Drug design

Last updated

Drug design, often referred to as rational drug design or simply rational design, is the inventive process of finding new medications based on the knowledge of a biological target. [1] The drug is most commonly an organic small molecule that activates or inhibits the function of a biomolecule such as a protein, which in turn results in a therapeutic benefit to the patient. In the most basic sense, drug design involves the design of molecules that are complementary in shape and charge to the biomolecular target with which they interact and therefore will bind to it. Drug design frequently but not necessarily relies on computer modeling techniques. [2] This type of modeling is sometimes referred to as computer-aided drug design. Finally, drug design that relies on the knowledge of the three-dimensional structure of the biomolecular target is known as structure-based drug design. [2] In addition to small molecules, biopharmaceuticals including peptides [3] [4] and especially therapeutic antibodies are an increasingly important class of drugs and computational methods for improving the affinity, selectivity, and stability of these protein-based therapeutics have also been developed. [5]

Contents

Definition

The phrase "drug design" is similar to ligand design (i.e., design of a molecule that will bind tightly to its target). [6] Although design techniques for prediction of binding affinity are reasonably successful, there are many other properties, such as bioavailability, metabolic half-life, and side effects, that first must be optimized before a ligand can become a safe and effictive drug. These other characteristics are often difficult to predict with rational design techniques.

Due to high attrition rates, especially during clinical phases of drug development, more attention is being focused early in the drug design process on selecting candidate drugs whose physicochemical properties are predicted to result in fewer complications during development and hence more likely to lead to an approved, marketed drug. [7] Furthermore, in vitro experiments complemented with computation methods are increasingly used in early drug discovery to select compounds with more favorable ADME (absorption, distribution, metabolism, and excretion) and toxicological profiles. [8]

Drug targets

A biomolecular target (most commonly a protein or a nucleic acid) is a key molecule involved in a particular metabolic or signaling pathway that is associated with a specific disease condition or pathology or to the infectivity or survival of a microbial pathogen. Potential drug targets are not necessarily disease causing but must by definition be disease modifying. [9] In some cases, small molecules will be designed to enhance or inhibit the target function in the specific disease modifying pathway. Small molecules (for example receptor agonists, antagonists, inverse agonists, or modulators; enzyme activators or inhibitors; or ion channel openers or blockers) [10] will be designed that are complementary to the binding site of target. [11] Small molecules (drugs) can be designed so as not to affect any other important "off-target" molecules (often referred to as antitargets) since drug interactions with off-target molecules may lead to undesirable side effects. [12] Due to similarities in binding sites, closely related targets identified through sequence homology have the highest chance of cross reactivity and hence highest side effect potential.

Most commonly, drugs are organic small molecules produced through chemical synthesis, but biopolymer-based drugs (also known as biopharmaceuticals) produced through biological processes are becoming increasingly more common. [13] In addition, mRNA-based gene silencing technologies may have therapeutic applications. [14] For example, nanomedicines based on mRNA can streamline and expedite the drug development process, enabling transient and localized expression of immunostimulatory molecules. [15] In vitro transcribed (IVT) mRNA allows for delivery to various accessible cell types via the blood or alternative pathways. The use of IVT mRNA serves to convey specific genetic information into a person's cells, with the primary objective of preventing or altering a particular disease. [16]

Drug discovery

Phenotypic drug discovery

Phenotypic drug discovery is a traditional drug discovery method, also known as forward pharmacology or classical pharmacology. It uses the process of phenotypic screening on collections of synthetic small molecules, natural products, or extracts within chemical libraries to pinpoint substances exhibiting beneficial therapeutic effects. This method is to first discover the in vivo or in vitro functional activity of drugs (such as extract drugs or natural products), and then perform target identification. Phenotypic discovery uses a practical and target-independent approach to generate initial leads, aiming to discover pharmacologically active compounds and therapeutics that operate through novel drug mechanisms. [17] This method allows the exploration of disease phenotypes to find potential treatments for conditions with unknown, complex, or multifactorial origins, where the understanding of molecular targets is insufficient for effective intervention. [18]

Rational drug discovery

Rational drug design (also called reverse pharmacology) begins with a hypothesis that modulation of a specific biological target may have therapeutic value. In order for a biomolecule to be selected as a drug target, two essential pieces of information are required. The first is evidence that modulation of the target will be disease modifying. This knowledge may come from, for example, disease linkage studies that show an association between mutations in the biological target and certain disease states. [19] The second is that the target is capable of binding to a small molecule and that its activity can be modulated by the small molecule. [20]

Once a suitable target has been identified, the target is normally cloned and produced and purified. The purified protein is then used to establish a screening assay. In addition, the three-dimensional structure of the target may be determined.

The search for small molecules that bind to the target is begun by screening libraries of potential drug compounds. This may be done by using the screening assay (a "wet screen"). In addition, if the structure of the target is available, a virtual screen may be performed of candidate drugs. Ideally, the candidate drug compounds should be "drug-like", that is they should possess properties that are predicted to lead to oral bioavailability, adequate chemical and metabolic stability, and minimal toxic effects. [21] Several methods are available to estimate druglikeness such as Lipinski's Rule of Five and a range of scoring methods such as lipophilic efficiency. [22] Several methods for predicting drug metabolism have also been proposed in the scientific literature. [23]

Due to the large number of drug properties that must be simultaneously optimized during the design process, multi-objective optimization techniques are sometimes employed. [24] Finally because of the limitations in the current methods for prediction of activity, drug design is still very much reliant on serendipity [25] and bounded rationality. [26]

Computer-aided drug design

The most fundamental goal in drug design is to predict whether a given molecule will bind to a target and if so how strongly. Molecular mechanics or molecular dynamics is most often used to estimate the strength of the intermolecular interaction between the small molecule and its biological target. These methods are also used to predict the conformation of the small molecule and to model conformational changes in the target that may occur when the small molecule binds to it. [3] [4] Semi-empirical, ab initio quantum chemistry methods, or density functional theory are often used to provide optimized parameters for the molecular mechanics calculations and also provide an estimate of the electronic properties (electrostatic potential, polarizability, etc.) of the drug candidate that will influence binding affinity. [27]

Molecular mechanics methods may also be used to provide semi-quantitative prediction of the binding affinity. Also, knowledge-based scoring function may be used to provide binding affinity estimates. These methods use linear regression, machine learning, neural nets or other statistical techniques to derive predictive binding affinity equations by fitting experimental affinities to computationally derived interaction energies between the small molecule and the target. [28] [29]

Ideally, the computational method will be able to predict affinity before a compound is synthesized and hence in theory only one compound needs to be synthesized, saving enormous time and cost. The reality is that present computational methods are imperfect and provide, at best, only qualitatively accurate estimates of affinity. In practice, it requires several iterations of design, synthesis, and testing before an optimal drug is discovered. Computational methods have accelerated discovery by reducing the number of iterations required and have often provided novel structures. [30] [31]

Computer-aided drug design may be used at any of the following stages of drug discovery:

  1. hit identification using virtual screening (structure- or ligand-based design)
  2. hit-to-lead optimization of affinity and selectivity (structure-based design, QSAR, etc.)
  3. lead optimization of other pharmaceutical properties while maintaining affinity
Flowchart of a Usual Clustering Analysis for Structure-Based Drug Design Wiki Clustering.png
Flowchart of a Usual Clustering Analysis for Structure-Based Drug Design

In order to overcome the insufficient prediction of binding affinity calculated by recent scoring functions, the protein-ligand interaction and compound 3D structure information are used for analysis. For structure-based drug design, several post-screening analyses focusing on protein-ligand interaction have been developed for improving enrichment and effectively mining potential candidates:

Types

Drug discovery cycle highlighting both ligand-based (indirect) and structure-based (direct) drug design strategies. Drug discovery cycle 2.png
Drug discovery cycle highlighting both ligand-based (indirect) and structure-based (direct) drug design strategies.

There are two major types of drug design. The first is referred to as ligand-based drug design and the second, structure-based drug design. [2]

Ligand-based

Ligand-based drug design (or indirect drug design) relies on knowledge of other molecules that bind to the biological target of interest. These other molecules may be used to derive a pharmacophore model that defines the minimum necessary structural characteristics a molecule must possess in order to bind to the target. [36] A model of the biological target may be built based on the knowledge of what binds to it, and this model in turn may be used to design new molecular entities that interact with the target. Alternatively, a quantitative structure-activity relationship (QSAR), in which a correlation between calculated properties of molecules and their experimentally determined biological activity, may be derived. These QSAR relationships in turn may be used to predict the activity of new analogs. [37]

Structure-based

Structure-based drug design (or direct drug design) relies on knowledge of the three dimensional structure of the biological target obtained through methods such as x-ray crystallography or NMR spectroscopy. [38] If an experimental structure of a target is not available, it may be possible to create a homology model of the target based on the experimental structure of a related protein. Using the structure of the biological target, candidate drugs that are predicted to bind with high affinity and selectivity to the target may be designed using interactive graphics and the intuition of a medicinal chemist. Alternatively, various automated computational procedures may be used to suggest new drug candidates. [39]

Current methods for structure-based drug design can be divided roughly into three main categories. [40] The first method is identification of new ligands for a given receptor by searching large databases of 3D structures of small molecules to find those fitting the binding pocket of the receptor using fast approximate docking programs. This method is known as virtual screening.

A second category is de novo design of new ligands. In this method, ligand molecules are built up within the constraints of the binding pocket by assembling small pieces in a stepwise manner. These pieces can be either individual atoms or molecular fragments. The key advantage of such a method is that novel structures, not contained in any database, can be suggested. [41] [42] [43] A third method is the optimization of known ligands by evaluating proposed analogs within the binding cavity. [40]

Binding site identification

Binding site identification is the first step in structure based design. [20] [44] If the structure of the target or a sufficiently similar homolog is determined in the presence of a bound ligand, then the ligand should be observable in the structure in which case location of the binding site is trivial. However, there may be unoccupied allosteric binding sites that may be of interest. Furthermore, it may be that only apoprotein (protein without ligand) structures are available and the reliable identification of unoccupied sites that have the potential to bind ligands with high affinity is non-trivial. In brief, binding site identification usually relies on identification of concave surfaces on the protein that can accommodate drug sized molecules that also possess appropriate "hot spots" (hydrophobic surfaces, hydrogen bonding sites, etc.) that drive ligand binding. [20] [44]

Scoring functions

Structure-based drug design attempts to use the structure of proteins as a basis for designing new ligands by applying the principles of molecular recognition. Selective high affinity binding to the target is generally desirable since it leads to more efficacious drugs with fewer side effects. Thus, one of the most important principles for designing or obtaining potential new ligands is to predict the binding affinity of a certain ligand to its target (and known antitargets) and use the predicted affinity as a criterion for selection. [45]

One early general-purposed empirical scoring function to describe the binding energy of ligands to receptors was developed by Böhm. [46] [47] This empirical scoring function took the form:

where:

  • ΔG0 – empirically derived offset that in part corresponds to the overall loss of translational and rotational entropy of the ligand upon binding.
  • ΔGhb – contribution from hydrogen bonding
  • ΔGionic – contribution from ionic interactions
  • ΔGlip – contribution from lipophilic interactions where |Alipo| is surface area of lipophilic contact between the ligand and receptor
  • ΔGrot – entropy penalty due to freezing a rotatable in the ligand bond upon binding

A more general thermodynamic "master" equation is as follows: [48]

where:

  • desolvation – enthalpic penalty for removing the ligand from solvent
  • motion – entropic penalty for reducing the degrees of freedom when a ligand binds to its receptor
  • configuration – conformational strain energy required to put the ligand in its "active" conformation
  • interaction – enthalpic gain for "resolvating" the ligand with its receptor

The basic idea is that the overall binding free energy can be decomposed into independent components that are known to be important for the binding process. Each component reflects a certain kind of free energy alteration during the binding process between a ligand and its target receptor. The Master Equation is the linear combination of these components. According to Gibbs free energy equation, the relation between dissociation equilibrium constant, Kd, and the components of free energy was built.

Various computational methods are used to estimate each of the components of the master equation. For example, the change in polar surface area upon ligand binding can be used to estimate the desolvation energy. The number of rotatable bonds frozen upon ligand binding is proportional to the motion term. The configurational or strain energy can be estimated using molecular mechanics calculations. Finally the interaction energy can be estimated using methods such as the change in non polar surface, statistically derived potentials of mean force, the number of hydrogen bonds formed, etc. In practice, the components of the master equation are fit to experimental data using multiple linear regression. This can be done with a diverse training set including many types of ligands and receptors to produce a less accurate but more general "global" model or a more restricted set of ligands and receptors to produce a more accurate but less general "local" model. [49]

Examples

A particular example of rational drug design involves the use of three-dimensional information about biomolecules obtained from such techniques as X-ray crystallography and NMR spectroscopy. Computer-aided drug design in particular becomes much more tractable when there is a high-resolution structure of a target protein bound to a potent ligand. This approach to drug discovery is sometimes referred to as structure-based drug design. The first unequivocal example of the application of structure-based drug design leading to an approved drug is the carbonic anhydrase inhibitor dorzolamide, which was approved in 1995. [50] [51]

Another case study in rational drug design is imatinib, a tyrosine kinase inhibitor designed specifically for the bcr-abl fusion protein that is characteristic for Philadelphia chromosome-positive leukemias (chronic myelogenous leukemia and occasionally acute lymphocytic leukemia). Imatinib is substantially different from previous drugs for cancer, as most agents of chemotherapy simply target rapidly dividing cells, not differentiating between cancer cells and other tissues. [52]

Additional examples include:

Drug screening

Types of drug screening include phenotypic screening, high-throughput screening, and virtual screening. Phenotypic screening is characterized by the process of screening drugs using cellular or animal disease models to identify compounds that alter the phenotype and produce beneficial disease-related effects. [54] [55]  Emerging technologies in high-throughput screening substantially enhance processing speed and decrease the required detection volume. [56] Virtual screening is completed by computer, enabling a large number of molecules can be screened with a short cycle and low cost. Virtual screening uses a range of computational methods that empower chemists to reduce extensive virtual libraries into more manageable sizes. [57]

Case studies

Criticism

It has been argued that the highly rigid and focused nature of rational drug design suppresses serendipity in drug discovery. [58]

See also

Related Research Articles

<span class="mw-page-title-main">Allosteric regulation</span> Regulation of enzyme activity

In biochemistry, allosteric regulation is the regulation of an enzyme by binding an effector molecule at a site other than the enzyme's active site.

<span class="mw-page-title-main">Binding site</span> Molecule-specific coordinate bonding area in biological systems

In biochemistry and molecular biology, a binding site is a region on a macromolecule such as a protein that binds to another molecule with specificity. The binding partner of the macromolecule is often referred to as a ligand. Ligands may include other proteins, enzyme substrates, second messengers, hormones, or allosteric modulators. The binding event is often, but not always, accompanied by a conformational change that alters the protein's function. Binding to protein binding sites is most often reversible, but can also be covalent reversible or irreversible.

<span class="mw-page-title-main">Receptor (biochemistry)</span> Protein molecule receiving signals for a cell

In biochemistry and pharmacology, receptors are chemical structures, composed of protein, that receive and transduce signals that may be integrated into biological systems. These signals are typically chemical messengers which bind to a receptor and produce physiological responses such as change in the electrical activity of a cell. For example, GABA, an inhibitory neurotransmitter, inhibits electrical activity of neurons by binding to GABAA receptors. There are three main ways the action of the receptor can be classified: relay of signal, amplification, or integration. Relaying sends the signal onward, amplification increases the effect of a single ligand, and integration allows the signal to be incorporated into another biochemical pathway.

<span class="mw-page-title-main">Protein–protein interaction</span> Physical interactions and constructions between multiple proteins

Protein–protein interactions (PPIs) are physical contacts of high specificity established between two or more protein molecules as a result of biochemical events steered by interactions that include electrostatic forces, hydrogen bonding and the hydrophobic effect. Many are physical contacts with molecular associations between chains that occur in a cell or in a living organism in a specific biomolecular context.

<span class="mw-page-title-main">Ligand (biochemistry)</span> Substance that forms a complex with a biomolecule

In biochemistry and pharmacology, a ligand is a substance that forms a complex with a biomolecule to serve a biological purpose. The etymology stems from Latin ligare, which means 'to bind'. In protein-ligand binding, the ligand is usually a molecule which produces a signal by binding to a site on a target protein. The binding typically results in a change of conformational isomerism (conformation) of the target protein. In DNA-ligand binding studies, the ligand can be a small molecule, ion, or protein which binds to the DNA double helix. The relationship between ligand and binding partner is a function of charge, hydrophobicity, and molecular structure.

<span class="mw-page-title-main">Docking (molecular)</span> Prediction method in molecular modeling

In the field of molecular modeling, docking is a method which predicts the preferred orientation of one molecule to a second when a ligand and a target are bound to each other to form a stable complex. Knowledge of the preferred orientation in turn may be used to predict the strength of association or binding affinity between two molecules using, for example, scoring functions.

<span class="mw-page-title-main">Enzyme inhibitor</span> Molecule that blocks enzyme activity

An enzyme inhibitor is a molecule that binds to an enzyme and blocks its activity. Enzymes are proteins that speed up chemical reactions necessary for life, in which substrate molecules are converted into products. An enzyme facilitates a specific chemical reaction by binding the substrate to its active site, a specialized area on the enzyme that accelerates the most difficult step of the reaction.

Protein–ligand docking is a molecular modelling technique. The goal of protein–ligand docking is to predict the position and orientation of a ligand when it is bound to a protein receptor or enzyme. Pharmaceutical research employs docking techniques for a variety of purposes, most notably in the virtual screening of large databases of available chemicals in order to select likely drug candidates. There has been rapid development in computational ability to determine protein structure with programs such as AlphaFold, and the demand for the corresponding protein-ligand docking predictions is driving implementation of software that can find accurate models. Once the protein folding can be predicted accurately along with how the ligands of various structures will bind to the protein, the ability for drug development to progress at a much faster rate becomes possible.

<span class="mw-page-title-main">Virtual screening</span>

Virtual screening (VS) is a computational technique used in drug discovery to search libraries of small molecules in order to identify those structures which are most likely to bind to a drug target, typically a protein receptor or enzyme.

In the fields of computational chemistry and molecular modelling, scoring functions are mathematical functions used to approximately predict the binding affinity between two molecules after they have been docked. Most commonly one of the molecules is a small organic compound such as a drug and the second is the drug's biological target such as a protein receptor. Scoring functions have also been developed to predict the strength of intermolecular interactions between two proteins or between protein and DNA.

Molecular binding is an attractive interaction between two molecules that results in a stable association in which the molecules are in close proximity to each other. It is formed when atoms or molecules bind together by sharing of electrons. It often, but not always, involves some chemical bonding.

Fragment-based lead discovery (FBLD) also known as fragment-based drug discovery (FBDD) is a method used for finding lead compounds as part of the drug discovery process. Fragments are small organic molecules which are small in size and low in molecular weight. It is based on identifying small chemical fragments, which may bind only weakly to the biological target, and then growing them or combining them to produce a lead with a higher affinity. FBLD can be compared with high-throughput screening (HTS). In HTS, libraries with up to millions of compounds, with molecular weights of around 500 Da, are screened, and nanomolar binding affinities are sought. In contrast, in the early phase of FBLD, libraries with a few thousand compounds with molecular weights of around 200 Da may be screened, and millimolar affinities can be considered useful. FBLD is a technique being used in research for discovering novel potent inhibitors. This methodology could help to design multitarget drugs for multiple diseases. The multitarget inhibitor approach is based on designing an inhibitor for the multiple targets. This type of drug design opens up new polypharmacological avenues for discovering innovative and effective therapies. Neurodegenerative diseases like Alzheimer’s (AD) and Parkinson’s, among others, also show rather complex etiopathologies. Multitarget inhibitors are more appropriate for addressing the complexity of AD and may provide new drugs for controlling the multifactorial nature of AD, stopping its progression.

<span class="mw-page-title-main">Cell surface receptor</span> Class of ligand activated receptors localized in surface of plama cell membrane

Cell surface receptors are receptors that are embedded in the plasma membrane of cells. They act in cell signaling by receiving extracellular molecules. They are specialized integral membrane proteins that allow communication between the cell and the extracellular space. The extracellular molecules may be hormones, neurotransmitters, cytokines, growth factors, cell adhesion molecules, or nutrients; they react with the receptor to induce changes in the metabolism and activity of a cell. In the process of signal transduction, ligand binding affects a cascading chemical change through the cell membrane.

Druggability is a term used in drug discovery to describe a biological target that is known to or is predicted to bind with high affinity to a drug. Furthermore, by definition, the binding of the drug to a druggable target must alter the function of the target with a therapeutic benefit to the patient. The concept of druggability is most often restricted to small molecules but also has been extended to include biologic medical products such as therapeutic monoclonal antibodies.

A thermal shift assay (TSA) measures changes in the thermal denaturation temperature and hence stability of a protein under varying conditions such as variations in drug concentration, buffer pH or ionic strength, redox potential, or sequence mutation. The most common method for measuring protein thermal shifts is differential scanning fluorimetry (DSF) or thermofluor, which utilizes specialized fluorogenic dyes.

A ligand binding assay (LBA) is an assay, or an analytic procedure, which relies on the binding of ligand molecules to receptors, antibodies or other macromolecules. A detection method is used to determine the presence and extent of the ligand-receptor complexes formed, and this is usually determined electrochemically or through a fluorescence detection method. This type of analytic test can be used to test for the presence of target molecules in a sample that are known to bind to the receptor.

<span class="mw-page-title-main">Targeted covalent inhibitors</span>

Targeted covalent inhibitors (TCIs) or Targeted covalent drugs are rationally designed inhibitors that bind and then bond to their target proteins. These inhibitors possess a bond-forming functional group of low chemical reactivity that, following binding to the target protein, is positioned to react rapidly with a proximate nucleophilic residue at the target site to form a bond.

Chemoproteomics entails a broad array of techniques used to identify and interrogate protein-small molecule interactions. Chemoproteomics complements phenotypic drug discovery, a paradigm that aims to discover lead compounds on the basis of alleviating a disease phenotype, as opposed to target-based drug discovery, in which lead compounds are designed to interact with predetermined disease-driving biological targets. As phenotypic drug discovery assays do not provide confirmation of a compound's mechanism of action, chemoproteomics provides valuable follow-up strategies to narrow down potential targets and eventually validate a molecule's mechanism of action. Chemoproteomics also attempts to address the inherent challenge of drug promiscuity in small molecule drug discovery by analyzing protein-small molecule interactions on a proteome-wide scale. A major goal of chemoproteomics is to characterize the interactome of drug candidates to gain insight into mechanisms of off-target toxicity and polypharmacology.

Molecular Operating Environment (MOE) is a drug discovery software platform that integrates visualization, modeling and simulations, as well as methodology development, in one package. MOE scientific applications are used by biologists, medicinal chemists and computational chemists in pharmaceutical, biotechnology and academic research. MOE runs on Windows, Linux, Unix, and macOS. Main application areas in MOE include structure-based design, fragment-based design, ligand-based design, pharmacophore discovery, medicinal chemistry applications, biologics applications, structural biology and bioinformatics, protein and antibody modeling, molecular modeling and simulations, virtual screening, cheminformatics & QSAR. The Scientific Vector Language (SVL) is the built-in command, scripting and application development language of MOE.

<span class="mw-page-title-main">Protein–ligand complex</span>

A protein–ligand complex is a complex of a protein bound with a ligand that is formed following molecular recognition between proteins that interact with each other or with other molecules. Formation of a protein-ligand complex is based on molecular recognition between biological macromolecules and ligands, where ligand means any molecule that binds the protein with high affinity and specificity. Molecular recognition is not a process by itself since it is part of a functionally important mechanism involving the essential elements of life like in self-replication, metabolism, and information processing. For example DNA-replication depends on recognition and binding of DNA double helix by helicase, DNA single strand by DNA-polymerase and DNA segments by ligase. Molecular recognition depends on affinity and specificity. Specificity means that proteins distinguish the highly specific binding partner from less specific partners and affinity allows the specific partner with high affinity to remain bound even if there are high concentrations of less specific partners with lower affinity.

References

  1. Madsen U, Krogsgaard-Larsen P, Liljefors T (2002). Textbook of Drug Design and Discovery. Washington, D.C.: Taylor & Francis. ISBN   978-0-415-28288-8.
  2. 1 2 3 Reynolds CH, Merz KM, Ringe D, eds. (2010). Drug Design: Structure- and Ligand-Based Approaches (1 ed.). Cambridge, UK: Cambridge University Press. ISBN   978-0521887236.
  3. 1 2 Fosgerau K, Hoffmann T (January 2015). "Peptide therapeutics: current status and future directions". Drug Discovery Today. 20 (1): 122–128. doi: 10.1016/j.drudis.2014.10.003 . PMID   25450771.
  4. 1 2 Ciemny M, Kurcinski M, Kamel K, Kolinski A, Alam N, Schueler-Furman O, Kmiecik S (August 2018). "Protein-peptide docking: opportunities and challenges". Drug Discovery Today. 23 (8): 1530–1537. doi: 10.1016/j.drudis.2018.05.006 . PMID   29733895.
  5. Shirai H, Prades C, Vita R, Marcatili P, Popovic B, Xu J, et al. (November 2014). "Antibody informatics for drug discovery". Biochimica et Biophysica Acta (BBA) - Proteins and Proteomics. 1844 (11): 2002–2015. doi:10.1016/j.bbapap.2014.07.006. PMID   25110827.
  6. Tollenaere JP (April 1996). "The role of structure-based ligand design and molecular modelling in drug discovery". Pharmacy World & Science. 18 (2): 56–62. doi:10.1007/BF00579706. PMID   8739258. S2CID   21550508.
  7. Waring MJ, Arrowsmith J, Leach AR, Leeson PD, Mandrell S, Owen RM, et al. (July 2015). "An analysis of the attrition of drug candidates from four major pharmaceutical companies". Nature Reviews. Drug Discovery. 14 (7): 475–486. doi:10.1038/nrd4609. PMID   26091267. S2CID   25292436.
  8. Yu H, Adedoyin A (September 2003). "ADME-Tox in drug discovery: integration of experimental and computational technologies". Drug Discovery Today. 8 (18): 852–861. doi:10.1016/S1359-6446(03)02828-9. PMID   12963322.
  9. Dixon SJ, Stockwell BR (December 2009). "Identifying druggable disease-modifying gene products". Current Opinion in Chemical Biology. 13 (5–6): 549–555. doi:10.1016/j.cbpa.2009.08.003. PMC   2787993 . PMID   19740696.
  10. Imming P, Sinning C, Meyer A (October 2006). "Drugs, their targets and the nature and number of drug targets". Nature Reviews. Drug Discovery. 5 (10): 821–834. doi:10.1038/nrd2132. PMID   17016423. S2CID   8872470.
  11. Anderson AC (September 2003). "The process of structure-based drug design". Chemistry & Biology. 10 (9): 787–797. doi: 10.1016/j.chembiol.2003.09.002 . PMID   14522049.
  12. Recanatini M, Bottegoni G, Cavalli A (December 2004). "In silico antitarget screening". Drug Discovery Today. Technologies. 1 (3): 209–215. doi:10.1016/j.ddtec.2004.10.004. PMID   24981487.
  13. Wu-Pong S, Rojanasakul Y (2008). Biopharmaceutical drug design and development (2nd ed.). Totowa, NJ Humana Press: Humana Press. ISBN   978-1-59745-532-9.
  14. Scomparin A, Polyak D, Krivitsky A, Satchi-Fainaro R (November 2015). "Achieving successful delivery of oligonucleotides--From physico-chemical characterization to in vivo evaluation". Biotechnology Advances. 33 (6 Pt 3): 1294–1309. doi:10.1016/j.biotechadv.2015.04.008. PMID   25916823.
  15. Youssef M, Hitti C, Puppin Chaves Fulber J, Kamen AA (October 2023). "Enabling mRNA Therapeutics: Current Landscape and Challenges in Manufacturing". Biomolecules. 13 (10): 1497. doi: 10.3390/biom13101497 . PMC   10604719 . PMID   37892179.
  16. Sahin U, Karikó K, Türeci Ö (October 2014). "mRNA-based therapeutics--developing a new class of drugs". Nature Reviews. Drug Discovery. 13 (10): 759–780. doi: 10.1038/nrd4278 . PMID   25233993. S2CID   27454546.
  17. Swinney DC, Lee JA (2020). "Recent advances in phenotypic drug discovery". F1000Research. 9: F1000 Faculty Rev–944. doi: 10.12688/f1000research.25813.1 . PMC   7431967 . PMID   32850117.
  18. Moffat JG, Vincent F, Lee JA, Eder J, Prunotto M (August 2017). "Opportunities and challenges in phenotypic drug discovery: an industry perspective". Nature Reviews. Drug Discovery. 16 (8): 531–543. doi: 10.1038/nrd.2017.111 . PMID   28685762. S2CID   6180139.
  19. Ganellin CR, Jefferis R, Roberts SM (2013). "The small molecule drug discovery process — from target selection to candidate selection". Introduction to Biological and Small Molecule Drug Research and Development: theory and case studies. Elsevier. ISBN   9780123971760.
  20. 1 2 3 Yuan Y, Pei J, Lai L (Dec 2013). "Binding site detection and druggability prediction of protein targets for structure-based drug design". Current Pharmaceutical Design. 19 (12): 2326–2333. doi:10.2174/1381612811319120019. PMID   23082974.
  21. Rishton GM (January 2003). "Nonleadlikeness and leadlikeness in biochemical screening". Drug Discovery Today. 8 (2): 86–96. doi:10.1016/s1359644602025722. PMID   12565011.
  22. Hopkins AL (2011). "Chapter 25: Pharmacological space". In Wermuth CG (ed.). The Practice of Medicinal Chemistry (3 ed.). Academic Press. pp. 521–527. ISBN   978-0-12-374194-3.
  23. Kirchmair J (2014). Drug Metabolism Prediction. Wiley's Methods and Principles in Medicinal Chemistry. Vol. 63. Wiley-VCH. ISBN   978-3-527-67301-8.
  24. Nicolaou CA, Brown N (September 2013). "Multi-objective optimization methods in drug design". Drug Discovery Today. Technologies. 10 (3): e427–e435. doi:10.1016/j.ddtec.2013.02.001. PMID   24050140.
  25. Ban TA (2006). "The role of serendipity in drug discovery". Dialogues in Clinical Neuroscience. 8 (3): 335–344. doi:10.31887/DCNS.2006.8.3/tban. PMC   3181823 . PMID   17117615.
  26. Ethiraj SK, Levinthal D (Sep 2004). "Bounded Rationality and the Search for Organizational Architecture: An Evolutionary Perspective on the Design of Organizations and Their Evolvability". Administrative Science Quarterly. Sage Publications, Inc. on behalf of the Johnson Graduate School of Management, Cornell University. 49 (3): 404–437. doi:10.2307/4131441. JSTOR   4131441. S2CID   142910916. SSRN   604123.
  27. Lewis RA (2011). "Chapter 4: The Development of Molecular Modelling Programs: The Use and Limitations of Physical Models". In Gramatica P, Livingstone DJ, Davis AM (eds.). Drug Design Strategies: Quantitative Approaches. RSC Drug Discovery. Royal Society of Chemistry. pp. 88–107. doi:10.1039/9781849733410-00088. ISBN   978-1849731669.
  28. Rajamani R, Good AC (May 2007). "Ranking poses in structure-based lead discovery and optimization: current trends in scoring function development". Current Opinion in Drug Discovery & Development. 10 (3): 308–315. PMID   17554857.
  29. de Azevedo WF, Dias R (December 2008). "Computational methods for calculation of ligand-binding affinity". Current Drug Targets. 9 (12): 1031–1039. doi:10.2174/138945008786949405. PMID   19128212.
  30. Singh J, Chuaqui CE, Boriack-Sjodin PA, Lee WC, Pontz T, Corbley MJ, et al. (December 2003). "Successful shape-based virtual screening: the discovery of a potent inhibitor of the type I TGFbeta receptor kinase (TbetaRI)". Bioorganic & Medicinal Chemistry Letters. 13 (24): 4355–4359. doi:10.1016/j.bmcl.2003.09.028. PMID   14643325.
  31. Becker OM, Dhanoa DS, Marantz Y, Chen D, Shacham S, Cheruku S, et al. (June 2006). "An integrated in silico 3D model-driven discovery of a novel, potent, and selective amidosulfonamide 5-HT1A agonist (PRX-00023) for the treatment of anxiety and depression". Journal of Medicinal Chemistry. 49 (11): 3116–3135. doi:10.1021/jm0508641. PMID   16722631.
  32. Liang S, Meroueh SO, Wang G, Qiu C, Zhou Y (May 2009). "Consensus scoring for enriching near-native structures from protein-protein docking decoys". Proteins. 75 (2): 397–403. doi:10.1002/prot.22252. PMC   2656599 . PMID   18831053.
  33. Oda A, Tsuchida K, Takakura T, Yamaotsu N, Hirono S (2006). "Comparison of consensus scoring strategies for evaluating computational models of protein-ligand complexes". Journal of Chemical Information and Modeling. 46 (1): 380–391. doi:10.1021/ci050283k. PMID   16426072.
  34. Deng Z, Chuaqui C, Singh J (January 2004). "Structural interaction fingerprint (SIFt): a novel method for analyzing three-dimensional protein-ligand binding interactions". Journal of Medicinal Chemistry. 47 (2): 337–344. doi:10.1021/jm030331x. PMID   14711306.
  35. Amari S, Aizawa M, Zhang J, Fukuzawa K, Mochizuki Y, Iwasawa Y, et al. (2006). "VISCANA: visualized cluster analysis of protein-ligand interaction based on the ab initio fragment molecular orbital method for virtual ligand screening". Journal of Chemical Information and Modeling. 46 (1): 221–230. doi:10.1021/ci050262q. PMID   16426058.
  36. Guner OF (2000). Pharmacophore Perception, Development, and use in Drug Design. La Jolla, Calif: International University Line. ISBN   978-0-9636817-6-8.
  37. Tropsha A (2010). "QSAR in Drug Discovery". In Reynolds CH, Merz KM, Ringe D (eds.). Drug Design: Structure- and Ligand-Based Approaches (1st ed.). Cambridge, UK: Cambridge University Press. pp. 151–164. ISBN   978-0521887236.
  38. Leach AR, Harren J (2007). Structure-based Drug Discovery. Berlin: Springer. ISBN   978-1-4020-4406-9.
  39. Mauser H, Guba W (May 2008). "Recent developments in de novo design and scaffold hopping". Current Opinion in Drug Discovery & Development. 11 (3): 365–374. PMID   18428090.
  40. 1 2 Klebe G (2000). "Recent developments in structure-based drug design". Journal of Molecular Medicine. 78 (5): 269–281. doi:10.1007/s001090000084. PMID   10954199. S2CID   21314020.
  41. Wang R, Gao Y, Lai L (2000). "LigBuilder: A Multi-Purpose Program for Structure-Based Drug Design". Journal of Molecular Modeling. 6 (7–8): 498–516. doi:10.1007/s0089400060498. S2CID   59482623.
  42. Schneider G, Fechner U (August 2005). "Computer-based de novo design of drug-like molecules". Nature Reviews. Drug Discovery. 4 (8): 649–663. doi:10.1038/nrd1799. PMID   16056391. S2CID   2549851.
  43. Jorgensen WL (March 2004). "The many roles of computation in drug discovery". Science. 303 (5665): 1813–1818. Bibcode:2004Sci...303.1813J. doi:10.1126/science.1096361. PMID   15031495. S2CID   1307935.
  44. 1 2 Leis S, Schneider S, Zacharias M (2010). "In silico prediction of binding sites on proteins". Current Medicinal Chemistry. 17 (15): 1550–1562. doi:10.2174/092986710790979944. PMID   20166931.
  45. Warren GL, Warren SD (2011). "Chapter 16: Scoring Drug-Receptor Interactions". In Gramatica P, Livingstone DJ, Davis AM (eds.). Drug Design Strategies: Quantitative Approaches. RSC Drug Discovery. Royal Society of Chemistry. pp. 440–457. doi:10.1039/9781849733410-00440. ISBN   978-1849731669.
  46. Böhm HJ (June 1994). "The development of a simple empirical scoring function to estimate the binding constant for a protein-ligand complex of known three-dimensional structure". Journal of Computer-Aided Molecular Design. 8 (3): 243–256. Bibcode:1994JCAMD...8..243B. doi:10.1007/BF00126743. PMID   7964925. S2CID   2491616.
  47. Liu J, Wang R (March 2015). "Classification of current scoring functions". Journal of Chemical Information and Modeling. 55 (3): 475–482. doi:10.1021/ci500731a. PMID   25647463.
  48. Murcko MA (December 1995). "Computational methods to predict binding free energy in ligand-receptor complexes". Journal of Medicinal Chemistry. 38 (26): 4953–4967. doi:10.1021/jm00026a001. PMID   8544170.
  49. Gramatica P (2011). "Chapter 17: Modeling Chemicals in the Environment". In Gramatica P, Livingstone DJ, Davis AM (eds.). Drug Design Strategies: Quantitative Approaches. RSC Drug Discovery. Royal Society of Chemistry. p. 466. doi:10.1039/9781849733410-00458. ISBN   978-1849731669.
  50. Greer J, Erickson JW, Baldwin JJ, Varney MD (April 1994). "Application of the three-dimensional structures of protein target molecules in structure-based drug design". Journal of Medicinal Chemistry. 37 (8): 1035–1054. doi:10.1021/jm00034a001. PMID   8164249.
  51. Timmerman H, Gubernator K, Böhm HJ, Mannhold R, Kubinyi H (1998). Structure-based Ligand Design (Methods and Principles in Medicinal Chemistry). Weinheim: Wiley-VCH. ISBN   978-3-527-29343-8.
  52. Capdeville R, Buchdunger E, Zimmermann J, Matter A (July 2002). "Glivec (STI571, imatinib), a rationally developed, targeted anticancer drug". Nature Reviews. Drug Discovery. 1 (7): 493–502. doi:10.1038/nrd839. PMID   12120256. S2CID   2728341.
  53. "AutoDock's role in Developing the First Clinically-Approved HIV Integrase Inhibitor". Press Release. The Scripps Research Institute. 2007-12-17.
  54. Prior M, Chiruta C, Currais A, Goldberg J, Ramsey J, Dargusch R, et al. (July 2014). "Back to the future with phenotypic screening". ACS Chemical Neuroscience. 5 (7): 503–513. doi:10.1021/cn500051h. PMC   4102969 . PMID   24902068.
  55. Kotz J (April 2012). "Phenotypic screening, take two". Science-Business EXchange. 5 (15): 380. doi: 10.1038/scibx.2012.380 . ISSN   1945-3477. S2CID   72519717.
  56. Hertzberg RP, Pope AJ (August 2000). "High-throughput screening: new technology for the 21st century". Current Opinion in Chemical Biology. 4 (4): 445–451. doi:10.1016/S1367-5931(00)00110-1. PMID   10959774.
  57. Walters WP, Stahl MT, Murcko MA (April 1998). "Virtual screening—an overview". Drug Discovery Today. 3 (4): 160–178. doi:10.1016/S1359-6446(97)01163-X.
  58. Klein DF (March 2008). "The loss of serendipity in psychopharmacology". JAMA. 299 (9): 1063–1065. doi:10.1001/jama.299.9.1063. PMID   18319418.