however, face two fundamental challenges. The first is the problem of combinatorial explosion, stemming from its unimaginable scale. The sequence → structure → function paradigm—the idea that a protein’s amino acid sequence encodes its three-dimensional fold, which in turn largely determines its biological function—is a longstanding central tenet of molecular biology [19,20]. For perspective, a mere 100-residue protein theoretically permits 20100 (≈1.27 × 10130) possible amino acid arrangements, exceeding the estimated number of atoms in the observable universe (~1080) by more than fifty orders of magnitude [21]. This renders the prior probability that a random sequence will fold stably and display useful activity vanishingly small. Given that experimental investigations are restricted and costly, unguided experimental screening is profoundly inefficient [22].
The second challenge arises from the constraints of natural evolution. Despite their functional richness, natural proteins are products of evolutionary pressures for biological fitness, and are not necessarily optimized as versatile tools for human utility. This so-called “evolutionary myopia” tends to lead to proteins that are optimized for survival in specific niches, potentially limiting properties such as stability, specificity, or suitability for industrial conditions. Comparative analyses suggest that known protein functions represent only a tiny subset of the diversity nature can produce [23]. Furthermore, the current evidence suggests that the known protein fold space may be nearing saturation, with recent functional innovations predominantly arising from domain rearrangements [24,25]. Integrating predicted structures reveals how environmental pressures profoundly shape natural protein diversity: evolutionary forces overwhelmingly favor domain recombination over de novo emergence of structural motifs or folds. Coupled with the accumulation of gradual genetic variation in natural populations, this selective paradigm reinforces an evolutionary trajectory that diversifies proteomes through reorganization and repurposing—thereby constraining the exploration of genuinely novel sequences and structures [24].
Despite considerable advances in mapping the existing sequence and structure spaces—exemplified by resources such as the MGnify Protein Database [26] (cataloging nearly 2.4 billion non-redundant sequences) and the Profluent Protein Atlas v1 [27] (encompassing over 3.4 billion full-length proteins)—alongside expanding structural repositories (including ~214 million models in the AlphaFold Protein Structure Database [28] and about 600 million predicted structures in the ESM Metagenomic Atlas [29], these datasets constitute an infinitesimally small portion within the theoretical protein functional space. Public datasets are also biased by evolutionary history and assay ability, which tends to channel data-driven methods toward well-explored regions of the sequence–structure space. Thus, vast regions of the sequence–structure space remain inaccessible, highlighting the need for a new paradigm that integrates advanced computation with experimental validation to unlock the immense latent functional potential within the uncharted protein universe [22,30,31].

3. Beyond Evolutionary Boundaries: Exploring the Functional Universe

This imperative for a new approach is underscored by the intrinsic limitations of conventional protein engineering strategies. While methods such as directed evolution have proven powerful for optimizing existing proteins [4,5], their workflow inherently constrains exploration. By necessitating a natural protein as a starting point, they remain tethered to evolutionary pressure and the requirements for the construction and experimental screening of immense variant libraries through iterative cycles of mutation and selection. This is not just labor-intensive and costly; more fundamentally, it confines discovery to the immediate “functional neighborhood” of the parent scaffold, performing a local search within the vastness of the protein functional universe. Consequently, these approaches are structurally biased and ill-equipped to access genuinely novel functional regions that lie beyond the boundaries of natural evolutionary pathways [32,33,34].
De novo protein design aims to transcend these limits by designing proteins from first principles to meet specified structural or functional objectives, rather than by modifying existing scaffolds [32]. This approach can, in principle, produce wholly new folds, bespoke active sites, and modular components with engineered properties (tunability, controllability, and modularity), offering a systematic route to functions that natural evolution has not explored [33,35,36,37]. This fundamental paradigm shift frees protein engineering from its historical reliance on natural templates and removes their inherent evolutionary limitations; exploration of the functional protein universe transitions from empirical trial-and-error explorations to systematic rational design, vastly expanding our access to the previously unimaginable diversity of biologically active folds and functions.
However, de novo design itself faces fundamental challenges. Functional proteins occupy an astronomically small subset of the possible sequence–structure space, and as the polypeptide length increases, the conformational freedom, synthesis cost, and assay complexity grow exponentially. The precise mapping between protein sequence and structural attributes and functional phenotypes defines the canonical “fitness landscape” [38]. In practical de novo design, searches must therefore be restricted to computationally tractable subspaces, and precise atomic constraints should be imposed for functional geometries to ensure orthogonality to the host biology and to avoid off-target interactions.

3.1. The AI-Driven Paradigm Shift in Protein Engineering

Historically, de novo protein design has relied heavily on empirical methods and physics-based modeling [39,40,41,42]. Rosetta is a typical example, operating on Anfinsen’s hypothesis that proteins fold into their lowest-energy state [43]. By taking a blueprint of secondary structure elements (e.g., α-helices, β-sheets), Rosetta employs fragment assembly and force-field energy minimization to fold proteins in silico and stitches together short peptide fragments from known proteins and performs conformational sampling (e.g., Monte Carlo with simulated annealing). The lowest-energy conformations under its force field are then selected as candidate designs [39]. In 2003, Kuhlman et al. used Rosetta to create Top7, a 93-residue protein with a novel fold not observed in nature [44]. Subsequent work extended Rosetta to design enzyme active sites [45,46] and drug-binding scaffolds [47,48,49], showcasing its versatility in rational protein engineering.
Nevertheless, these physics-based methodologies exhibit inherent drawbacks. Firstly, the underlying force fields retain an approximate character: despite contemporary refinements, accurately computing a protein’s comprehensive energy landscape remains challenging, particularly when incorporating elaborate side-chain packing and solvent effects. Even marginal inaccuracies in energy estimates can yield designs that misfold or fail to achieve intended functionality in vitro. Secondly, the associated computational expense is considerable: exhaustive sampling of even a constrained fraction of the sequence and structure space is frequently infeasible [32]. These constraints are acutely observable for large or structurally complex proteins, limiting both throughput and the practical exploration of distant regions of the protein functional universe.
In response, modern AI-augmented strategies have emerged to complement and extend physics-based design [50,51]. Machine learning (ML) models trained on large-scale biological datasets that can establish high-dimensional mappings learned directly from sequence–structure–function can capture intricate interdependent fitness relationships [52,53]. These learned priors accelerate sampling and scoring, enable scalable generative workflows, and guide experiments by prioritizing candidates with higher a priori chances of success. This capability has been fueled by decades of experimental progress—high-throughput DNA sequencing that has exponentially expanded sequence repositories; structural biology innovations such as cryo-electron microscopy (cryo-EM) that uncover atomic-resolution structures [54]; and deep mutational scanning enabling systematic functional profiling [55]. AI-driven de novo protein design commonly integrates hybrid pipelines—generating backbones or sequences, rescoring with structure predictors and energy models, and iteratively incorporating experimental feedback—to make it both more efficient and more expansive in scope. These strategies do not replace experiments, but they transform them from brute-force searches into focused, data-efficient campaigns that stand a realistic chance of revealing novel, functional proteins.
By partitioning the combinatorial search into tractable subspaces and enforcing target-specific constraints, these pipelines enable focused searches for foldable, functional proteins while also promoting orthogonality with the host biology through designing interaction sites specific to the intended targets through rapid in silico screening (e.g., docking/binding [56,57], epitope and immune-response [58], and solubility prediction [59]).

3.2. Main Paradigms of AI-Driven De Novo Protein Design

AI-driven approaches provide a rational, systematic framework for de novo protein design, which can be broadly classified into three main paradigms [33] (Figure 2):
  • Two-Stage Generative Design
Modern AI-driven de novo design typically involves two stages: First, generative models such as diffusion networks construct novel protein backbone geometries or scaffold topologies tailored to specific functional or structural requirements. Next, sequence-design algorithms assign amino acid identities predicted to fold stably into these backbones and exhibit the desired activity.
2.
Sequence-Guided Language Methods
Inspired by breakthroughs in natural language processing, large pre-trained protein language models trained solely on sequence data have opened new avenues for de novo protein design by generating and evaluating candidate proteins without explicit structural inputs.
3.
Sequence–Structure Co-Guided Methods.
Moreover, based on the fact that protein function arises from intricate sequence–structure interdependencies, recent methods jointly model both structure generation and sequence optimization. These co-design frameworks improve foldability and functional accuracy, leading to higher-confidence de novo protein design.
Collectively, these methodologies advance de novo protein design toward on-demand biomolecular engineering, which would transform drug discovery, synthetic biology, and biomaterials science. AI-driven frameworks transcend natural evolutionary constraints, enabling the de novo design of proteins with novel functions inaccessible via natural evolution. Unified can approaches map multidimensional functional landscapes—from synthetic organelles to biological quantum sensors—effectively rendering the protein universe a programmable platform for medicine, energy, and synthetic biology.

4. The AI Toolbox for De Novo Protein Design

In practice, AI-driven de novo protein design typically couples generative models with predictive models in an iterative “digital evolution” loop [18]. Predictive tools originally developed for structural determination, such as RoseTTAFold All-Atom [60] and AlphaFold3 (AF3) [56], have become indispensable components of the AI-driven de novo protein design toolbox. These predictive tools provide atomic-precision analyses of target binding sites and employ reverse-engineering of active-site relationships to optimize dynamic stability and functional compatibility [61]. Together with generative models, sequence-design algorithms, and predictive structure tools form a powerful, integrated AI-driven toolbox for de novo protein design (Figure 3).
For readers new to computational protein design, Table 1 summarizes the practical workflows for each toolkit; each row summarizes the goal, the typical inputs, and the typical outputs that beginners can follow as an actionable starting point. The following section provides a concise overview of advancements in the AI-driven de novo protein design toolbox, categorizing its components into five functional classes (Table 2): (i) Protein Structure Prediction (Figure 3A); (ii) De novo Backbone Generation (Figure 3B); (iii) “Fixed-backbone” Sequence Design (Figure 3C); (iv) Sequence Generation (Figure 3D); and (v) Sequence–Structure co-design (Figure 3E).

4.1. Protein Structure Prediction

Accurately predicting the three-dimensional (3D) structure of a protein from its primary amino acid sequence has been a longstanding challenge in computational and structural biology [62]. Traditional experimental methods such as nuclear magnetic resonance (NMR) [63], X-ray crystallography [64], and cryo-electron microscopy (cryo-EM) [54] are limited by their low throughput and challenges in capturing dynamic information [65]. AI has transformed this field: at the 14th Critical Assessment of Protein Structure Prediction (CASP14) in 2020, AlphaFold2 (AF2) achieved an atomic-level prediction accuracy comparable to experimental crystallographic resolutions [66,67,68]. AF2 employs coevolutionary signals from multiple sequence alignments (MSAs) within a two-stage framework: the Evoformer network refines evolutionary and geometric constraints, then the structure module assembles backbone coordinates [67]. This breakthrough effectively resolved a decades-long protein-folding enigma and ushered in a transformative era for structural biology. Inspired by AF2, many other groups have developed high-performance prediction models. The Baker laboratory developed RoseTTAFold [69], which uses a three-track network to simultaneously process MSAs, residue pair distances, and 3D coordinates. RoseTTAFold achieves a comparable accuracy to AF2 while offering great flexibility in some settings. Variants such as ColabFold [70] provide an optimized pipeline that reduces computational costs with minimal loss in accuracy. OpenFold [71] provides an open-source reimplementation of AF2 with a comparable accuracy but a faster speed and lower memory usage, while SPIRED [72] can address high-throughput demands by delivering significant speed improvements and reduced computational costs while achieving a performance comparable to OmegaFold [73]. AFsample2 [74] introduces stochastic masking of MSA columns to attenuate co-evolutionary constraints, thereby enhancing the structural diversity in models generated by AF2. More recently, Zheng et al. developed D-I-TASSER [75], which combines deep learning with classical physics-based folding simulations to tackle multidomain proteins. Benchmarking and CASP15 results show that D-I-TASSER outperforms both AF2 and AF3 on single and multidomain targets. Despite these advances, the models mentioned above primarily depend on MSAs to extract coevolutionary signals and guide structure predictions, rendering them ineffective for rapidly evolving proteins or wholly synthetic sequences that lack sufficient homologous data. To address this limitation, pre-trained protein language models (PLMs) like OmegaFold [73] and ESMfold [29] learn evolutionary information patterns directly from large quantities of raw protein sequences, removing the requirement for MSAs. These PLM-based models enable faster, single-sequence structure inference and are particularly valuable for orphan, rapidly evolving genes or synthetic sequences. However, benchmarking studies reveal a clear trade-off: PLMs can be competitive in some low-MSA-depth scenarios, even approaching or exceeding MSA-based performance, but when deep MSAs or strong template information are available, MSA-dependent models generally exhibit a higher accuracy [37,76]. Practically, de novo design workflows therefore often feature a tiered strategy: PLM/MSA-free predictors are used for a high-throughput, first-pass screening of novel sequences and orphan designs, and MSA-dependent models (or rescoring with high-confidence structure models) are applied when homologous data or templates exist and atomic detail is required to increase confidence in foldability and function [77].
Biological function is frequently mediated by biomolecular complexes or macromolecular assemblies consisting of proteins, nucleic acids, and small molecules. Consequently, beyond predicting the structures of single proteins, AI-driven models are increasingly capable of predicting complex multi-component assemblies [78]. AlphaFold-Multimer [79] extends the AF2 architecture to accurately predict the structures of multi-chain protein complexes. RoseTTAFoldNA [80] further broadens this capability by expanding the scope of structural prediction to include interactions between proteins and nucleic acids. RoseTTAFold All-Atom [60] and AF3 [56] have enhanced the capabilities of modeling comprehensive biological molecular systems through unified frameworks, enabling accurate structure predictions for complexes containing proteins, small molecules, nucleic acids, ions, and post-translational modifications. These breakthroughs provide a more comprehensive understanding of biological activities in diverse macromolecular systems. Building on AF3, Chai-1 [81] incorporates protein language model embeddings and advanced structural constraints, yielding greater flexibility and precision in predicting biomolecular assemblies. Boltz-1 [82] enhances the core AF3 architecture with targeted optimizations to boost accuracy and computational efficiency for multi-component assemblies. Building upon this framework, Boltz-2 [83] not only predicts the structure of protein–biomolecule interactions but also quantitatively estimates their binding affinity.
Advances in AI-driven structure prediction have driven the progression from accurate models of single protein structures to complex multi-component assemblies and dynamic conformational ensembles [84,85]. When integrated with downstream design algorithms, they form a unified workflow for de novo engineering: providing in silico validation and enabling rapid virtual screening of designed sequences and complexes to evaluate structural integrity, binding site fidelity, and interface compatibility, and thereby significantly reducing the experimental burden. Moreover, by combining high-throughput structure prediction with clustering methods such as Foldseek [86], researchers can systematically survey billions of candidate models, differentiate genuinely new structures from those cataloged in the Protein Data Bank (PDB), and validate new protein architectures with potential functional innovations across structure space [25].

4.2. De Novo Backbone Generation

In de novo protein design, generating novel folds that transcend evolutionary and natural templates is critical for accessing the protein structure space. While structure prediction models can accurately model proteins given a sequence, they are not designed to sample diverse backbone ensembles under tailored biophysical constraints [67]. Conversely, physics-based design engines like Rosetta rely on global energy minimization and therefore inherently limit both the scaffold diversity and design flexibility [39].
Diffusion models overcome these limitations by learning complex structural distributions directly from data, enabling the generation of highly diverse and precise backbone geometries while allowing the incorporation of custom constraints during training [87]. The Baker laboratory’s RFdiffusion [88] (a denoising diffusion probabilistic model fine-tuned from RoseTTAFold) has emerged as a benchmark for de novo backbone generation, supporting both unconditional scaffold generation and topology-constrained design across diverse applications. RFdiffusion [88] has demonstrated efficacy in designing high-affinity binders, assembling symmetric oligomers, and constructing scaffolds and functional motifs. Building upon this core framework, subsequent extensions of RFdiffusion address a spectrum of design challenges: RFdiffusion All-Atom [60] integrates atomic-level geometric constraints and explicit small-molecule binding specifications to generate protein pockets with tailored ligand affinities; RFantibody [89] adapts the model for de novo nanobody scaffold design; RFdiffusion-IDP Binder [90] and RFdiffusion β-Strand Binder [91], respectively, generate binders targeting intrinsically disordered regions and complement irregular β-sheet features (e.g., twists, bends, and bulges); RFpeptides [92] enhances the generation of macrocyclic peptides with defined topologies, and most recently, RFdiffusion2 [93] enables the direct scaffolding of atomically defined enzyme active sites without the need to pre-specify residue indices or enumerate side-chain rotamers, representing a significant advance in computational enzyme engineering. Furthermore, FrameDiff [94] implements diffusion-based generation without reliance on pretrained prediction networks, thereby potentially avoiding biases toward known natural structures.
Chroma [95] features a programmable generative process that integrates biophysically informed diffusion mechanisms with quasi-linear graph neural networks, enabling constraint-guided protein structure generation based on geometric constraints, symmetry, topology, and semantic prompts. ROS [96] employs a hallucination-based protein design strategy that operates within a relaxed sequence space [97,98], outperforming existing methods such as RFdiffusion in the design of large proteins (>600 amino acids). Flow-matching approaches eliminate the need to simulate stepwise diffusion during training, greatly reducing the computational cost for high-dimensional data and improving the training efficiency. Proteína [99] leverages flow matching for protein backbone generation and employs a non-equivariant Transformer architecture with 400 million parameters (five times larger than RFdiffusion) and can generate protein structures up to 800 amino acids in length.

4.3. “Fixed-Backbone” Sequence Design

After generating novel backbones, the core challenge in de novo protein design becomes inverse folding—identifying amino acid sequences that will reliably fold into a specified backbone while achieving the desired function. A standard metric for evaluating inverse-folding methods is the natural sequence recovery (NSR) rate, defined as the fraction of designed residues that match the native sequence of a naturally occurring protein with a similar fold. NSR is a widely used benchmark for assessing inverse folding models and has been widely used to compare methods. However, it has important limitations [100]. Inverse folding is inherently one-to-many: distinct, non-homologous sequences can stabilize the same backbone through alternative packing and compensatory substitutions, so strict recovery of a single “native” sequence can restrict novel yet biophysically valid solutions. Moreover, natural sequences encode evolutionary constraints (expression, regulation, promiscuous interactions, cellular context, immune selection, etc.) that may not align with specific engineering objectives (for example, enhanced thermostability or bespoke binding specificity); consequently, high NSR does not guarantee the intended functional phenotype. Most current inverse folding models also report structure-prediction metrics for the designed sequences (e.g., RMSD and pLDDT/pTM) and sequence novelty as a reference; we recommend, where possible, validating candidate sequences with task-relevant experimental assays (expression, solubility/stability, binding affinity, or deep mutational scanning).
ESM-IF [101] leverages a hybrid architecture combining protein language and structural modeling to generate proteins with substantially divergent sequences from natural evolutionary distributions. Furthermore, developed by the Baker laboratory, ProteinMPNN [102] has become a broadly adopted inverse-folding model. It implements a graph-based message-passing neural network (MPNN) architecture that includes a three-layer equivariant encoder to embed backbone atoms (N, C, O, Cα, Cβ) as a distance-weighted graph and a sequence-agnostic stochastic decoder to sample optimal amino acids at each position, producing diverse, topology-informed designs. It achieved a notable NSR of 52.4% for a test set of 402 monomer backbones (compared with Rosetta’s 32.9%) and has been experimentally validated, showing that its designed sequences reliably fold as intended [103,104]. LigandMPNN [105] extends the ProteinMPNN [102] framework and further incorporates non-protein ligands such as small molecules, nucleic acids, and metal ions into the design process, enabling optimization of protein ligand interfaces. Inspired by AlphaFold, CarbonDesign [106] employs an “Inverseformer” network that integrates multimodal constraints from structural features and ESM2-derived evolutionary embeddings. This architecture delivers breakthroughs in designing long-chain and novel proteins, outperforming established methods such as ProteinMPNN [102] and ESM-IF [101] across independent benchmarks, including CAMEO and CASP15, and de novo backbones generated by RFdiffusion [88]. CARBonARa [107] further enhances inverse folding by jointly modeling atomic coordinates and molecular environment constraints, significantly improving sequence-prediction accuracy in functional interface regions.
The capacity to engineer proteins with precise structural constraints is a prerequisite for the rational design of novel proteins with tailored functions. De novo backbone generation models overcome the fixed-topology limitations of classical design methods by systematically sampling the entire protein structure space, thereby uncovering previously inaccessible structures. In parallel, inverse folding algorithms identify amino acid sequences that not only adopt these specific backbones but also confer the desired functions. This synergistic, two-stage design strategy broadens the exploration of the sequence and structure space and, by extension, the functional universe, establishing an end-to-end pipeline from functional objective to tailored structure to sequence realization.

4.4. Sequence Generation

Recent advances in large language models (LLMs) have propelled the development of sequence generation models for de novo protein design, including ProtGPT2 [108], ProGen [109], ProGen2 [110], and TourSynbio [111]. These protein language models (PLMs) are pre-trained on millions of natural protein sequences, and treat each sequence as a “sentence” and each amino acid as a “token”. PLMs can generate novel sequences that conform to the learned biophysical and thermodynamic “grammar” of natural proteins, and can be fine-tuned for specific tasks such as enzyme or binder design [52]. However, the ability of PLMs to produce truly novel, broadly useful sequences is constrained by biases in their training data. Public sequence resources and curated databases are taxonomically and experimentally uneven (over-representing model organisms, pathogen- and human-related sequences, and protein families that are easy to assay), while structural and functional annotation lags far behind raw sequence growth. These sampling biases influence model likelihoods and sampling behavior, tending to push generation toward well-represented motifs and taxa rather than underexplored regions of the protein space [112,113]. To mitigate such artifacts, groups commonly augment or reweight the training corpus with metagenomic sequences and apply controlled fine-tuning on target design families; combined with structure- and energy-based rescoring and targeted experimental validation, these steps increase the likelihood that generated sequences are both novel and functional [109,114]. ProGen [109] is a 1.2 billion parameter neural network trained on 280 million sequences from over 19,000 annotated families that does not require structural data, while ProGen3 [27] scales this approach to 46 billion parameters pre-trained on 1.5 trillion amino-acid tokens drawn from 3.4 billion full-length proteins. Meanwhile, xTrimoPGLM [115] is a protein general language model that employs a dual objective training scheme using 100 billion parameters and 1 trillion tokens to jointly optimize protein understanding and design. It was pre-trained on 940 million unique protein sequences, comprising roughly 200 billion residues.
Protein sequences encode the blueprint for both protein structure and function [116]. Sequence generation models are trained on billion-scale protein databases, treating amino acid chains as a “language” whose grammar reflects the intrinsic evolutionary patterns and sequence features of protein structures. By leveraging this protein “grammar” to explore the sequence space and generate novel sequences, these models in turn drive the discovery and expansion of the protein functional universe.

4.5. Sequence–Structure Co-Design

Most de novo design methods typically generate sequences or structures separately, limiting their ability to capture the complex, bidirectional dependencies between them that govern protein folding functions [117]. Sequence–structure co-generation methods address this gap by jointly modeling both modalities, combining structure and sequence-driven insights to explore new regions of the functional universe. Several co-generation frameworks have recently emerged, including Multiflow [118], ProteinGenerator [119], ESM3 [120], and Pinal [121]. ProteinGenerator [119] is a RoseTTAFold-based sequence-space diffusion model that generates sequence–structure pairs through iterative denoising guided by specified sequence and structural attributes. ESM3 [120] is a frontier multimodal generative language model with 98 billion parameters, trained on 2.78 billion protein sequences and 771 billion unique tokens. It can jointly reason on protein sequences, structure, and function, yielding improved representation and generative evaluation across all three modalities. Pinal [121] employs a two-stage pipeline in which a natural language functional description is first converted into structural constraints and then conditioned sequences are generated on both the description and the resulting structure, achieving higher quality protein designs by operating within a smaller structural space.
By integrating sequence and structure within a unified generative framework, co-generation methods overcome the intrinsic limitations of two-stage generative design approaches; information flows unidirectionally—first through the generation of backbones and then through mapping them to sequences—hindering a full traversal of the sequence-structure-function landscape. Likewise, sequence generation methods neglect the critical structural constraints that underpin protein function. These separations fail to capture bidirectional dependencies and often lead to suboptimal designs. In contrast, sequence–structure co-generation enables coordinated exploration of the vast sequence–structure space, revealing previously inaccessible regions of the protein functional universe and directly yielding novel proteins with defined folds and tailored activities.
Table 2. Classification and Key-Attribute Comparison of AI-driven de novo Protein Design Tools.
Table 2. Classification and Key-Attribute Comparison of AI-driven de novo Protein Design Tools.
Model NameRelease DateExperimental ValidationModel
Description
AlgorithmsEfficiencyAdvantageLimitationIs the Code Publicly
Available?
Ref.
Protein Structure Prediction
Alphafold215 July 2021/The first model predicts protein structures with atomic accuracy.EvoformerLowAtomic accuracyResource-intensive; MSA-dependenthttps://github.com/google-deepmind/alphafold (accessed on 5 September 2025)[67]
RoseTTAFold19 August 2021/Accurately predicts protein structures and interactions.3D TransformerMiddleFast, flexible; accuracyLess accurate than AF2 on hard targetshttps://github.com/RosettaCommons/RoseTTAFold (accessed on 5 September 2025)[69]
ColabFold30 May 2022/Fast and easy for the prediction of protein structures and complexes.EvoformerHighFast, accessibleDepends on the AF2 back endhttps://github.com/sokrypton/ColabFold (accessed on 5 September 2025)[70]
OmegaFold2 July 2022/Predict orphan proteins and rapidly evolving antibodies.PLMHighFast, MSA-freeSlightly lower accuracy on some targetshttps://github.com/HeliXonProtein/OmegaFold (accessed on 5 September 2025)[73]
ESMFold16 March 2023/Predict protein structure with atomic precision using language models.PLMHighFast, MSA-freeLower atomic precision vs. AF2https://github.com/mit-ll/ESMFold (accessed on 5 September 2025)[29]
OpenFold14 May 2024/An open-source, trainable implementation of AF2.EvoformerMiddleOpen, reproducible AF2 implementationSimilar computational needs to AF2https://github.com/aqlaboratory/openfold (accessed on 5 September 2025)[71]
SPIRED27 August 2024/Enhances prediction speed, reduces training consumption.Unit-basedMiddleFaster training/inference; efficient designAccuracy below AF2https://github.com/Gonglab-THU/SPIRED-Fitness (accessed on 5 September 2025)[72]
AFsample25 March 2025/Expands the structural diversity of AF2’s generative models.Stochastic samplingLowProduces conformational ensemblesMuch higher computation for samplinghttps://github.com/iamysk/AFsample2 (accessed on 5 September 2025)[74]
D-I-TASSER23 May 2025/Predicts large multidomain protein structures.Deep and physicsLowGood for multi-domain proteinsSlow; template/MSA-dependenthttps://zhanggroup.org/D-I-TASSER/download/ (accessed on 5 September 2025)[75]
Predicted multimers structure
AlphaFold Multimer10 October 2021/Predicts the structure of protein complexes.EvoformerLowImproved multimer predictionsHigher computation requirementshttps://github.com/jcheongs/alphafold-multimer (accessed on 5 September 2025)[79]
RoseTTAFoldNA23 November 2023/Predicts the structures of protein-nucleic acid complexes.3D TransformerMiddlePredicts protein–nucleic acid complexesLimited NA training datahttps://github.com/uw-ipd/RoseTTAFold2NA (accessed on 5 September 2025)[80]
RoseTTAFold All-Atom7 March 2024/Predicts the structures of biomolecular assemblies containing proteins, nucleic acids, small molecules, metals, and chemical modifications.3D TransformerLowModels small molecules/ionsComputationally demanding, high memoryhttps://github.com/baker-laboratory/RoseTTAFold-All-Atom (accessed on 5 September 2025)[60]
AlphaFold38 May 2024/Predicts biomolecular complexes such as proteins, nucleic acids, small molecules, ions, and modified residues.EvoformerLowMultimolecule unified modelingLarge computational requirementshttps://github.com/google-deepmind/alphafold3 (accessed on 5 September 2025)[56]
Chai-111 October 2024/Predicts the structures of protein–ligand complexes and protein multimer.PLMHighAims MSA-free multimodal predictionPublic details limitedhttps://github.com/chaidiscovery/chai-lab (accessed on 5 September 2025)[81]
Boltz-120 November 2024/Open-source AF3-level precision prediction model.EvoformerHighOptimized AF3 architecturePublic details limitedhttps://github.com/jwohlwend/boltz (accessed on 5 September 2025)[82]
Boltz-26 June 2025/Simultaneous prediction of protein–small molecule complex structures and binding affinity.EvoformerHighAdds affinity estimation to the structurePublic details limitedhttps://github.com/jwohlwend/boltz (accessed on 5 September 2025)[83]
De novo protein backbone generation
RFdiffusion11 July 2023YesUnconditional/topology monomers; binders; symmetric oligomers; enzyme scaffolds; motif scaffolds.DiffusionMiddleVersatile backbone generationSampling is computationally allyintensivehttps://github.com/RosettaCommons/RFdiffusion (accessed on 5 September 2025)[88]
FrameDiff22 May 2023NoIndependent monomer generation with up to 500 amino acids without pretraining.DiffusionHighEfficient; no pretrained predictor neededNeeds more validationhttps://github.com/jasonkyuyim/se3_diffusion (accessed on 5 September 2025)
[94]
Chroma15 November 2023YesProgrammable protein generation via symmetry, shape, class, or text inputs.DiffusionHighEfficient, scalableLimited benchmarkshttps://github.com/generatebio/chroma (accessed on 5 September 2025)[95]
FoldingDiff5 February 2024NoUnconditionally generates highly realistic protein structures.DiffusionHighScales to long chainsIndirect side-chain generationhttps://github.com/microsoft/foldingdiff (accessed on 5 September 2025)[122]
RFdiffusion All-Atom7 March 2024YesLigand-guided de novo protein scaffold design.DiffusionMiddleAtomistic pockets and ligand designComputationally costlyhttps://github.com/baker-laboratory/rf_diffusion_all_atom (accessed on 5 September 2025)[60]
RSO24 October 2024YesRelaxed-sequence optimization enabling large-scale protein design without retraining.Hallucination-basedLowJoint sequence/structure optimizationLocal optimum riskhttps://github.com/sokrypton/ColabDesign (accessed on 5 September 2025)[96]
SCUBA-D21 November 2024YesUnconditional generation; generation based on sketch input; motif scaffolding.DiffusionMiddleSample novel foldsExperimental validation neededhttps://github.com/liuyf020419/SCUBA-D (accessed on 5 September 2025)[123]
Proteina2 March 2025NoUnconditional/class-conditional generation; motif scaffolding.Flow-matchingHighGenerates long chains (up to 800 aa)Needs side-chain stephttps://github.com/NVIDIA-Digital-Bio/proteina (accessed on 5 September 2025)[99]
RFdiffusion210 April 2025YesAtom-level active site scaffolding without residue indexing or rotamer sampling.DiffusionMiddleDirect active-site scaffoldingComputationally costlyhttps://github.com/RosettaCommons/RFdiffusion2 (accessed on 5 September 2025)[93]
ProtComposer6 March 2025NoEllipsoid-guided protein generation with customizable layouts.Flow-matchingHighConditional layout controlComplex implementationhttps://github.com/NVlabs/protcomposer (accessed on 5 September 2025)[124]
TopoDiff18 June 2025YesEnabling both unconditional and controllable diffusion-based protein generation.DiffusionHighTopology-controlled designPublic details sparsehttps://github.com/meneshail/TopoDiff/tree/main (accessed on 5 September 2025)[125]
‘Fixed-backbone’ sequence design
ESM-IF10 April 2022YesInverse folding (protein complexes, partially masked structures, binding interfaces, and multiple states).PLMMiddlePLM priors improve noveltyOnly backbone designhttps://github.com/facebookresearch/esm (accessed on 5 September 2025)[101]
ProteinMPNN15 September 2022YesInverse folding (monomers, cyclic oligomers, protein nanoparticles, and protein-protein interfaces).MPNNHighFast; strong inverse-folding performanceIgnores the ligand contexthttps://github.com/dauparas/ProteinMPNN (accessed on 5 September 2025)[102]
ProRefiner16 November 2023YesStructure-guided residue sequence inpainting with entropy-based global noise filtering.TransformerMiddleImproves model outputsTraining complexityhttps://github.com/veghen/ProRefiner (accessed on 5 September 2025)[126]
CarbonDesign23 May 2024NoInverse folding, zero-shot prediction of mutational effects on protein function.TransformerMiddleMultimodal constraint integrationNeeds broader benchmarkinghttps://github.com/carbon-design-system/carbon (accessed on 5 September 2025)[106]
CARBonAra25 July 2024YesDesigns protein sequences under the constraints of specific molecular interaction environments.TransformerMiddleHandles ligand/metal contextsTraining complexityhttps://github.com/LBM-EPFL/CARBonARa (accessed on 5 September 2025)[107]
LigandMPNN28 March 2025YesSimultaneously outputs ligand-binding sequences and sidechain conformations for interaction analysis.MPNNMiddleDesigns ligand interfacesRequires ligand coordinateshttps://github.com/dauparas/LigandMPNN (accessed on 5 September 2025)[105]
FAMPNN17 February 2025NoFull-atom protein sequence design.MPNNHighFull-atom sequence and sidechain outputLimited public detailshttps://github.com/richardshuai/fampnn (accessed on 5 September 2025)[127]
Methods generating sequences
ProtGPT227 July 2022NoHigh-throughput de novo protein sequence generation.TransformerHighFast generationLimited controlhttps://github.com/TeletcheaLab/protGPT2 (accessed on 5 September 2025)[108]
ProGen26 January 2023YesGenerates functional artificial proteins across families based on a conditional language model.TransformerMiddleConditional generation possibleComputationally demandinghttps://github.com/salesforce/progen (accessed on 5 September 2025)[109]
ESM216 May 2023YesLearns evolutionary patterns for accurate structure–function prediction.PLMHighExcellent embeddings; fastNot primarily generativehttps://github.com/facebookresearch/esm (accessed on 5 September 2025)[29]
ProGen215 November 2023NoEvolutionary modeling, de novo generation, and zero-shot fitness prediction.TransformerLowStrong generative powerResource heavyhttps://github.com/anonymized-research/progen2 (accessed on 5 September 2025)[110]
xTrimoPGLM3 April 2025NoLarge-scale language models for protein analysis and design.HybridLowScales to large tokensTraining/inference costlyhttps://github.com/ONERAI/xTrimoPGLM (accessed on 5 September 2025)[115]
ProGen316 April 2025YesIts scale enables broader viable protein generation.TransformerLowSuper-scale generative modelExtremely resource-intensivehttps://github.com/Profluent-AI/progen3 (accessed on 5 September 2025)[27]
Sequence–structure co-design
Multiflow7 February 2024NoDFMs and multiflow enable protein co-design.Flow-matchingHighAccurate joint sequence–structureNo side-chain outputhttps://github.com/jasonkyuyim/multiflow (accessed on 5 September 2025)[118]
ProteinGenerator25 September 2024YesGenerates diverse de novo proteins under customizable sequence constraints.DiffusionMiddleProperty-guided seq–structure co-designWeak on long proteinhttps://github.com/RosettaCommons/protein_generator (accessed on 5 September 2025)[119]
ESM316 January 2025YesSupports multi-modal prompt control (sequences, structures, and functions) for generating proteins.PLMLowMultimodal reasoningComputationally demandinghttps://github.com/Cogibra/esm3 (accessed on 5 September 2025)[120]
Pinal31 March 2025NoProtein structure and language co-constrained sequence design.TransformerLowText → structure → sequence pipelineLimited by training distribution biaseshttps://github.com/westlake-repl/Denovo-Pinal (accessed on 5 September 2025)[121]

5. AI as an Engine for Protein Functional Universe Exploration

AI is rapidly emerging as the principal driver for systematic protein functional exploration by incorporating accurate structure prediction, de novo design, and iterative experimental feedback into a seamless discovery engine. Generative frameworks now propose novel fold topologies and candidate sequences that expand the accessible sequence–structure space, while predictive models trained on biochemical and phenotypic assays provide rapid in silico triaging of catalytic efficiencies, binding affinity, and other functional indicators [128,129,130,131,132]. These components are commonly integrated into a closed-loop “design–predict–test–learn” workflow: (i) an acquisition policy (often based on a predicted score, uncertainty, or an acquisition function) selects a small, high-value set of candidates for synthesis and assay; (ii) experimental readouts (e.g., binding KD, kcat/Km, stability, expression) are quality-controlled and preprocessed to account for assay noise and batch effects; and (iii) the resulting labeled data are used to update models via targeted fine-tuning, retraining of surrogate predictors, or Bayesian/active-learning schemes that prioritize the next experiments. In practice, effective feedback requires uncertainty-aware models (to avoid overconfident selection), multi-fidelity modeling, and methods for domain adaptation when designs depart from the training distribution. Major challenges remain—the limited experimental throughput and cost, inconsistent assays, label sparsity for rare functions, and potential distribution shifts between training data and designed sequences—and these practical constraints shape algorithmic choices. Automation (robotics, standardized metadata, and data pipelines) and careful reporting of model scores and assay conditions improve the speed and reliability of the loop, but exploration must still be balanced with pragmatic developability and biosafety considerations. Through this seamless cycle, AI transforms de novo protein design from a laborious trial-and-error process into a rational, guided, data-centric exploration, unlocking completely new biochemical transformations, substrate specificities, and regulatory mechanisms within the protein functional universe and accelerating the development of next-generation therapeutic proteins [133,134], biocatalysts [93,135,136], biosensors [137,138], and self-assembling materials [139,140]. This underpins the assembly of ever more sophisticated synthetic biology circuits for precise cellular regulation (Figure 4).

5.1. Exploring Novel Folds and Topologies

Yeo et al. clustered 821 million structures from AFDB and ESMatlas by sequence and structural similarity, yielding 5.12 million non-singleton clusters. Strikingly, despite encompassing over 600 million predicted structures, ESMatlas yielded only one novel fold when compared to the more than 200 million entries in AFDB, representing only a minute fraction of the theoretical landscape of possible protein folds and underscoring that vast regions of the structure space remain unexplored [25]. AI-driven methods have begun to fill these gaps by uncovering entirely new structures. RFdiffusion All-Atom produced functional small-molecule binders for digoxigenin, heme, and bilin that adopt non-natural backbone geometries with a near-zero sequence homology to any PDB entry yet achieve high-affinity binding via unprecedented folds [60]. Similarly, SCUBA-D’s unconditional sampling of just 500 backbones revealed multiple topologies absent from nature, confirming that vast uncharted regions persist within the structure space [123]. TopoDiff further expanded this space by designing proteins composed exclusively of β strands and coils—none of which resemble known structures [125]. AI methods also enable the creation of modular twistless helix-repeat (THR) models that assemble similar molecular “Lego” to form polygons, rings, cages, and tubes of a defined size [139]. These results demonstrate that AI methods can systematically explore and experimentally realize protein structures far beyond the limits of natural evolution.

5.2. Designing Functional Sites De Novo

AI-driven de novo protein design achieves functional diversification through de novo construction of functional sites such as tailored binding pockets, catalytic centers, and allosteric regulators—transforming the function-first design blueprint from concept to reality.
For instance, Wu et al. integrated physics-based and deep-learning approaches (e.g., RFdiffusion) to build binding pockets for intrinsically disordered-region sequences. Screening 22 designs per target across 39 unstructured targets, they observed binders with affinities of 100 pM-100 nM for 34 targets. Glögl et al. then tackled the challenge of flat, polar interfaces by designing TNFR1 antagonists (Kd < 10 pM) that effectively inhibit TNF-α signaling and demonstrate efficacy against inflammatory cascades previously resistant to inhibition [133]. Beyond the binding pocket, Pillai et al. created switchable assemblies whose oligomeric state transitions in response to small-molecule effectors, enabling allosteric control systems for drug delivery and dynamic regulation of synthetic cellular pathways [140]. On the catalytic site design front, Hou et al. developed a class of artificial protein catalysts termed NovoChromes, which bind both heme and synthetic porphyrins to efficiently catalyze nonnatural reactions, including cyclopropanation and silylation. Through de novo design and directed evolution, these catalysts achieve high efficiency and stereoselectivity, exhibiting remarkable stability in concentrated organic solvent conditions (up to 70% ethanol) and under high thermal resistance (Tm > 90 °C) [141].
Together, these examples illustrate how AI tools, by specifying only the desired function, can autonomously sculpt atomically precise functional geometries within synthetic frameworks, unlocking previously inaccessible regions of the protein functional universe.

5.3. Exploring Sequence–Structure–Function Landscapes

Proteins exist in a high-dimensional sequence–structure–function landscape whose topology is extremely rugged. Each protein sequence corresponds to a point defined by its three-dimensional structure and its functional properties. Due to pervasive epistasis (mutations interacting in non-additive ways), the fitness landscape is pockmarked with many local optima [142,143]. In practical terms, this means that changing one amino acid can have very different effects depending on the rest of the sequence; thus, evolutionary or traditional protein engineering searches often become trapped in local peaks and fail to find globally optimal solutions. Most computational design methods must “navigate a rugged fitness landscape incrementally”, much like natural evolution does [95,144]. Directed evolution or physics-based design thus tends to explore only a small sub-region of the sequence space.
By contrast, modern AI approaches learn continuous latent embeddings of protein sequences and structures that capture underlying biophysical and evolutionary constraints [120]. Within these latent spaces, AI models can propose novel sequences and topological structures and direct probabilistic sampling toward regions of higher predicted fitness, thereby efficiently navigating the complex sequence–structure–function landscape [119].
AI-driven de novo protein design has already explored novel regions of protein function that nature never sampled. For example, a multimodal language model ESM3 that considers sequence, structure, and function was prompted to generate fluorescent proteins. Among the designs was “esmGFP”, a bright green fluorescent protein only 58% identical to any known example, placing it in a region of the functional landscape that natural evolution never explored [120]. Similarly, Gao et al. introduced AiCE, a framework that samples from inverse-folding models with structural and evolutionary constraints. AiCE tackled eight distinct engineering challenges—covering deaminase enzymes, nucleases, base editors, and more—achieving success rates between 11% and 88% in identifying improved variants and efficiently navigating the fitness landscape across multiple structures and functions [145].
In addition, AI-driven de novo protein design still depends on the integration of experimental feedback; as Biswas et al. showed, latent models guided only by evolutionary priors tend to avoid nonfunctional regions in fitness landscapes. In addition, without any real experimental data, they may not find the highest-activity variants [38].
AI-driven de novo protein design continually sharpens our view of the sequence–structure–function landscape and empowers us to move beyond incremental tweaks. By iterating between AI-driven exploration and wet-lab validation, initial models can delineate the boundaries of promising sequence regions, with targeted lab validation filling in the gaps and correcting biases. This feedback loop refines the latent embeddings, making the model’s landscape smoother and more accurate around high-fitness areas [146]. We can progressively reveal and exploit the global peaks of protein functionality to not only overcome local traps in the sequence–structure–function landscape but also to steadily expand the frontiers of the protein functional universe.

5.4. AI-Driven De Novo Protein Design for Applications in Biotechnology and Synthetic Biology

AI-driven de novo protein design is rapidly maturing from a method-focused discipline into a practical engine for biotechnology and synthetic biology. In this Section 5.4, we summarize representative application domains—therapeutic proteins, enzyme engineering, biosensors/materials, and synthetic biology regulators—and summarize design objectives and computational workflows, together with developability and performance metrics (Table 3).
In therapeutics, target/epitope specification has led to the production of compact proteins and peptides with high affinity, exceptional stability, and demonstrable efficacy in cellular and animal models. Vázquez Torres et al. engineered miniproteins that neutralize snake venom toxins, achieving 100% survival in envenomed mice with exceptional thermal stability (Tm > 95 °C) and nanomolar affinity [134]. Mahling et al. used Colabdesign [147] to produce a 21-residue peptide (ELIXIR) that selectively inhibits pathological late or persistent Na+ current (INaL) by enhancing NaV1.5 channel inactivation. ELIXIR binds to the NaV1.5 C-terminal domain with KD = 0.89 ± 0.25 μM, achieving > 90% INaL inhibition in the disease models tested. Functionally, ELIXIR restores elevated INaL toward healthy levels in patient-derived iPSC cardiomyocytes and markedly reduces INaL and shortens QTc in transgenic mouse models [148].
In enzyme engineering, Listov et al. developed a computational enzyme-design workflow that produced efficient Kemp eliminase without further experimental optimization. Of the 73 designs tested, 3 (~4.1%) exhibited measurable Kemp-elimination activities. The top variant, Des27.7, achieved kcat/Km = 1.27 × 104 M−1s−1, an approximately 60-fold improvement over the initial design; the experimentally observed active-site conformation matched the design model to within 0.5 Å [149]. Lauko et al. employed a deep learning-driven approach to de novo enzyme design, resulting in a serine hydrolase with a previously unobserved protein fold. The designed enzyme demonstrated a catalytic efficiency (kcat/Km) of up to 2.2 × 105 M−1s−1. Among the 132 designed variants, 20% exhibited detectable hydrolytic activity [136]. Munsamy et al. used the ZymCTRL framework to design 20 carbonic anhydrase variants, each with <50% sequence identity to known enzymes; 7 of these (35%) exhibited measurable catalytic activity. Applying a similar design workflow to lactate dehydrogenase yielded 20 candidates, of which 14 (70%) displayed detectable enzymatic activity [150].
In synthetic biology, researchers aspire to “program” living systems by designing predictable genetic parts that drive cells to perform desired functions, and AI has emerged as a new enabling engine for this effort. Zhang et al. exemplified this approach by engineering intracellular Ras–GTP activity sensors (Ras-LOCKR-S) and proximity-labeling modules (Ras-LOCKR-PL) that operate with a subcellular spatial resolution; these tools were subsequently used to dissect mechanisms of resistance to Ras-G12C inhibitors [137,138].
Table 3. AI-driven de novo Protein Design—Biotech Applications Overview.
Table 3. AI-driven de novo Protein Design—Biotech Applications Overview.
Molecule NumberMoleculeTarget and ActivityMethodIndications/FunctionRef.
Therapeutic Proteins
1 and 2 and 3SHRTShort-chain α-neurotoxins (ScNtx)
KD = 0.9 nM, Tm = 78 °C
RFdiffusion ProteinMPNNSnake venom toxins.[134]
LNGLong-chain α-neurotoxin
(P01391)
KD = 1.9 nM, Tm > 95 °C
CYTXCytotoxins (Naja pallida)
KD = 271 nM, Tm = 61 °C
4TNFR1_mb2_pd1The tumor necrosis factor receptor 1(TNFR1)
KD (TNFR1) < 10 pM
RFdiffusion ProteinMPNNInflammatory disease.[133]
5 and 6 and 723R-91Interleukin (IL)-23R
KD < 1 pM
RosettaAutoinflammatory diseases.[48]
17–53IL-17
KD = 10 pM
8ELIXIRNaV1.5 carboxy-terminal domain
KD = 0.89 ± 0.25 μM
AfDesignCardiac arrhythmias and epilepsy.[148]
Enzyme Engineering
9Serine hydrolaseskcat/Km = 2.2 × 105 M−1s−1RFdiffusion LigandMPNN
PLACER
Catalyze ester hydrolysis with catalytic.[136]
10Kemp eliminasekcat/Km = 1.27 × 104 M−1s−1Rosetta, PROSS
FuncLib,
AlphaFold2
Kemp elimination.[149]
11Metallohydrolaseskcat/Km = 2.3 × 104 M−1s−1RFam,
ProteinMPNN
AlphaFold2
Catalyzes some difficult hydrolysis reactions.[151]
12Retroaldolasekcat/Km = 1.1 × 104 M−1min−1ChemNet,
Rosetta
LigandMPNN
Catalyze the reverse aldol reaction.[152]
13 and 14Carbonic anhydrases
Lactate dehydrogenases
NAZymCTRLThe fastest enzymes known in nature.
Primarily in lactic acid production.
[150]
Synthetic biological components
15Ras-LOCKR-S/PLRas-GTPRosetta,
AlphaFold
Sensor for Ras activity.
Ras activity-dependent Proximity Labeler.
[137]
16THR/Rosetta,
ProteinMPNN
Enables modular nanomaterial design.[139]
17Allosterically protein assemblies/Rosetta,
ProteinMPNN,
RFDiffusion
AlphaFold2
Allosteric modulation.[140]

6. Conclusions

AI-driven de novo protein design has substantially transformed our ability to explore the protein functional universe, enabling the rational engineering of novel folds, bespoke functional sites, and proteins with tailored biophysical properties—with a growing number of experimentally validated successes. To translate these advances into robust, scalable pipelines, we must address several practical challenges.
Model performance highly depends on the quality, diversity, and annotation of training data. Public sequence and structure resources are extensive but biased toward model organisms and readily assayed families, which limits generalization to under-represented families and wholly synthetic sequences. The experimental throughput, noisy or heterogeneous assays, and the distribution shift between training corpora and designed sequences further complicate the reliable translation from in silico designs to functional molecules. In addition, biosafety and validation must be integrated early: designs that introduce novel folds or activities can have unintended biological effects (toxicity, off-target interactions, or immune responses) and therefore require rigorous multiomic and phenotypic validation, together with explicit risk-mitigation strategies, prior to any in vivo work. Explicit risk assessment and traceable data sources are prerequisites for clinical or industrial applications [153,154,155].
Looking forward, models should be improved to better capture the multi-state conformational ensembles, allostery, and dynamic interactions that underpin biological function; they should also provide calibrated uncertainty estimates and more interpretable outputs to support users’ design choices. Progress will depend on improving dataset curation (for example, adding quality-filtered metagenomic sequences and richer functional labels), standardizing assay and reporting practices, and sharing community benchmarking datasets to increase robustness and comparability. We further recommend adopting unified scoring standards to enable direct, intuitive comparison among models. Finally, the computational infrastructure remains a key enabler. GPU-accelerated parallelism and optimized distributed training/inference stacks reduce iteration times and permit larger models and broader in silico screenings. Emerging quantum computing approaches may eventually assist specialized optimization [156,157,158].
In conclusion, AI has already opened new horizons for programmable biology. Realizing its full promise will require parallel progress on data, algorithms, computing, interpretability, and biological validation—combined with explicit biosafety safeguards—to ensure those horizons are explored responsibly and efficiently.

Author Contributions

G.Z., C.L. and J.L.: writing—original draft preparation; S.Z. and L.Z.: writing—review and editing. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by the NUDT Research Program 22-TDRCJH-02-015 (L.Z.) and 2023-lxy-fhjj-005 and 24-ZZCX-JDZ-02 and the National Natural Science Foundation of China project (No. 32401056).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

No new data were created or analyzed in this study.

Conflicts of Interest

The authors declare no conflicts of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript; or in the decision to publish the results.

References

  1. Liu, J.; Wang, X.; Ye, X.; Chen, D. Improved health outcomes of nasopharyngeal carcinoma patients 3 years after treatment by the AI-assisted home enteral nutrition management. Front. Nutr. 2025, 11, 1481073. [Google Scholar] [CrossRef]
  2. Yin, M.; Feng, C.; Yu, Z.; Zhang, Y.; Li, Y.; Wang, X.; Song, C.; Guo, M.; Li, C. sc2GWAS: A comprehensive platform linking single cell and GWAS traits of human. Nucleic Acids Res. 2025, 53, D1151–D1161. [Google Scholar] [CrossRef]
  3. Ulmer, K.M. Protein engineering. Science 1983, 219, 666–671. [Google Scholar] [CrossRef]
  4. Arnold, F.H. Design by directed evolution. Acc. Chem. Res. 1998, 31, 125–131. [Google Scholar] [CrossRef]
  5. Jäckel, C.; Kast, P.; Hilvert, D. Protein design by directed evolution. Annu. Rev. Biophys. 2008, 37, 153–173. [Google Scholar] [CrossRef]
  6. Koshland, D.E. Application of a theory of enzyme specificity to protein synthesis. Proc. Natl. Acad. Sci. USA 1958, 44, 98–104. [Google Scholar] [CrossRef] [PubMed]
  7. Fersht, A.R.; Shi, J.-P.; Knill-Jones, J.; Lowe, D.M.; Wilkinson, A.J.; Blow, D.M.; Brick, P.; Carter, P.; Waye, M.M.Y.; Winter, G. Hydrogen bonding and biological specificity analysed by protein engineering. Nature 1985, 314, 235–238. [Google Scholar] [CrossRef] [PubMed]
  8. Palczewski, K. G Protein–coupled receptor rhodopsin. Annu. Rev. Biochem. 2006, 75, 743–767. [Google Scholar] [CrossRef]
  9. Hunter, T. Signaling—2000 and beyond. Cell 2000, 100, 113–127. [Google Scholar] [CrossRef] [PubMed]
  10. Wittinghofer, A.; Vetter, I.R. Structure-function relationships of the g domain, a canonical switch motif. Annu. Rev. Biochem. 2011, 80, 943–971. [Google Scholar] [CrossRef]
  11. Wilson, I.A.; Skehel, J.J.; Wiley, D.C. Structure of the haemagglutinin membrane glycoprotein of influenza virus at 3 A resolution. Nature 1981, 289, 366–373. [Google Scholar] [CrossRef] [PubMed]
  12. Garcia, K.C.; Degano, M.; Stanfield, R.L.; Brunmark, A.; Jackson, M.R.; Peterson, P.A.; Teyton, L.; Wilson, I.A. An Aβ T cell receptor structure at 2.5 Å and its orientation in the TCR-MHC complex. Science 1996, 274, 209–219. [Google Scholar] [CrossRef] [PubMed]
  13. Mitchison, T.; Kirschner, M. Dynamic instability of microtubule growth. Nature 1984, 312, 237–242. [Google Scholar] [CrossRef]
  14. Shoulders, M.D.; Raines, R.T. Collagen structure and stability. Annu. Rev. Biochem. 2009, 78, 929–958. [Google Scholar] [CrossRef]
  15. Orgel, J.P.R.O.; Irving, T.C.; Miller, A.; Wess, T.J. Microfibrillar structure of type I collagen in situ. Proc. Natl. Acad. Sci. USA 2006, 103, 9001–9005. [Google Scholar] [CrossRef]
  16. Walport Mark, J. Complement. N. Engl. J. Med. 2001, 344, 1058–1066. [Google Scholar] [CrossRef]
  17. Janeway, C.A.J. Approaching the asymptote? Evolution and revolution in immunology. Cold Spring Harb. Symp. Quant. Biol. 1989, 54 Pt 1, 1–13. [Google Scholar] [CrossRef]
  18. Zhang, P.; Wei, L.; Li, J.; Wang, X. Artificial intelligence-guided strategies for next-generation biological sequence design. Natl. Sci. Rev. 2024, 11, nwae343. [Google Scholar] [CrossRef]
  19. Saikia, B.; Baruah, A. Recent advances in de novo computational design and redesign of intrinsically disordered proteins and intrinsically disordered protein regions. Arch. Biochem. Biophys. 2024, 752, 109857. [Google Scholar] [CrossRef]
  20. Anfinsen, C.B. Principles that govern the folding of protein chains. Science 1973, 181, 223–230. [Google Scholar] [CrossRef] [PubMed]
  21. Faure, A.J.; Martí-Aranda, A.; Hidalgo-Carcedo, C.; Beltran, A.; Schmiedel, J.M.; Lehner, B. The genetic architecture of protein stability. Nature 2024, 634, 995–1003. [Google Scholar] [CrossRef]
  22. Keefe, A.D.; Szostak, J.W. Functional proteins from a random-sequence library. Nature 2001, 410, 715–718. [Google Scholar] [CrossRef] [PubMed]
  23. Copp, J.N.; Akiva, E.; Babbitt, P.C.; Tokuriki, N. Revealing unexplored sequence-function space using sequence similarity networks. Biochemistry 2018, 57, 4651–4662. [Google Scholar] [CrossRef]
  24. Lemke, O.; Heineike, B.M.; Viknander, S.; Cohen, N.; Li, F.; Steenwyk, J.L.; Spranger, L.; Agostini, F.; Lee, C.T.; Aulakh, S.K.; et al. The role of metabolism in shaping enzyme structures over 400 million years. Nature 2025, 644, 280–289. [Google Scholar] [CrossRef] [PubMed]
  25. Yeo, J.; Han, Y.; Bordin, N.; Lau, A.M.; Kandathil, S.M.; Kim, H.; Karin, E.L.; Mirdita, M.; Jones, D.T.; Orengo, C.; et al. Metagenomic-scale analysis of the predicted protein structure universe. bioRxiv 2025. [Google Scholar] [CrossRef]
  26. Richardson, L.; Allen, B.; Baldi, G.; Beracochea, M.; Bileschi, M.L.; Burdett, T.; Burgin, J.; Caballero-Pérez, J.; Cochrane, G.; Colwell, L.J.; et al. MGnify: The microbiome sequence data analysis resource in 2023. Nucleic Acids Res. 2023, 51, D753–D759. [Google Scholar] [CrossRef]
  27. Bhatnagar, A.; Jain, S.; Beazer, J.; Curran, S.C.; Hoffnagle, A.M.; Ching, K.; Martyn, M.; Nayfach, S.; Ruffolo, J.A.; Madani, A. Scaling unlocks broader generation and deeper functional understanding of proteins. bioRxiv 2025. [Google Scholar] [CrossRef]
  28. Varadi, M.; Bertoni, D.; Magana, P.; Paramval, U.; Pidruchna, I.; Radhakrishnan, M.; Tsenkov, M.; Nair, S.; Mirdita, M.; Yeo, J.; et al. AlphaFold Protein Structure Database in 2024: Providing structure coverage for over 214 million protein sequences. Nucleic Acids Res. 2024, 52, D368–D375. [Google Scholar] [CrossRef]
  29. Lin, Z.; Akin, H.; Rao, R.; Hie, B.; Zhu, Z.; Lu, W.; Smetanin, N.; Verkuil, R.; Kabeli, O.; Shmueli, Y.; et al. Evolutionary-scale prediction of atomic-level protein structure with a language model. Science 2023, 379, 1123–1130. [Google Scholar] [CrossRef]
  30. Linsky, T.W.; Noble, K.; Tobin, A.R.; Crow, R.; Carter, L.; Urbauer, J.L.; Baker, D.; Strauch, E.-M. Sampling of structure and sequence space of small protein folds. Nat. Commun. 2022, 13, 7151. [Google Scholar] [CrossRef]
  31. Minami, S.; Kobayashi, N.; Sugiki, T.; Nagashima, T.; Fujiwara, T.; Tatsumi-Koga, R.; Chikenji, G.; Koga, N. Exploration of novel αβ-protein folds through de novo design. Nat. Struct. Mol. Biol. 2023, 30, 1132–1140. [Google Scholar] [CrossRef] [PubMed]
  32. Huang, P.-S.; Boyken, S.E.; Baker, D. The coming of age of de novo protein design. Nature 2016, 537, 320–327. [Google Scholar] [CrossRef]
  33. Kortemme, T. De novo protein design—From new structures to programmable functions. Cell 2024, 187, 526–544. [Google Scholar] [CrossRef] [PubMed]
  34. Sellés Vidal, L.; Isalan, M.; Heap, J.T.; Ledesma-Amaro, R. A primer to directed evolution: Current methodologies and future directions. RSC Chem. Biol. 2023, 4, 271–291. [Google Scholar] [CrossRef] [PubMed]
  35. Regan, L.; DeGrado, W.F. Characterization of a helical protein designed from first principles. Science 1988, 241, 976–978. [Google Scholar] [CrossRef]
  36. Kuhlman, B.; Bradley, P. Advances in protein structure prediction and design. Nat. Rev. Mol. Cell Biol. 2019, 20, 681–697. [Google Scholar] [CrossRef]
  37. Jänes, J.; Beltrao, P. Deep learning for protein structure prediction and design—Progress and applications. Mol. Syst. Biol. 2024, 20, 162–169. [Google Scholar] [CrossRef]
  38. Muir, D.F.; Asper, G.P.R.; Notin, P.; Posner, J.A.; Marks, D.S.; Keiser, M.J.; Pinney, M.M. Evolutionary-scale enzymology enables exploration of a rugged catalytic landscape. Science 2025, 388, eadu1058. [Google Scholar] [CrossRef]
  39. Leaver-Fay, A.; Tyka, M.; Lewis, S.M.; Lange, O.F.; Thompson, J.; Jacak, R.; Kaufman, K.; Renfrew, P.D.; Smith, C.A.; Sheffler, W.; et al. ROSETTA3: An object-oriented software suite for the simulation and design of macromolecules. Methods Enzymol. 2011, 487, 545–574. [Google Scholar] [CrossRef]
  40. Negron, C.; Keating, A.E. Multistate protein design using CLEVER and CLASSY. Methods Enzymol. 2013, 523, 171–190. [Google Scholar] [CrossRef]
  41. Smadbeck, J.; Peterson, M.B.; Khoury, G.A.; Taylor, M.S.; Floudas, C.A. Protein WISDOM: A workbench for In silico De novo design of biomolecules. J. Vis. Exp. 2013, 77, e50476. [Google Scholar] [CrossRef] [PubMed]
  42. Wood, C.W.; Bruning, M.; Ibarra, A.Á.; Bartlett, G.J.; Thomson, A.R.; Sessions, R.B.; Brady, R.L.; Woolfson, D.N. CCBuilder: An interactive web-based tool for building, designing and assessing coiled-coil protein assemblies. Bioinformatics 2014, 30, 3029–3035. [Google Scholar] [CrossRef] [PubMed]
  43. Epstein, C.; Goldberger, R.; Anfinsen, C. The genetic control of tertiary protein structure: Studies with model systems. Cold Spring Harb. Symp. Quant. Biol. 1963, 28, 439–449. [Google Scholar] [CrossRef]
  44. Kuhlman, B.; Dantas, G.; Ireton, G.C.; Varani, G.; Stoddard, B.L.; Baker, D. Design of a novel globular protein fold with atomic-level accuracy. Science 2003, 302, 1364–1368. [Google Scholar] [CrossRef]
  45. Röthlisberger, D.; Khersonsky, O.; Wollacott, A.M.; Jiang, L.; DeChancie, J.; Betker, J.; Gallaher, J.L.; Althoff, E.A.; Zanghellini, A.; Dym, O.; et al. Kemp elimination catalysts by computational enzyme design. Nature 2008, 453, 190–195. [Google Scholar] [CrossRef]
  46. Siegel, J.B.; Zanghellini, A.; Lovick, H.M.; Kiss, G.; Lambert, A.R.; St Clair, J.L.; Gallaher, J.L.; Hilvert, D.; Gelb, M.H.; Stoddard, B.L.; et al. Computational design of an enzyme catalyst for a stereoselective bimolecular diels-alder reaction. Science 2010, 329, 309–313. [Google Scholar] [CrossRef]
  47. Roy, A.; Shi, L.; Chang, A.; Dong, X.; Fernandez, A.; Kraft, J.C.; Li, J.; Le, V.Q.; Winegar, R.V.; Cherf, G.M.; et al. De novo design of highly selective miniprotein inhibitors of integrins αvβ6 and αvβ8. Nat. Commun. 2023, 14, 5660. [Google Scholar] [CrossRef] [PubMed]
  48. Berger, S.; Seeger, F.; Yu, T.-Y.; Aydin, M.; Yang, H.; Rosenblum, D.; Guenin-Macé, L.; Glassman, C.; Arguinchona, L.; Sniezek, C.; et al. Preclinical proof of principle for orally delivered Th17 antagonist miniproteins. Cell 2024, 187, 4305–4317.e18. [Google Scholar] [CrossRef] [PubMed]
  49. Huang, B.; Coventry, B.; Borowska, M.T.; Arhontoulis, D.C.; Exposit, M.; Abedi, M.; Jude, K.M.; Halabiya, S.F.; Allen, A.; Cordray, C.; et al. De novo design of miniprotein antagonists of cytokine storm inducers. Nat. Commun. 2024, 15, 7064. [Google Scholar] [CrossRef] [PubMed]
  50. Notin, P.; Rollins, N.; Gal, Y.; Sander, C.; Marks, D. Machine learning for functional protein design. Nat. Biotechnol. 2024, 42, 216–228. [Google Scholar] [CrossRef]
  51. Hsu, C.; Fannjiang, C.; Listgarten, J. Generative models for protein structures and sequences. Nat. Biotechnol. 2024, 42, 196–199. [Google Scholar] [CrossRef]
  52. Strokach, A.; Kim, P.M. Deep generative modeling for protein design. Curr. Opin. Struct. Biol. 2022, 72, 226–236. [Google Scholar] [CrossRef]
  53. Chandra, A.; Tünnermann, L.; Löfstedt, T.; Gratz, R. Transformer-based deep learning for predicting protein properties in the life sciences. eLife 2023, 12, e82819. [Google Scholar] [CrossRef] [PubMed]
  54. Earl, L.A.; Falconieri, V.; Milne, J.L.; Subramaniam, S. Cryo-EM: Beyond the microscope. Curr. Opin. Struct. Biol. 2017, 46, 71–78. [Google Scholar] [CrossRef] [PubMed]
  55. Araya, C.L.; Fowler, D.M. Deep mutational scanning: Assessing protein function on a massive scale. Trends Biotechnol. 2011, 29, 435–442. [Google Scholar] [CrossRef]
  56. Abramson, J.; Adler, J.; Dunger, J.; Evans, R.; Green, T.; Pritzel, A.; Ronneberger, O.; Willmore, L.; Ballard, A.J.; Bambrick, J.; et al. Accurate structure prediction of biomolecular interactions with AlphaFold 3. Nature 2024, 630, 493–500. [Google Scholar] [CrossRef]
  57. Yan, Y.; Tao, H.; He, J.; Huang, S.-Y. The HDOCK server for integrated protein-protein docking. Nat. Protoc. 2020, 15, 1829–1852. [Google Scholar] [CrossRef]
  58. Reynisson, B.; Alvarez, B.; Paul, S.; Peters, B.; Nielsen, M. NetMHCpan-4.1 and NetMHCIIpan-4.0: Improved predictions of MHC antigen presentation by concurrent motif deconvolution and integration of MS MHC eluted ligand data. Nucleic Acids Res. 2020, 48, W449–W454. [Google Scholar] [CrossRef]
  59. Hon, J.; Marusiak, M.; Martinek, T.; Kunka, A.; Zendulka, J.; Bednar, D.; Damborsky, J. SoluProt: Prediction of soluble protein expression in Escherichia coli. Bioinformatics 2021, 37, 23–28. [Google Scholar] [CrossRef]
  60. Krishna, R.; Wang, J.; Ahern, W.; Sturmfels, P.; Venkatesh, P.; Kalvet, I.; Lee, G.R.; Morey-Burrows, F.S.; Anishchenko, I.; Humphreys, I.R.; et al. Generalized biomolecular modeling and design with RoseTTAFold All-Atom. Science 2024, 384, eadl2528. [Google Scholar] [CrossRef] [PubMed]
  61. Abriata, L.A. The Nobel Prize in Chemistry: Past, present, and future of AI in biology. Commun. Biol. 2024, 7, 1409. [Google Scholar] [CrossRef]
  62. Marks, D.S.; Hopf, T.A.; Sander, C. Protein structure prediction from sequence variation. Nat. Biotechnol. 2012, 30, 1072–1080. [Google Scholar] [CrossRef]
  63. Wüthrich, K. Protein structure determination in solution by NMR spectroscopy. J. Biol. Chem. 1990, 265, 22059–22062. [Google Scholar] [CrossRef] [PubMed]
  64. Shi, Y. A glimpse of structural biology through X-ray crystallography. Cell 2014, 159, 995–1014. [Google Scholar] [CrossRef] [PubMed]
  65. Yang, Z.; Zeng, X.; Zhao, Y.; Chen, R. AlphaFold2 and its applications in the fields of biology and medicine. Signal Transduct. Target. Ther. 2023, 8, 115. [Google Scholar] [CrossRef]
  66. Pakhrin, S.C.; Shrestha, B.; Adhikari, B.; Kc, D.B. Deep learning-based advances in protein structure prediction. Int. J. Mol. Sci. 2021, 22, 5553. [Google Scholar] [CrossRef]
  67. Jumper, J.; Evans, R.; Pritzel, A.; Green, T.; Figurnov, M.; Ronneberger, O.; Tunyasuvunakool, K.; Bates, R.; Žídek, A.; Potapenko, A.; et al. Highly accurate protein structure prediction with AlphaFold. Nature 2021, 596, 583–589. [Google Scholar] [CrossRef]
  68. Kryshtafovych, A.; Schwede, T.; Topf, M.; Fidelis, K.; Moult, J. Critical assessment of methods of protein structure prediction (CASP)-Round XIV. Proteins 2021, 89, 1607–1617. [Google Scholar] [CrossRef] [PubMed]
  69. Baek, M.; DiMaio, F.; Anishchenko, I.; Dauparas, J.; Ovchinnikov, S.; Lee, G.R.; Wang, J.; Cong, Q.; Kinch, L.N.; Schaeffer, R.D.; et al. Accurate prediction of protein structures and interactions using a three-track neural network. Science 2021, 373, 871–876. [Google Scholar] [CrossRef]
  70. Mirdita, M.; Schütze, K.; Moriwaki, Y.; Heo, L.; Ovchinnikov, S.; Steinegger, M. ColabFold: Making protein folding accessible to all. Nat. Methods 2022, 19, 679–682. [Google Scholar] [CrossRef]
  71. Ahdritz, G.; Bouatta, N.; Floristean, C.; Kadyan, S.; Xia, Q.; Gerecke, W.; O’Donnell, T.J.; Berenberg, D.; Fisk, I.; Zanichelli, N.; et al. OpenFold: Retraining AlphaFold2 yields new insights into its learning mechanisms and capacity for generalization. Nat. Methods 2024, 21, 1514–1524. [Google Scholar] [CrossRef] [PubMed]
  72. Chen, Y.; Xu, Y.; Liu, D.; Xing, Y.; Gong, H. An end-to-end framework for the prediction of protein structure and fitness from single sequence. Nat. Commun. 2024, 15, 7400. [Google Scholar] [CrossRef] [PubMed]
  73. Wu, R.; Ding, F.; Wang, R.; Shen, R.; Zhang, X.; Luo, S.; Su, C.; Wu, Z.; Xie, Q.; Berger, B.; et al. High-resolution de novo structure prediction from primary sequence. bioRxiv 2022. [Google Scholar] [CrossRef]
  74. Kalakoti, Y.; Wallner, B. AFsample2 predicts multiple conformations and ensembles with AlphaFold2. Commun. Biol. 2025, 8, 373. [Google Scholar] [CrossRef]
  75. Zheng, W.; Wuyun, Q.; Li, Y.; Liu, Q.; Zhou, X.; Peng, C.; Zhu, Y.; Freddolino, L.; Zhang, Y. Deep-learning-based single-domain and multidomain protein structure prediction with D-I-TASSER. Nat. Biotechnol. 2025. [Google Scholar] [CrossRef]
  76. Elofsson, A. Progress at protein structure prediction, as seen in CASP15. Curr. Opin. Struct. Biol. 2023, 80, 102594. [Google Scholar] [CrossRef]
  77. Moussad, B.; Roche, R.; Bhattacharya, D. The transformative power of transformers in protein structure prediction. Proc. Natl. Acad. Sci. USA 2023, 120, e2303499120. [Google Scholar] [CrossRef]
  78. Li, H.; Lei, Y.; Zeng, J. Revolutionizing biomolecular structure determination with artificial intelligence. Natl. Sci. Rev. 2024, 11, nwae339. [Google Scholar] [CrossRef]
  79. Evans, R.; O’Neill, M.; Pritzel, A.; Antropova, N.; Senior, A.; Green, T.; Žídek, A.; Bates, R.; Blackwell, S.; Yim, J.; et al. Protein complex prediction with AlphaFold-Multimer. bioRxiv 2021. [Google Scholar] [CrossRef]
  80. Baek, M.; McHugh, R.; Anishchenko, I.; Jiang, H.; Baker, D.; DiMaio, F. Accurate prediction of protein–nucleic acid complexes using RoseTTAFoldNA. Nat. Methods 2024, 21, 117–121. [Google Scholar] [CrossRef]
  81. Discovery, C.; Boitreaud, J.; Dent, J.; McPartlon, M.; Meier, J.; Reis, V.; Rogozhnikov, A.; Wu, K. Chai-1: Decoding the molecular interactions of life. bioRxiv 2024. [Google Scholar] [CrossRef]
  82. Wohlwend, J.; Corso, G.; Passaro, S.; Getz, N.; Reveiz, M.; Leidal, K.; Swiderski, W.; Atkinson, L.; Portnoi, T.; Chinn, I.; et al. Boltz-1 democratizing biomolecular interaction modeling. bioRxiv 2024. [Google Scholar] [CrossRef]
  83. Passaro, S.; Corso, G.; Wohlwend, J.; Reveiz, M.; Thaler, S.; Somnath, V.R.; Getz, N.; Portnoi, T.; Roy, J.; Stark, H.; et al. Boltz-2: Towards accurate and efficient binding affinity prediction. bioRxiv 2025. [Google Scholar] [CrossRef]
  84. Zheng, S.; He, J.; Liu, C.; Shi, Y.; Lu, Z.; Feng, W.; Ju, F.; Wang, J.; Zhu, J.; Min, Y.; et al. Predicting equilibrium distributions for molecular systems with deep learning. Nat. Mach. Intell. 2024, 6, 558–567. [Google Scholar] [CrossRef]
  85. Lewis, S.; Hempel, T.; Jiménez-Luna, J.; Gastegger, M.; Xie, Y.; Foong, A.Y.K.; Satorras, V.G.; Abdin, O.; Veeling, B.S.; Zaporozhets, I.; et al. Scalable emulation of protein equilibrium ensembles with generative deep learning. Science 2025, 389, eadv9817. [Google Scholar] [CrossRef] [PubMed]
  86. Van Kempen, M.; Kim, S.S.; Tumescheit, C.; Mirdita, M.; Lee, J.; Gilchrist, C.L.M.; Söding, J.; Steinegger, M. Fast and accurate protein structure search with Foldseek. Nat. Biotechnol. 2024, 42, 243–246. [Google Scholar] [CrossRef] [PubMed]
  87. Yim, J.; Stärk, H.; Corso, G.; Jing, B.; Barzilay, R.; Jaakkola, T.S. Diffusion models in protein structure and docking. WIREs Comput. Mol. Sci. 2024, 14, e1711. [Google Scholar] [CrossRef]
  88. Watson, J.L.; Juergens, D.; Bennett, N.R.; Trippe, B.L.; Yim, J.; Eisenach, H.E.; Ahern, W.; Borst, A.J.; Ragotte, R.J.; Milles, L.F.; et al. De novo design of protein structure and function with RFdiffusion. Nature 2023, 620, 1089–1100. [Google Scholar] [CrossRef]
  89. Bennett, N.R.; Watson, J.L.; Ragotte, R.J.; Borst, A.J.; See, D.L.; Weidle, C.; Biswas, R.; Yu, Y.; Shrock, E.L.; Ault, R.; et al. Atomically accurate de novo design of antibodies with RFdiffusion. bioRxiv 2024. [Google Scholar] [CrossRef] [PubMed]
  90. Wu, K.; Jiang, H.; Hicks, D.R.; Liu, C.; Muratspahić, E.; Ramelot, T.A.; Liu, Y.; McNally, K.; Gaur, A.; Coventry, B.; et al. Sequence-specific targeting of intrinsically disordered protein regions. bioRxiv 2024. [Google Scholar] [CrossRef]
  91. Sappington, I.; Toul, M.; Lee, D.S.; Robinson, S.A.; Goreshnik, I.; McCurdy, C.; Chan, T.C.; Buchholz, N.; Huang, B.; Vafeados, D.; et al. Improved protein binder design using beta-pairing targeted RFdiffusion. bioRxiv 2024. [Google Scholar] [CrossRef]
  92. Rettie, S.A.; Juergens, D.; Adebomi, V.; Bueso, Y.F.; Zhao, Q.; Leveille, A.N.; Liu, A.; Bera, A.K.; Wilms, J.A.; Üffing, A.; et al. Accurate de novo design of high-affinity protein-binding macrocycles using deep learning. Nat. Chem. Biol. 2025. [Google Scholar] [CrossRef] [PubMed]
  93. Ahern, W.; Yim, J.; Tischer, D.; Salike, S.; Woodbury, S.M.; Kim, D.; Kalvet, I.; Kipnis, Y.; Coventry, B.; Altae-Tran, H.R.; et al. Atom level enzyme active site scaffolding using RFdiffusion2. bioRxiv 2025. [Google Scholar] [CrossRef]
  94. Yim, J.; Trippe, B.L.; Bortoli, V.D.; Mathieu, E.; Doucet, A.; Barzilay, R.; Jaakkola, T. SE(3) diffusion model with application to protein backbone generation. In Proceedings of Machine Learning Research, Proceedings of the 40th International Conference on Machine Learning, Honolulu, HI, USA, 23–29 July 2023; Krause, A., Brunskill, E., Cho, K., Engelhardt, B., Sabato, S., Scarlett, J., Eds.; PMLR: Cambridge, MA, USA, 2023; Volume 202, pp. 40001–40039. [Google Scholar]
  95. Ingraham, J.B.; Baranov, M.; Costello, Z.; Barber, K.W.; Wang, W.; Ismail, A.; Frappier, V.; Lord, D.M.; Ng-Thow-Hing, C.; Van Vlack, E.R.; et al. Illuminating protein space with a programmable generative model. Nature 2023, 623, 1070–1078. [Google Scholar] [CrossRef]
  96. Frank, C.; Khoshouei, A.; Fuβ, L.; Schiwietz, D.; Putz, D.; Weber, L.; Zhao, Z.; Hattori, M.; Feng, S.; De Stigter, Y.; et al. Scalable protein design using optimization in a relaxed sequence space. Science 2024, 386, 439–445. [Google Scholar] [CrossRef]
  97. Anishchenko, I.; Pellock, S.J.; Chidyausiku, T.M.; Ramelot, T.A.; Ovchinnikov, S.; Hao, J.; Bafna, K.; Norn, C.; Kang, A.; Bera, A.K.; et al. De novo protein design by deep network hallucination. Nature 2021, 600, 547–552. [Google Scholar] [CrossRef]
  98. Wicky, B.I.M.; Milles, L.F.; Courbet, A.; Ragotte, R.J.; Dauparas, J.; Kinfu, E.; Tipps, S.; Kibler, R.D.; Baek, M.; DiMaio, F.; et al. Hallucinating symmetric protein assemblies. Science 2022, 378, 56–61. [Google Scholar] [CrossRef]
  99. Geffner, T.; Didi, K.; Zhang, Z.; Reidenbach, D.; Cao, Z.; Yim, J.; Geiger, M.; Dallago, C.; Kucukbenli, E.; Vahdat, A.; et al. Proteina: Scaling Flow-based Protein Structure Generative Models. arXiv 2025. [Google Scholar] [CrossRef]
  100. Castorina, L.V.; Petrenas, R.; Subr, K.; Wood, C.W. PDBench: Evaluating computational methods for protein-sequence design. Bioinformatics 2023, 39, btad027. [Google Scholar] [CrossRef] [PubMed]
  101. Hsu, C.; Verkuil, R.; Liu, J.; Lin, Z.; Hie, B.; Sercu, T.; Lerer, A.; Rives, A. Learning inverse folding from millions of predicted structures. In Proceedings of Machine Learning Research, Proceedings of the 39th International Conference on Machine Learning, Baltimore, MD, USA, 17–23 July 2022; Chaudhuri, K., Jegelka, S., Song, L., Szepesvari, C., Niu, G., Sabato, S., Eds.; PMLR: Cambridge, MA, USA, 2022; Volume 162, pp. 8946–8970. [Google Scholar]
  102. Dauparas, J.; Anishchenko, I.; Bennett, N.; Bai, H.; Ragotte, R.J.; Milles, L.F.; Wicky, B.I.M.; Courbet, A.; de Haas, R.J.; Bethel, N.; et al. Robust deep learning-based protein sequence design using ProteinMPNN. Science 2022, 378, 49–56. [Google Scholar] [CrossRef]
  103. Sumida, K.H.; Núñez-Franco, R.; Kalvet, I.; Pellock, S.J.; Wicky, B.I.M.; Milles, L.F.; Dauparas, J.; Wang, J.; Kipnis, Y.; Jameson, N.; et al. Improving protein expression, stability, and function with ProteinMPNN. J. Am. Chem. Soc. 2024, 146, 2054–2061. [Google Scholar] [CrossRef] [PubMed]
  104. Wang, T.; Jin, X.; Lu, X.; Min, X.; Ge, S.; Li, S. Empirical validation of ProteinMPNN’s efficiency in enhancing protein fitness. Front. Genet. 2024, 14, 1347667. [Google Scholar] [CrossRef]
  105. Dauparas, J.; Lee, G.R.; Pecoraro, R.; An, L.; Anishchenko, I.; Glasscock, C.; Baker, D. Atomic context-conditioned protein sequence design using LigandMPNN. Nat. Methods 2025, 22, 717–723. [Google Scholar] [CrossRef]
  106. Ren, M.; Yu, C.; Bu, D.; Zhang, H. Accurate and robust protein sequence design with CarbonDesign. Nat. Mach. Intell. 2024, 6, 536–547. [Google Scholar] [CrossRef]
  107. Krapp, L.F.; Meireles, F.A.; Abriata, L.A.; Devillard, J.; Vacle, S.; Marcaida, M.J.; Dal Peraro, M. Context-aware geometric deep learning for protein sequence design. Nat. Commun. 2024, 15, 6273. [Google Scholar] [CrossRef]
  108. Ferruz, N.; Schmidt, S.; Höcker, B. ProtGPT2 is a deep unsupervised language model for protein design. Nat. Commun. 2022, 13, 4348. [Google Scholar] [CrossRef]
  109. Madani, A.; Krause, B.; Greene, E.R.; Subramanian, S.; Mohr, B.P.; Holton, J.M.; Olmos, J.L.; Xiong, C.; Sun, Z.Z.; Socher, R.; et al. Large language models generate functional protein sequences across diverse families. Nat. Biotechnol. 2023, 41, 1099–1106. [Google Scholar] [CrossRef]
  110. Nijkamp, E.; Ruffolo, J.A.; Weinstein, E.N.; Naik, N.; Madani, A. ProGen2: Exploring the boundaries of protein language models. Cell Syst. 2023, 14, 968–978.e3. [Google Scholar] [CrossRef] [PubMed]
  111. Shen, Y.; Chen, Z.; Mamalakis, M.; Liu, Y.; Li, T.; Su, Y.; He, J.; Liò, P.; Wang, Y.G. TourSynbio: A multi-modal large model and agent framework to bridge text and protein sequences for protein engineering. arXiv 2024. [Google Scholar] [CrossRef]
  112. Orlando, G.; Raimondi, D.; Vranken, W.F. Observation selection bias in contact prediction and its implications for structural bioinformatics. Sci. Rep. 2016, 6, 36679. [Google Scholar] [CrossRef] [PubMed]
  113. Derry, A.; Carpenter, K.A.; Altman, R.B. Training data composition affects performance of protein structure analysis algorithms. Pac. Symp. Biocomput. 2022, 27, 10–21. [Google Scholar] [CrossRef]
  114. Schmirler, R.; Heinzinger, M.; Rost, B. Fine-tuning protein language models boosts predictions across diverse tasks. Nat. Commun. 2024, 15, 7407. [Google Scholar] [CrossRef]
  115. Chen, B.; Cheng, X.; Li, P.; Geng, Y.; Gong, J.; Li, S.; Bei, Z.; Tan, X.; Wang, B.; Zeng, X.; et al. xTrimoPGLM: Unified 100-billion-parameter pretrained transformer for deciphering the language of proteins. Nat. Methods 2025, 22, 1028–1039. [Google Scholar] [CrossRef] [PubMed]
  116. Edsall, J.T. The molecular basis of evolution (Anfinsen, Christian B.). J. Chem. Educ. 1960, 37, 107. [Google Scholar] [CrossRef][Green Version]
  117. Zhou, J.; Panaitiu, A.E.; Grigoryan, G. A general-purpose protein design framework based on mining sequence–structure relationships in known protein structures. Proc. Natl. Acad. Sci. USA 2020, 117, 1059–1068. [Google Scholar] [CrossRef]
  118. Campbell, A.; Yim, J.; Barzilay, R.; Rainforth, T.; Jaakkola, T. Generative flows on discrete state-spaces: Enabling multimodal flows with applications to protein co-design. arXiv 2024. [Google Scholar] [CrossRef]
  119. Lisanza, S.L.; Gershon, J.M.; Tipps, S.W.K.; Sims, J.N.; Arnoldt, L.; Hendel, S.J.; Simma, M.K.; Liu, G.; Yase, M.; Wu, H.; et al. Multistate and functional protein design using RoseTTAFold sequence space diffusion. Nat. Biotechnol. 2024, 43, 1288–1298. [Google Scholar] [CrossRef]
  120. Hayes, T.; Rao, R.; Akin, H.; Sofroniew, N.J.; Oktay, D.; Lin, Z.; Verkuil, R.; Tran, V.Q.; Deaton, J.; Wiggert, M.; et al. Simulating 500 million years of evolution with a language model. Science 2025, 387, 850–858. [Google Scholar] [CrossRef]
  121. Dai, F.; Fan, Y.; Su, J.; Wang, C.; Han, C.; Zhou, X.; Liu, J.; Qian, H.; Wang, S.; Zeng, A.; et al. Toward de novo protein design from natural language. bioRxiv 2024. [Google Scholar] [CrossRef]
  122. Wu, K.E.; Yang, K.K.; Van Den Berg, R.; Alamdari, S.; Zou, J.Y.; Lu, A.X.; Amini, A.P. Protein structure generation via folding diffusion. Nat. Commun. 2024, 15, 1059. [Google Scholar] [CrossRef]
  123. Liu, Y.; Wang, S.; Dong, J.; Chen, L.; Wang, X.; Wang, L.; Li, F.; Wang, C.; Zhang, J.; Wang, Y.; et al. De novo protein design with a denoising diffusion network independent of pretrained structure prediction models. Nat. Methods 2024, 21, 2107–2116. [Google Scholar] [CrossRef]
  124. Stark, H.; Jing, B.; Geffner, T.; Yim, J.; Jaakkola, T.; Vahdat, A.; Kreis, K. ProtComposer: Compositional protein structure generation with 3D ellipsoids. arXiv 2025. [Google Scholar] [CrossRef]
  125. Zhang, Y.; Liu, Y.; Ma, Z.; Li, M.; Xu, C.; Gong, H. Improving diffusion-based protein backbone generation with global-geometry-aware latent encoding. Nat. Mach. Intell. 2025, 7, 1104–1118. [Google Scholar] [CrossRef]
  126. Zhou, X.; Chen, G.; Ye, J.; Wang, E.; Zhang, J.; Mao, C.; Li, Z.; Hao, J.; Huang, X.; Tang, J.; et al. ProRefiner: An entropy-based refining strategy for inverse protein folding with global graph attention. Nat. Commun. 2023, 14, 7434. [Google Scholar] [CrossRef]
  127. Shuai, R.W.; Widatalla, T.; Huang, P.-S.; Hie, B.L. Sidechain conditioning and modeling for full-atom protein sequence design with FAMPNN. bioRxiv 2025. [Google Scholar] [CrossRef]
  128. Jiang, K.; Yan, Z.; Di Bernardo, M.; Sgrizzi, S.R.; Villiger, L.; Kayabolen, A.; Kim, B.J.; Carscadden, J.K.; Hiraizumi, M.; Nishimasu, H.; et al. Rapid in silico directed evolution by a protein language model with EVOLVEpro. Science 2025, 387, eadr6006. [Google Scholar] [CrossRef] [PubMed]
  129. Gligorijević, V.; Renfrew, P.D.; Kosciolek, T.; Leman, J.K.; Berenberg, D.; Vatanen, T.; Chandler, C.; Taylor, B.C.; Fisk, I.M.; Vlamakis, H.; et al. Structure-based protein function prediction using graph convolutional networks. Nat. Commun. 2021, 12, 3168. [Google Scholar] [CrossRef] [PubMed]
  130. Bileschi, M.L.; Belanger, D.; Bryant, D.H.; Sanderson, T.; Carter, B.; Sculley, D.; Bateman, A.; DePristo, M.A.; Colwell, L.J. Using deep learning to annotate the protein universe. Nat. Biotechnol. 2022, 40, 932–937. [Google Scholar] [CrossRef]
  131. Bordin, N.; Dallago, C.; Heinzinger, M.; Kim, S.; Littmann, M.; Rauer, C.; Steinegger, M.; Rost, B.; Orengo, C. Novel machine learning approaches revolutionize protein knowledge. Trends Biochem. Sci. 2023, 48, 345–359. [Google Scholar] [CrossRef]
  132. Wang, W.; Shuai, Y.; Zeng, M.; Fan, W.; Li, M. DPFunc: Accurately predicting protein function via deep learning with domain-guided structure information. Nat. Commun. 2025, 16, 70. [Google Scholar] [CrossRef]
  133. Glögl, M.; Krishnakumar, A.; Ragotte, R.J.; Goreshnik, I.; Coventry, B.; Bera, A.K.; Kang, A.; Joyce, E.; Ahn, G.; Huang, B.; et al. Target-conditioned diffusion generates potent TNFR superfamily antagonists and agonists. Science 2024, 386, 1154–1161. [Google Scholar] [CrossRef] [PubMed]
  134. Vázquez Torres, S.; Benard Valle, M.; Mackessy, S.P.; Menzies, S.K.; Casewell, N.R.; Ahmadi, S.; Burlet, N.J.; Muratspahić, E.; Sappington, I.; Overath, M.D.; et al. De novo designed proteins neutralize lethal snake venom toxins. Nature 2025, 639, 225–231. [Google Scholar] [CrossRef] [PubMed]
  135. Ming, Y.; Wang, W.; Yin, R.; Zeng, M.; Tang, L.; Tang, S.; Li, M. A review of enzyme design in catalytic stability by artificial intelligence. Brief. Bioinform. 2023, 24, bbad065. [Google Scholar] [CrossRef]
  136. Lauko, A.; Pellock, S.J.; Sumida, K.H.; Anishchenko, I.; Juergens, D.; Ahern, W.; Jeung, J.; Shida, A.F.; Hunt, A.; Kalvet, I.; et al. Computational design of serine hydrolases. Science 2025, 388, eadu2454. [Google Scholar] [CrossRef]
  137. Zhang, J.Z.; Nguyen, W.H.; Greenwood, N.; Rose, J.C.; Ong, S.-E.; Maly, D.J.; Baker, D. Computationally designed sensors detect endogenous Ras activity and signaling effectors at subcellular resolution. Nat. Biotechnol. 2024, 42, 1888–1898. [Google Scholar] [CrossRef]
  138. Zhang, J.Z.; Ong, S.-E.; Baker, D.; Maly, D.J. Single-cell sensor analyses reveal signaling programs enabling Ras-G12C drug resistance. Nat. Chem. Biol. 2025, 21, 47–58. [Google Scholar] [CrossRef] [PubMed]
  139. Huddy, T.F.; Hsia, Y.; Kibler, R.D.; Xu, J.; Bethel, N.; Nagarajan, D.; Redler, R.; Leung, P.J.Y.; Weidle, C.; Courbet, A.; et al. Blueprinting extendable nanomaterials with standardized protein blocks. Nature 2024, 627, 898–904. [Google Scholar] [CrossRef]
  140. Pillai, A.; Idris, A.; Philomin, A.; Weidle, C.; Skotheim, R.; Leung, P.J.Y.; Broerman, A.; Demakis, C.; Borst, A.J.; Praetorius, F.; et al. De novo design of allosterically switchable protein assemblies. Nature 2024, 632, 911–920. [Google Scholar] [CrossRef]
  141. Hou, K.; Huang, W.; Qi, M.; Tugwell, T.H.; Alturaifi, T.M.; Chen, Y.; Zhang, X.; Lu, L.; Mann, S.I.; Liu, P.; et al. De novo design of porphyrin-containing proteins as efficient and stereoselective catalysts. Science 2025, 388, 665–670. [Google Scholar] [CrossRef]
  142. Poelwijk, F.J.; Kiviet, D.J.; Weinreich, D.M.; Tans, S.J. Empirical fitness landscapes reveal accessible evolutionary paths. Nature 2007, 445, 383–386. [Google Scholar] [CrossRef]
  143. Starr, T.N.; Thornton, J.W. Epistasis in protein evolution. Protein Sci. 2016, 25, 1204–1218. [Google Scholar] [CrossRef] [PubMed]
  144. Smith, J.M. Natural selection and the concept of a protein space. Nature 1970, 225, 563–564. [Google Scholar] [CrossRef]
  145. Fei, H.; Li, Y.; Liu, Y.; Wei, J.; Chen, A.; Gao, C. Advancing protein evolution with inverse folding models integrating structural and evolutionary constraints. Cell 2025, 188, 4674–4692. [Google Scholar] [CrossRef]
  146. Biswas, S.; Khimulya, G.; Alley, E.C.; Esvelt, K.M.; Church, G.M. Low-N protein engineering with data-efficient deep learning. Nat. Methods 2021, 18, 389–396. [Google Scholar] [CrossRef]
  147. Wang, J.; Lisanza, S.; Juergens, D.; Tischer, D.; Watson, J.L.; Castro, K.M.; Ragotte, R.; Saragovi, A.; Milles, L.F.; Baek, M.; et al. Scaffolding protein functional sites using deep learning. Science 2022, 377, 387–394. [Google Scholar] [CrossRef] [PubMed]
  148. Mahling, R.; Hegyi, B.; Cullen, E.R.; Cho, T.M.; Rodriques, A.R.; Fossier, L.; Yehya, M.; Yang, L.; Chen, B.-X.; Katchman, A.N.; et al. De novo design of a peptide modulator to reverse sodium channel dysfunction linked to cardiac arrhythmias and epilepsy. Cell S0092-8674(25)00860-8. [CrossRef]
  149. Listov, D.; Vos, E.; Hoffka, G.; Hoch, S.Y.; Berg, A.; Hamer-Rogotner, S.; Dym, O.; Kamerlin, S.C.L.; Fleishman, S.J. Complete computational design of high-efficiency Kemp elimination enzymes. Nature 2025, 643, 1421–1427. [Google Scholar] [CrossRef]
  150. Munsamy, G.; Illanes-Vicioso, R.; Funcillo, S.; Nakou, I.T.; Lindner, S.; Ayres, G.; Sheehan, L.S.; Moss, S.; Eckhard, U.; Lorenz, P.; et al. Conditional language models enable the efficient design of proficient enzymes. bioRxiv 2024. [Google Scholar] [CrossRef]
  151. Kim, D.; Woodbury, S.M.; Ahern, W.; Tischer, D.; Hanikel, N.; Salike, S.; Yim, J.; Pellock, S.J.; Lauko, A.; Kalvet, I.; et al. Computational design of metallohydrolases. bioRxiv 2024. [Google Scholar] [CrossRef]
  152. Anishchenko, I.; Kipnis, Y.; Kalvet, I.; Zhou, G.; Krishna, R.; Pellock, S.J.; Lauko, A.; Lee, G.R.; An, L.; Dauparas, J.; et al. Modeling protein-small molecule conformational ensembles with ChemNet. bioRxiv 2024. [Google Scholar] [CrossRef]
  153. Liu, Z.; Zhao, Z.; Xie, L.; Xiao, Z.; Li, M.; Li, Y.; Luo, T. Proteomic analysis reveals chromatin remodeling as a potential therapeutical target in neuroblastoma. J. Transl. Med. 2025, 23, 234. [Google Scholar] [CrossRef]
  154. Zhang, G.; Song, C.; Yin, M.; Liu, L.; Zhang, Y.; Li, Y.; Zhang, J.; Guo, M.; Li, C. TRAPT: A multi-stage fused deep learning framework for predicting transcriptional regulators based on large-scale epigenomic data. Nat. Commun. 2025, 16, 3611. [Google Scholar] [CrossRef]
  155. Wang, M.; Zhang, Z.; Singh Bedi, A.; Guerra, S.; Lin-Gibson, S.; Cong, L.; Chakraborty, S.; Qu, Y.; Ma, J.; Xing, E.; et al. A call for built-in biosecurity safeguards for generative AI tools. Preprint 2025, 43, 845–847. [Google Scholar] [CrossRef] [PubMed]
  156. Irbäck, A.; Knuthson, L.; Mohanty, S.; Peterson, C. Using quantum annealing to design lattice proteins. Phys. Rev. Res. 2024, 6, 13162. [Google Scholar] [CrossRef]
  157. Pandey, M.; Fernandez, M.; Gentile, F.; Isayev, O.; Tropsha, A.; Stern, A.C.; Cherkasov, A. The transformational role of GPU computing and deep learning in drug discovery. Nat. Mach. Intell. 2022, 4, 211–221. [Google Scholar] [CrossRef]
  158. Lee, T.-S.; Cerutti, D.S.; Mermelstein, D.; Lin, C.; LeGrand, S.; Giese, T.J.; Roitberg, A.; Case, D.A.; Walker, R.C.; York, D.M. GPU-accelerated molecular dynamics and free energy methods in Amber18: Performance enhancements and new features. J. Chem. Inf. Model. 2018, 58, 2043–2050. [Google Scholar] [CrossRef]
Figure 1. Schematic of the protein functional universe. The vast, high-dimensional mapping between sequence and structure spaces defines an immense protein functional universe. The yellow region denotes the sequence/structure/function space already explored via natural evolution and experimental characterization; the shaded gray region denotes largely unexplored regions of the sequence/structure/functional space. Each circle represents an individual protein characterized by its sequence, structure, and function. Red circles indicate proteins discovered and characterized via AI-driven de novo protein design; blue circles indicate proteins obtained and characterized through traditional protein engineering or evolutionary methods; gray circles indicate sequences/structures/functions that remain unknown or uncharacterized. Traditional engineering and evolution largely sample the well-explored neighborhood, whereas AI-driven de novo design can systematically probe and populate distant, previously inaccessible regions, enabling the discovery of novel sequences, folds, and functions.
Figure 1. Schematic of the protein functional universe. The vast, high-dimensional mapping between sequence and structure spaces defines an immense protein functional universe. The yellow region denotes the sequence/structure/function space already explored via natural evolution and experimental characterization; the shaded gray region denotes largely unexplored regions of the sequence/structure/functional space. Each circle represents an individual protein characterized by its sequence, structure, and function. Red circles indicate proteins discovered and characterized via AI-driven de novo protein design; blue circles indicate proteins obtained and characterized through traditional protein engineering or evolutionary methods; gray circles indicate sequences/structures/functions that remain unknown or uncharacterized. Traditional engineering and evolution largely sample the well-explored neighborhood, whereas AI-driven de novo design can systematically probe and populate distant, previously inaccessible regions, enabling the discovery of novel sequences, folds, and functions.
Biology 14 01268 g001
Figure 2. Illustration of the classic sequence → structure → function paradigm and how de novo design inverts it. (A) In the classical view, permutations of the 20 amino acids generate a vast sequence space, (B) which gives rise to a diverse structural/fold space (C) and thereby underpins functional diversity. De novo protein design fundamentally inverts the classic sequence → structure → function paradigm: it starts from a desired function and works backward to derive compatible structures and sequences. Current methodologies fall into three categories: (1) Two-Stage Generative Design, (2) Sequence-Guided Language Methods, and (3) Sequence–Structure Co-Guided Methods.
Figure 2. Illustration of the classic sequence → structure → function paradigm and how de novo design inverts it. (A) In the classical view, permutations of the 20 amino acids generate a vast sequence space, (B) which gives rise to a diverse structural/fold space (C) and thereby underpins functional diversity. De novo protein design fundamentally inverts the classic sequence → structure → function paradigm: it starts from a desired function and works backward to derive compatible structures and sequences. Current methodologies fall into three categories: (1) Two-Stage Generative Design, (2) Sequence-Guided Language Methods, and (3) Sequence–Structure Co-Guided Methods.
Biology 14 01268 g002
Figure 3. According to their roles in de novo protein design, we categorize the AI-driven de novo protein design toolbox into five classes: (A) Protein Structure Prediction models; (B) De novo Backbone Generation models; (C) “Fixed-backbone” Sequence Design models; (D) Sequence Generation models; (E) Sequence–structure co-design models.
Figure 3. According to their roles in de novo protein design, we categorize the AI-driven de novo protein design toolbox into five classes: (A) Protein Structure Prediction models; (B) De novo Backbone Generation models; (C) “Fixed-backbone” Sequence Design models; (D) Sequence Generation models; (E) Sequence–structure co-design models.
Biology 14 01268 g003
Figure 4. AI-driven de novo protein design can explore the structure space to uncover novel folds and topologies, engineer bespoke functional sites for defined activities, and thereby perform the global exploration of the sequence–structure–function landscape. Free from evolutionary constraints, it enables access to entirely new protein functions and more readily reaches global fitness optima—innovations that have already powered diverse synthetic biology applications.
Figure 4. AI-driven de novo protein design can explore the structure space to uncover novel folds and topologies, engineer bespoke functional sites for defined activities, and thereby perform the global exploration of the sequence–structure–function landscape. Free from evolutionary constraints, it enables access to entirely new protein functions and more readily reaches global fitness optima—innovations that have already powered diverse synthetic biology applications.
Biology 14 01268 g004
Table 1. Quick workflows (Design goal → inputs → outputs summaries) for AI-driven de novo protein-design toolkits.
Table 1. Quick workflows (Design goal → inputs → outputs summaries) for AI-driven de novo protein-design toolkits.
ToolkitGoalInputsOutputs
Protein structure prediction (Section 4.1)Produce 3D models and confidence estimates for single-chain proteins or complexes to assess foldability and guide design.One or more amino-acid sequences, optional partner sequences, ligands, oligomeric state, templates.Predicted coordinates (PDB); per-residue and global confidence scores (pLDDT, pTM, PAE); interface metrics for complexes.
De novo backbone generation (Section 4.2)Generate novel backbone geometries or scaffolds that satisfy specified geometric/functional constraints.Design constraints (motif/active sites coordinates, desired topology/symmetry, pocket geometry, binder, anchor residues).Ensemble of candidate backbone coordinates (atomic models).
‘Fixed-backbone’ sequence design (Section 4.3)Design sequences that fold to a given backbone and meet developability.Target backbone coordinates (PDB); optional side-chain/motif constraints.Ranked sets of candidate sequences.
Sequence generation (Section 4.4)Produce diverse candidate sequences de novo (unconditionally or conditionally guided).Conditioning information (family or functional labels, motif, structural constraints, or descriptor prompts) and sampling parameters.Batches of candidate amino-acid sequences annotated with model scores, novelty metrics and basic developability annotations.
Sequence–structure co-design (Section 4.5)Jointly generate matched sequence–structure pairs that satisfy functional constraints.Functional constraints (motif geometry, binding interface, text prompt).Paired sequence–structure candidates.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zhang, G.; Liu, C.; Lu, J.; Zhang, S.; Zhu, L. The Role of AI-Driven De Novo Protein Design in the Exploration of the Protein Functional Universe. Biology 2025, 14, 1268. https://doi.org/10.3390/biology14091268

AMA Style

Zhang G, Liu C, Lu J, Zhang S, Zhu L. The Role of AI-Driven De Novo Protein Design in the Exploration of the Protein Functional Universe. Biology. 2025; 14(9):1268. https://doi.org/10.3390/biology14091268

Chicago/Turabian Style

Zhang, Guohao, Chuanyang Liu, Jiajie Lu, Shaowei Zhang, and Lingyun Zhu. 2025. "The Role of AI-Driven De Novo Protein Design in the Exploration of the Protein Functional Universe" Biology 14, no. 9: 1268. https://doi.org/10.3390/biology14091268

APA Style

Zhang, G., Liu, C., Lu, J., Zhang, S., & Zhu, L. (2025). The Role of AI-Driven De Novo Protein Design in the Exploration of the Protein Functional Universe. Biology, 14(9), 1268. https://doi.org/10.3390/biology14091268

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

Article Metrics

Back to TopTop