Chemoproteomics

Source: Wikipedia, the free encyclopedia.

Chemoproteomics (also known as chemical proteomics) entails a broad array of techniques used to identify and interrogate

Chemoproteomics assays can be stratified into three basic types. Solution-based approaches involve the use of drug analogs that chemically modify target proteins in

lead optimization. Several targets of high profile drugs have been identified using chemoproteomics, and the continued improvement of mass spectrometer sensitivity and chemical probe technology indicates that chemoproteomics will play a large role in future drug discovery
.

Background

Context

An example quantitative proteomics workflow. Protein extracts from different samples are extracted and digested using trypsin. Separate samples are labeled using individual isobaric tandem mass tags (TMTs), then labeled samples are pooled. The sample origin of each peptide can be discerned from the TMT attached to it. Labeled peptides are then detected and fragmented by LC-MS/MS, and quantified by comparing relative amounts of TMT fragments in each mass spectrum. This image was adapted from BioRender.com.

The conclusion of

overexpression systems are simplistic. Spatially and temporally conditional knock-out/knock-in systems have improved the level of nuance in in vivo analysis of protein function, but still fail to completely parallel the systemic breadth of pharmacological action.[3] For example, drugs often act through multiple mechanisms, and often work best by engaging targets partially.[3] Chemoproteomic tools offer a solution to bridge the gap between a genetic understanding of disease and a pharmacological understanding of drug action
by identifying the many proteins involved in therapeutic success.

Basic tools

The chemoproteomic toolkit is anchored by

LC-MS/MS. For more accurate quantification, different samples can be reacted with isobaric tandem mass tags (TMTs), a form of chemical barcode that allows for sample multiplexing, and then pooled.[5]

Solution-based approaches

A prototypical activity-based protein profiling probe. A covalent warhead and reporter tag are connected by a linker group. The warhead covalently bonds with the active site of an enzyme and the reporter tag is used to enrich or detect the labeled protein. Fluorophosphonate-biotin is an example of an activity-based probe that targets serine hydrolases. It is connected to a biotinylated enrichment handle by an alkane chain. This image was made using BioRender.com.

Broad

biochemical readout
, chemical tools are required to detect drug-protein interactions.

Activity-based protein profiling

Activity-based protein profiling (ABPP, also

polypeptides.[5]

Under the assumption that enzymes vary in their structure, function, and associations depending on a system's

affinity to their targets will prevent binding of the probe, and the degree of probe binding can be used as an indication of compound affinity. Because ABPP probes label classes of enzymes, this approach can also be used to profile drug selectivity, as highly selective compounds will ideally outcompete probes at only a small number of proteins.[8]

A prototypical photoaffinity probe. A drug scaffold acts as the first interaction site between probe and protein. A photoreactive group, here an arylazide, can be activated by light to form a reactive intermediate that bonds with a non-specific site on the protein. A tag can then be used to enrich and identify or image and detect the target. This image was made using BioRender.com.

Photoaffinity labeling

Unlike ABPP, which results in protein labeling upon probe binding,

fluorescent dyes are used when using a gel-based imaging method, such as SDS-PAGE, to validate interaction with a target.[7][9]

Phenylazide, a photoreactive group commonly used in photoaffinity labeling.

Because photoaffinity probes are multifunctional, they are difficult to design. Chemists incorporate the same principles of

buffers, at various pH levels, and in living systems to ensure that labeling occurs only when exposed to light. Activation by light must also be fine-tuned, as radiation can damage cells.[9]

Immobilization-based approaches

LC-MS/MS. This image was adapted from BioRender.com
.

Immobilization-based chemoproteomic techniques encompass variations on microbead-based affinity pull-down, which is similar to immunoprecipitation, and affinity chromatography. In both cases, a solid support is used as an immobilization surface bearing a bait molecule. The bait molecule can be a potential drug if the investigator is trying to identify targets, or a target, such as an immobilized enzyme, if the investigator is screening for small molecules.[10] The bait is exposed to a mixture of potential binding partners, which can be identified after removing non-binding components.

Microbead-based immobilization

lysate.[12]

Hybrid solution- and immobilization-based strategies have been applied, in which

induced fit. Another drawback is non-specific adsorption of both proteins and small molecules to the bead surface, which has the potential to generate false positives.[10]

Affinity chromatography

Affinity chromatography emerged in the 1950s as a rarely used method used to purify enzymes; it has since seen mainstream use and is the oldest among chemoproteomic approaches.

stationary phase, and other compounds pass through the column unretained.[14] In both cases, retained analytes can be eluted
from the column and identified using mass spectrometry. A table of elution strategies is provided below.

Affinity Chromatography Elution Strategies[14]
Strategy Description Mechanism(s) Advantages Disadvantages
Non-specific elution Bound ligands are released after a change in
mobile phase
.
polarity
.
Can be fine tuned

to elute specific

proteins or ligands.

Elution conditions may not be strong enough to elute unoptimized compounds.
Isocratic elution The elution buffer is identical to the application buffer, ligands move readily through the column but at different speeds according to their underlying affinity for the target. Transient interactions with
stationary phase
.
Allows for comparison of compound affinities. Can be time consuming; requires continuous fraction collection.
Biospecific elution Analytes are eluted by adding a high concentration of a high-affinity binding partner to the mobile phase, bound analytes are competed off of the stationary phase; the additive can be either a protein that scavenges bound ligands or a small molecule that displaces them.
Competitive binding
.
Allows for concentration of analytes with immediate elution; high recovery. Compounds not previously tested may have higher affinity than the elution additive.

Derivatization-free approaches

An example thermal proteome profiling workflow. Binding of a drug to a protein often leads to ligand-induced stabilization of the protein (1), which can be measured by comparing the amount of non-denatured protein remaining in a drug-treated sample to an untreated control. The change in protein stability can be visualized as a rightward shift in its stability curve (2). This image was adapted from BioRender.com.

While the approaches above have shown success, they are inherently limited by their need for

steric hindrance.[10] Immobilized ligands and targets are limited in their ability to move freely through space in a way that replicates the native protein-ligand interaction, and conformational change from induced fit is often limited when proteins or drugs are immobilized. Probe-based approaches also alter the three-dimensional nature of the ligand-protein interaction by introducing functional groups to the ligand, which can alter compound activity. Derivatization-free approaches aim to infer interactions by proxy, often through observations of changes to protein stability upon binding, and sometimes through chromatographic co-elution.[15]

The stability-based methods below are thought to work due to

and chemical degradation. Some examples of stability-based derivatization-free approaches follow.

Thermal proteome profiling (TPP)

Thermal proteome profiling (also,

An example drug affinity responsive target stability (DARTS) workflow. Binding of a drug to a protein often leads to ligand-induced stabilization of the protein. In DARTS, drug and control treated proteins are subjected to limited proteolysis and the extent of protein digestion can either be visualized on a gel or measured by mass spectrometry. Drug binding is expected to result in an increase in signal of the stabilized protein. This image was made using BioRender.com.

Drug affinity responsive target stability (DARTS)

The Drug Affinity Responsive Target Stability

cell lysate is incubated with a small molecule of interest, the sample is split into aliquots, and each aliquot goes through limited proteolysis after addition of protease. Limited proteolysis is critical, since complete proteolysis would render even a ligand-bound protein completely digested. Samples are then analyzed via SDS-PAGE to assess differences in extent of digestion, and bands are then excised and analyzed via mass spectrometry to confirm the identities of proteins that are resist proteolysis. Alternatively, if the target is already suspected and is being tested for validation, a western blot protocol can be used to identify protein directly.[18]

folded form of a protein. Extent of oxidation can be monitored by mass spectrometry and used to generate stability curves. This image was made using BioRender.com
.

Stability of proteins from rates of oxidation (SPROX)

Stability of Proteins from Rates of Oxidation also rests upon the assumption that

LC-MS/MS analysis of these peptides is challenging, as the contribution of other sample components to mass spectrometer noise can drown out relevant signal. Therefore, SPROX samples require fractionation to concentrate peptides of interest prior to LC-MS/MS analysis.[16]

Affinity selection-mass spectrometry

While adoption of affinity selection-mass spectrometry (AS-MS) has led to an expansion of assay formats,

compounds per experiment have been reported in the literature, and one group has reported assaying chemical libraries against heterogeneous protein pools.[20]
The basic steps of AS-MS are described in more detail below.

gel filtration column. Protein-bound compounds move around the beads and exit the column quickly. Unbound compounds are small enough to travel through beads and take a longer path before elution. This image was made using BioRender.com
.

Affinity selection

A generalized AS-MS workflow begins with the pre-incubation of

millimolar solutions, effectively capping the number of compounds screened in the thousands.[20] After appropriate test compounds and targets are selected and incubated, ligand-protein complexes can be separated by a variety of means.[19]

Separation of unbound small molecules and ligand-protein complexes

Affinity selection is followed by the removal of unbound

throughput, including pressure-based, centrifugal, and precipitation-based ultrafiltration.[19] Under both pressure-based and centrifugal formats, unbound small molecules are forced through a semipermeable membrane that excludes proteins on the basis of size. Multiple washing steps are required after ultrafiltration to ensure complete removal of unbound small molecules.[19] Ultrafiltration can also be confounded by non-specific adsorption of unbound small molecules to the membrane.[19] A group at the University of Illinois published a screening strategy involving amyloid-beta, in which ligands were used to stabilize the protein and prevent its aggregation. Ultrafiltration was used to precipitate aggregated amyloid-beta and remove unbound ligands, while the ligand-stabilized protein was detected and quantified using mass spectrometry.[19]

gel filtration chromatography, which allows for simultaneous removal of unbound ligands from up to 96 samples.[21] Samples are also passed through porous beads, but centrifugation is used to move the sample through the column.[21] SpeedScreen is not coupled to an LC-MS system and requires further processing prior to final analysis. In this case, ligands
must be freed from their targets and analyzed separately.

Analysis of bound ligands

The final step requires bioanalytical separation of bound ligands from their targets, and subsequent identification of ligands using liquid chromatography-mass spectrometry.[20] AS-MS offers means for identifying small molecule-protein interactions either directly - through top-down proteomic detection of intact complexes - or indirectly - through denaturation of small molecule-protein complexes followed by identification of small molecules using mass spectrometry.[19] The top-down approach requires direct infusion of the complex into an electrospray ionization mass spectrometry source under conditions gentle enough to preserve the interaction and maintain its integrity in the transition from liquid to gas.[19] While this was shown to be possible by Ganem and Henion in 1991, it is not suitable for high throughput.[19] Interestingly, electron capture dissociation, which is typically used in structure elucidation of peptides, has been used to identify ligand binding sites during top-down analysis.[19] A simpler method for analysis of bound ligands uses a protein precipitation extraction to denature proteins and release ligands into the precipitation solution, which can then be diluted and identified on an LC-MS system.

LC-MS/MS. Drug and protein elution profiles are constructed and correlated. Target identification is supported by a strong correlation in elution profile between a drug and a protein. This image was made using BioRender.com
.

Target identification by chromatographic co-elution (TICC)

Target identification by

ion-exchange chromatography system and fractionated. Lysate proteins are eluted along an ionic strength gradient and fractions are collected over short time intervals. Each fraction is analyzed by LC-MS/MS for both protein and drug content. Individual proteins elute with characteristic profiles, and pre-incubated drugs should mirror the elution profiles of the targets they complex with. Correlation of drug and protein elution profiles allows for targets to be narrowed down and inferred.[22]

Computational approaches

Molecular docking simulations

The development and application of bench-top chemoproteomics assays is often time consuming and cost-prohibitive.

Molecular docking strategies are categorized by the type of information that is already known about the ligand and protein of interest.[23]

Ligand-based methods

An example pharmacophore model. Each sphere represents a different scorable feature.

When a

least-squares regression for superimposition, but this requires user-selected anchor points and therefore introduces human bias into the process. Pharmacophore models require training data sets, giving rise to another challenge—selection of the appropriate library of compounds to adequately train models. Data set size and chemical diversity significantly affect performance of the downstream product.[25]

Structure-based methods

Ideally, the structure of a drug target is known, which allows for

quantitative structure-activity relationship (QSAR) profiling. Accurate QSAR models rely on inclusion of many potential targets, not just the therapeutic target. For example, important pharmacophores may yield high-affinity interactions with therapeutic targets, but they may also lead to undesirable off-target activity, and they may also be substrates of metabolic enzymes, such as Cytochrome P450s. Therefore, pharmacophore modeling against therapeutic targets is only one component of the compound's total structure-activity relationship.[24]

Applications

Druggability

ubiquitination
of the target by the E3 ubiquitin ligase, leading to target degradation.

Chemoproteomic strategies have been used to expand the scope of

E3 ubiquitin ligase. The interaction brings the E3 ubiquitin ligase close enough to the target that the target is labeled for degradation. The existence of potential covalent binding sites across the proteome suggests that many drugs can be covalently targeted using such a modality.[26]

Drug repurposing

Chemoproteomics is at the forefront of

drug repurposing. This is particularly relevant in the era of COVID-19, which saw a dire need to rapidly identify FDA approved drugs that have antiviral activity.[27] In this context, a phenotypic screen is usually employed to identify drugs with a desired effect in vitro, such as inhibition of viral plaque formation. If a drug produces a positive test, the next step is to determine whether it is acting on a known or novel target. Chemoproteomics is thus a follow-up to phenotypic screening. In the case of COVID-19, Friman et al investigated off-target effects of the broad-spectrum antiviral Remdesivir, which was among the first repurposed drugs to be used in the pandemic.[27] Remdesivir was tested via thermal proteome profiling in a HepG2 cellular thermal shift assay, along with the controversial drug hydroxychloroquine, and investigators discovered TRIP13 as a potential off-target of Remdesivir.[27]

High-throughput screening

Approved drugs are never identified as hits in

proteins.[20] This is due in large part to the sheer volume of ligands that can be screened in a single assay. Researchers at the iHuman Institute at ShanghaiTech University employed of scheme in which 20,000 compounds per pool were screened against A2AR, a difficult G-protein coupled receptor to drug, with a 0.12% hit rate, leading to several high affinity ligands.[20]

Real World Applications of Chemoproteomic Techniques
Technique Bioactive Ligand Target Description Reference
Activity based protein profiling ABX1431 MAGL Cesar et al optimized MAGL inhibitor ABX1431 using a competitive inhibition system. [28]
Photoaffinity labeling Chloroquine PfCRT Lekostaj et al identified the chloroquine binding site on the Plasmodium chloroquine resistance transporter using photoaffinity labeling. [7]
Affinity chromatography Purvanolol CDKs, CK1 Knockaert et al confirmed CDKs and discovered CK1 as targets of purvanolol; CK1 is a Plasmodium protein, and purvanolol analogs have activity against Plasmodium. [29]
Thermal proteome profiling (
CETSA
)
Remdesivir TRIP13 Friman et al identified TRIP13 as an off-target of Remdesivir after an investigation into off-target effects of repurposed COVID-19 drugs. [27]
Drug affinity responsive target stability Resveratrol EIF4A Lomenick et al discovered EIF4A as a target of the longevity-associated small molecule resveratrol. [18]
Stability of proteins from rates of oxidation Manassantin A EEF1A1 Geer Wallace et al discovered
antineoplastic natural product
Manassantin A
[15]
Affinity selection-mass spectrometry Screening compounds STING Merck & Co. used AS-MS to successfully screen for ligands against the difficult to drug STING protein, which has a large binding site and strong competing endogenous ligand [20]
Target identification by chromatographic co-elution 4513-0042 ERG6P Chan et al identified the yeast protein ERG6P as a target of the antifungal natural product 4513-0042 and followed up with validation experiments [22]

See also

References

  1. ^ Wang, Lei (2020-08-21). "Chemical Tools and Mass Spectrometry-based Approaches for Exploring Reactivity and Selectivity of Small Molecules in Complex Proteomes". Doctoral Dissertations.
  2. ^
    S2CID 231910437
    .
  3. ^ .
  4. .
  5. ^ .
  6. ^ .
  7. ^ .
  8. ^ .
  9. ^ .
  10. ^ .
  11. .
  12. ^ .
  13. .
  14. ^ .
  15. ^ .
  16. ^ .
  17. .
  18. ^ .
  19. ^ .
  20. ^ .
  21. ^ .
  22. ^ .
  23. .
  24. ^ .
  25. .
  26. ^ .
  27. ^ .
  28. .
  29. .