Bioconjugation

Last updated

Bioconjugation is a chemical strategy to form a stable covalent link between two molecules, at least one of which is a biomolecule.

Contents

Overview

Function

Recent advances in the understanding of biomolecules enabled their application to numerous fields like medicine, diagnostics, biocatalysis and materials. Synthetically modified biomolecules can have diverse functionalities, such as tracking cellular events, revealing enzyme function, determining protein biodistribution, imaging specific biomarkers, and delivering drugs to targeted cells. [1] [2] [3] [4] Bioconjugation is a crucial strategy that links these modified biomolecules with different substrates. Besides applications in biomedical research, bioconjugation has recently also gained importance in nanotechnology such as bioconjugated quantum dots.

Types of Conjugated Molecules

The most common types of bioconjugation include coupling of a small molecule (such as biotin or a fluorescent dye) to a protein. Antibody-drug conjugates such as Brentuximab vedotin and Gemtuzumab ozogamicin are examples falling into this category. [5]

Protein-protein conjugations, such as the coupling of an antibody to an enzyme, or the linkage of protein complexes, is also facilitated via bioconjugations. [6] [7]

Other less common molecules used in bioconjugation are oligosaccharides, nucleic acids, synthetic polymers such as polyethylene glycol, [8] and carbon nanotubes. [9]

Common Bioconjugation Reactions

Synthesis of bioconjugates involves a variety of challenges, ranging from the simple and nonspecific use of a fluorescent dye marker to the complex design of antibody drug conjugates. [1] [3] Various bioconjugation reactions have been developed to chemically modify proteins. Common types of bioconjugation reactions on proteins are coupling of lysine, cysteine, and tyrosine amino acid residues, as well as modification of tryptophan residues and of the N- and C- terminus. [1] [3] [4]

However, these reactions often lack chemoselectivity and efficiency, because they depend on the presence of native amino acids, which are present in large quantities that hinder selectivity. There is an increasing need for chemical strategies that can effectively attach synthetic molecules site specifically to proteins. One strategy is to first install a unique functional group onto a protein, and then a bioorthogonal reaction is used to couple a biomolecule with this unique functional group. [1] The bioorthogonal reactions targeting non-native functional groups are widely used in bioconjugation chemistry. Some important reactions are modification of ketone and aldehydes, Staudinger ligation with organic azides, copper-catalyzed Huisgen cycloaddition of azides, and strain promoted Huisgen cycloaddition of azides. [10] [11] [12] [13]

On Natural Amino Acids

Reactions of lysines

The nucleophilic lysine residue is commonly targeted site in protein bioconjugation, typically through amine-reactive N-hydroxysuccinimidyl (NHS) esters. [3] To obtain optimal number of deprotonated lysine residues, the pH of the aqueous solution must be below the pKa of the lysine ammonium group, which is around 10.5, so the typical pH of the reaction is about 8 and 9. The common reagent for the coupling reaction is NHS-ester (shown in the first reaction below in Figure 1), which reacts with nucleophilic lysine through a lysine acylation mechanism. Other similar reagents are isocyanates and isothiocyanates that undergo a similar mechanism (shown in the second and third reactions in Figure 1 below). [1] Benzoyl fluorides (shown in the last reaction below in Figure 1), which allows for lysine modification of proteins under mild conditions (low temperature, physiological pH), were recently proposed as an alternative to classically used lysine specific reagents. [14]

Bioconjugation strategies for lysine residues.tif

Reactions of cysteines

Because free cysteine rarely occurs on protein surface, it is an excellent choice for chemoselective modification. [15] Under basic condition, the cysteine residues will be deprotonated to generate a thiolate nucleophile, which will react with soft electrophiles, such as maleimides and iodoacetamides (shown in the first two reactions in Figure 2 below). As a result, a carbon-sulfur bond is formed. Another modification of cysteine residues involves the formation of disulfide bond (shown in the third reaction in Figure 2). The reduced cysteine residues react with exogenous disulfides, generating new a disulfide bond on the protein. An excess of disulfides is often used to drive the reaction, such as 2-thiopyridone and 3-carboxy-4-nitrothiophenol. [1] [3] Electron-deficient alkynes were demonstrated to selectively react with cysteine residues of proteins in the presence of other nucleophilic amino acid residues. Depending on the alkyne substitution, these reactions can produce either cleavable (when alkynone derivatives are used), [16] or hydrolytically stable bioconjugates (when 3-arylpropiolonitriles are used; the last reaction below in Figure 2). [17]

Bioconjugation strategies for cysteine residues.tif

Reactions of tyrosines

Tyrosine residues are relatively unreactive; therefore they have not been a popular targets for bioconjugation. Recent development has shown that the tyrosine can be modified through electrophilic aromatic substitutions (EAS) reactions, and it is selective for the aromatic carbon adjacent to the phenolic hydroxyl group. [1] This becomes particularly useful in the case that cysteine residues cannot be targeted. Specifically, diazonium effectively couples with tyrosine residues (diazonium salt shown as reagent in the first reaction in Figure 3 below), and an electron withdrawing substituent in the 4-position of diazonium salt can effectively increase the efficiency of the reaction. Cyclic diazodicarboxyamide derivative like 4-Phenyl-1,2,4-triazole-3,5-dione (PTAD) were reported for selective bioconjugation on tyrosine residues (the second reaction in Figure 3 below). [18] A three-component Mannich-type reaction with aldehydes and anilines (the last reaction in Figure 3) was also described to be relatively tyrosine-selective under mild optimised reaction conditions. [19]

Bioconjugation strategies for tyrosine residues.tif

Reactions of N- and C- termini

Since natural amino acid residues are usually present in large quantities, it is often difficult to modify one single site. Strategies targeting the termini of protein have been developed, because they greatly enhanced the site selectivity of protein modification. One of the N- termini modifications involves the functionalization of the terminal amino acid. The oxidation of N-terminal serine and threonine residues are able to generate N-terminal aldehyde, which can undergo further bioorthogonal reactions (shown in the first reaction in Figure 4). Another type of modification involves the condensation of N-terminal cysteine with aldehyde, generating thiazolidine that is stable at high pH (second reaction in Figure 4). Using pyridoxal phosphate (PLP), several N-terminal amino acids can undergo transamination to yield N-terminal aldehyde, such as glycine and aspartic acid (third reaction in Figure 4).

Figure 4. Bioconjugation strategies for N-terminus.tif

An example of C-termini modification is the native chemical ligation (NCL), which is the coupling between a C-terminal thioester and a N-terminal cysteine (Figure 5).

Figure 5. Bioconjugation strategies for C-terminus.tif

Bioorthogonal Reactions: On Unique Functional Groups

Modification of ketones and aldehydes

A ketone or aldehyde can be attached to a protein through the oxidation of N-terminal serine residues or transamination with PLP. Additionally, they can be introduced by incorporating unnatural amino acids via the Tirrell method or Schultz method. [10] They will then selectively condense with an alkoxyamine and a hydrazine, producing oxime and hydrazone derivatives (shown in the first and second reactions, respectively, in Figure 6). This reaction is highly chemoselective in terms of protein bioconjugation, but the reaction rate is slow. The mechanistic studies show that the rate determining step is the dehydration of tetrahedral intermediate, so a mild acidic solution is often employed to accelerate the dehydration step. [2]

Figure 6. Bioconjugation strategies for targeting ketones and aldehydes.jpg

The introduction of nucleophilic catalyst can significantly enhance reaction rate (shown in Figure 7). For example, using aniline as a nucleophilic catalyst, a less populated protonated carbonyl becomes a highly populated protonated Schiff base. [20] In other words, it generates a high concentration of reactive electrophile. The oxime ligation can then occur readily, and it has been reported that the rate increased up to 400 times under mild acidic condition. [20] The key of this catalyst is that it can generate a reactive electrophile without competing with desired product.

Figure 7. Nucleophilic catalysis of oxime ligation.jpg

Recent developments that exploit proximal functional groups have enabled hydrazone condensations [21] to operate at 20 M−1s−1 at neutral pH while oxime condensations have been discovered which proceed at 500-10000 M−1s−1 at neutral pH without added catalysts. [22] [23]

Staudinger ligation with azides

The Staudinger ligation of azides and phosphine has been used extensively in field of chemical biology. Because it is able to form a stable amide bond in living cells and animals, it has been applied to modification of cell membrane, in vivo imaging, and other bioconjugation studies. [24] [25] [26] [27]


Figure 8. Staudinger Ligation with Azides.jpg

Contrasting with the classic Staudinger reaction, Staudinger ligation is a second order reaction in which the rate-limiting step is the formation of phosphazide (specific reaction mechanism shown in Figure 9). The triphenylphosphine first reacts with the azide to yield an azaylide through a four-membered ring transition state, and then an intramolecular reaction leads to the iminophosphorane intermediate, which will then give the amide-linkage under hydrolysis. [28]

Figure 9. Mechanism of Staudinger Ligation.jpg

Huisgen cyclization of azides

Copper catalyzed Huisgen cyclization of azides

Azide has become a popular target for chemoselective protein modification, because they are small in size and have a favorable thermodynamic reaction potential. One such azide reactions is the [3+2] cycloaddition reaction with alkyne, but the reaction requires high temperature and often gives mixtures of regioisomers.

Figure 10. Copper-catalyzed cyclization of Azides.jpg

An improved reaction developed by chemist Karl Barry Sharpless involves the copper (I) catalyst, which couples azide with terminal alkyne that only give 1,4 substituted 1,2,3 triazoles in high yields (shown below in Figure 11). The mechanistic study suggests a stepwise reaction. [13] The Cu (I) first couples with acetylenes, and then it reacts with azide to generate a six-membered intermediate. The process is very robust that it occurs at pH ranging from 4 to 12, and copper (II) sulfate is often used as a catalyst in the presence of a reducing agent. [13]

Figure 11. Mechanism for Copper-catalyzed cyclization of Azides.jpg

Strain promoted Huisgen cyclization of azides

Even though Staudinger ligation is a suitable bioconjugation in living cells without major toxicity, the phosphine's sensitivity to air oxidation and its poor solubility in water significantly hinder its efficiency. The copper(I) catalyzed azide-alkyne coupling has reasonable reaction rate and efficiency under physiological conditions, but copper poses significant toxicity and sometimes interferes with protein functions in living cells. In 2004, chemist Carolyn R. Bertozzi's lab developed a metal free [3+2] cycloaddition using strained cyclooctyne and azide. Cyclooctyne, which is the smallest stable cycloalkyne, can couple with azide through [3+2] cycloaddition, leading to two regioisomeric triazoles (Figure 12). [11] The reaction occurs readily at room temperature and therefore can be used to effectively modify living cells without negative effects. It has also been reported that the installation of fluorine substituents on a cyclic alkyne can greatly accelerate the reaction rate. [2] [29]

Figure 12. Strain promoted cycloaddition of azides and cyclooctynes.jpg

Transition Metal-Mediated Bioconjugation Reactions

Transition metal-based bioconjugation had been challenging due to the nature of biological conditions – aqueous solution, room temperature, mild pH, and low substrate concentrations – which are generally challenging for organometallic reactions. However, recently, besides copper-catalyzed [3 + 2] azide alkyne cycloaddition reaction, more and more diverse transition metal-mediated chemical transformations have been applied for bioconjugation reactions, introducing olefin metathesis, alkylation, C–H arylation, C–C, C–S, and C–N cross-coupling reactions. [30] [31]

Alkylation

On Natural Amino Acids

  • Rh-catalyzed Trp and Cys alkylation [32] [33]

Using in situ generated RhII-carbenoid by activation of vinyl-substituted diazo compounds with Rh2(OAc)4, tryptophans and cysteines were shown to be selectively alkylated in aqueous media.

However, this method is limited to surface tryptophans and cysteines possibly because of steric constraints. [34]

Rh-carbenoid 2.png

  • Ir-catalyzed Lys and N-terminus (reductive) alkylation [35]

Imines formed from the condensation of aldehydes with lysines or the N-terminus can be reduced efficient by an water-stable [Cp*Ir(bipy)(H2O)]SO4 complex in the presence of formate ions (serving as the hydride source). The reaction happens readily under physiologically relevant conditions and results in high conversion for various aromatic aldehydes.

Ir-catalyzed Lys (reductive) alkylation.png

  • Pd-catalyzed Tyr O-alkylation [36]

By using a pre-formed electrophilic π-allylpalladium(II) reagent derived from allylic acetate or carbamate precursors, selective allylic alkylation of tyrosines can be achieved in aqueous solution at room temperature and in the presence of cysteines.

Pd-catalyzed Tyr O-alkylation.png

  • Au-catalyzed Cys alkylation [37]

Cysteine-containing peptides have been shown to undergo 1,2-addition to allenes in the presence of gold(I) and/or silver(I) salts, producing hydroxyl substituted vinyl thioethers. The reaction with peptides proceeds with high yields and is selective for cysteines over other nucleophilic residues.

However, the reactivity towards proteins is much decreased, potentially due to the coordination of gold to the protein backbone.

Au-catalyzed Cys alkylation.png

Arylation

On Natural Amino Acids

  • Trp arylation

Multiple methods have been reported to achieve tryptophan C–H arylation, where diverse electrophiles such as aryl halides [38] [39] and aryl boronic acids [40] (an example shown below) have been used to transfer the aryl groups.

However, current tryptophan C–H arylation reaction conditions remain relatively harsh, requiring organic solvents, low pH and/or high temperatures.

Trp arylation.png

  • Cys arylation

Free thiols has been considered unfavorable for Pd-mediated reactions due to Pd-catalyst decomposition. [41] However, PdII oxidative addition complexes (OACs) supported by dialkylbiaryl phosphine ligands have shown to work efficiently towards cysteine S-arylation.

The first example is the use of PdII OAC with RuPhos: [42] The PdII complex resulting from the oxidative addition of aryl halides or trifluoromethanesulfonates and using RuPhos as the ligand could chemoselectively modify cysteines in various buffer with 5% organic co-solvent under neutral pH. This method has been shown to modify peptides and proteins, achieve peptide macrocyclization (by using bis-palladium reagent and peptides with two unprotected cysteines) [43] and synthesizing antibody-drug conjugates (ADCs). Changing the ligand to sSPhos supports the PdII complex to be sufficiently water soluble to achieve cysteine S-arylation under cosolvent-free aqueous conditions. [44]

Cys arylation (RuPhos) -2.png

There are other applications of this method where the PdII complexes were generated as PdII-peptide OACs by introducing 4-halophenylalanine into peptides during SPPS to achieve peptide-peptide or peptide-protein ligation. [45]

Pd-peptide OAC 2.png

Alternate to directly oxidative addition to the peptide, the Pd OACs could also be transferred to the protein through amine-selective acylation reaction via NHS ester. The latter has been applied to selectively label surface lysine residues of a protein (forming PdII-protein OACs) and oligonucleotides (forming PdII-oligonucleotide OACs), which could then be linked to cysteine-containing peptides or proteins. [46]

Pd-biomolecule OAC 2.png

Another example of protein-protein cross-coupling is achieved through converting cysteine residues into an electrophilic S-aryl–Pd–X OAC by utilizing an intramolecular oxidative addition strategy. [47]

Similar to cysteine, lysine N-arylation could be achieved through Pd OACs with different dialkylbiaryl phosphine ligands. Due to weaker nucleophilicity and slower reductive elimination rate compared to cysteine, the selection of supporting ligands is shown to be critical. The bulky BrettPhos and t-BuBrettPhos ligands in conjunction with mildly basic sodium phenoxide have been used as the strategy to functionalize lysines on peptide substrates. The reaction happens in mild conditions and is selective over most other nucleophilic amino acid residues.

Lys arylation.png

On Unnatural Amino Acids

Pd-mediated Sonogashira, Heck, and Suzuki-Miyaura cross-coupling reactions have been applied widely to modify peptides and proteins, where diverse Pd reagents have been developed for the application in aqueous solutions. [49] Those reactions require the protein or peptide substrate bearing unnatural functional groups such as alkyne, [50] [51] [52] aryl halides, [53] [54] [55] [56] and aryl boronic acids, [57] which can be achieved through genetic code expansion or post-translational modifications.

UAA arylation.png

Examples of Applied Bioconjugation Techniques

Growth Factors

Bioconjugation of TGF-β to iron oxide nanoparticles and its activation through magnetic hyperthermia in-vitro has been reported. [58] This was done by using 1-(3-dimethylaminopropyl)ethylcarbodiimide combined with N-Hydroxysuccinimide to form primary amide bonds with the free primary amines on the growth factor. Carbon nanotubes have been successfully used in conjunction with bioconjugation to link TGF-β followed by an activation with near-infrared light. [59] Typically, these reactions have involved the use of a crosslinker, but some of these add molecular space between the compound of interest and base material and in turn causes higher degrees of non-specific binding and unwanted reactivity. [60]

See also

Related Research Articles

<span class="mw-page-title-main">Protein primary structure</span> Linear sequence of amino acids in a peptide or protein

Protein primary structure is the linear sequence of amino acids in a peptide or protein. By convention, the primary structure of a protein is reported starting from the amino-terminal (N) end to the carboxyl-terminal (C) end. Protein biosynthesis is most commonly performed by ribosomes in cells. Peptides can also be synthesized in the laboratory. Protein primary structures can be directly sequenced, or inferred from DNA sequences.

<span class="mw-page-title-main">Post-translational modification</span> Biological processes

Post-translational modification (PTM) is the covalent process of changing proteins following protein biosynthesis. PTMs may involve enzymes or occur spontaneously. Proteins are created by ribosomes translating mRNA into polypeptide chains, which may then change to form the mature protein product. PTMs are important components in cell signalling, as for example when prohormones are converted to hormones.

Native Chemical Ligation (NCL) is an important extension of the chemical ligation concept for constructing a larger polypeptide chain by the covalent condensation of two or more unprotected peptides segments. Native chemical ligation is the most effective method for synthesizing native or modified proteins of typical size.

<span class="mw-page-title-main">Protein sequencing</span> Sequencing of amino acid arrangement in a protein

Protein sequencing is the practical process of determining the amino acid sequence of all or part of a protein or peptide. This may serve to identify the protein or characterize its post-translational modifications. Typically, partial sequencing of a protein provides sufficient information to identify it with reference to databases of protein sequences derived from the conceptual translation of genes.

The Heck reaction is the chemical reaction of an unsaturated halide with an alkene in the presence of a base and a palladium catalyst to form a substituted alkene. It is named after Tsutomu Mizoroki and Richard F. Heck. Heck was awarded the 2010 Nobel Prize in Chemistry, which he shared with Ei-ichi Negishi and Akira Suzuki, for the discovery and development of this reaction. This reaction was the first example of a carbon-carbon bond-forming reaction that followed a Pd(0)/Pd(II) catalytic cycle, the same catalytic cycle that is seen in other Pd(0)-catalyzed cross-coupling reactions. The Heck reaction is a way to substitute alkenes.

<span class="mw-page-title-main">Peptide synthesis</span> Production of peptides

In organic chemistry, peptide synthesis is the production of peptides, compounds where multiple amino acids are linked via amide bonds, also known as peptide bonds. Peptides are chemically synthesized by the condensation reaction of the carboxyl group of one amino acid to the amino group of another. Protecting group strategies are usually necessary to prevent undesirable side reactions with the various amino acid side chains. Chemical peptide synthesis most commonly starts at the carboxyl end of the peptide (C-terminus), and proceeds toward the amino-terminus (N-terminus). Protein biosynthesis in living organisms occurs in the opposite direction.

<span class="mw-page-title-main">Chemical biology</span> Scientific discipline

Chemical biology is a scientific discipline between the fields of chemistry and biology. The discipline involves the application of chemical techniques, analysis, and often small molecules produced through synthetic chemistry, to the study and manipulation of biological systems. Although often confused with biochemistry, which studies the chemistry of biomolecules and regulation of biochemical pathways within and between cells, chemical biology remains distinct by focusing on the application of chemical tools to address biological questions.

In chemical synthesis, click chemistry is a class of simple, atom-economy reactions commonly used for joining two molecular entities of choice. Click chemistry is not a single specific reaction, but describes a way of generating products that follow examples in nature, which also generates substances by joining small modular units. In many applications, click reactions join a biomolecule and a reporter molecule. Click chemistry is not limited to biological conditions: the concept of a "click" reaction has been used in chemoproteomic, pharmacological, biomimetic and molecular machinery applications. However, they have been made notably useful in the detection, localization and qualification of biomolecules.

The Hiyama coupling is a palladium-catalyzed cross-coupling reaction of organosilanes with organic halides used in organic chemistry to form carbon–carbon bonds. This reaction was discovered in 1988 by Tamejiro Hiyama and Yasuo Hatanaka as a method to form carbon-carbon bonds synthetically with chemo- and regioselectivity. The Hiyama coupling has been applied to the synthesis of various natural products.

The Lowry protein assay is a biochemical assay for determining the total level of protein in a solution. The total protein concentration is exhibited by a color change of the sample solution in proportion to protein concentration, which can then be measured using colorimetric techniques. It is named for the biochemist Oliver H. Lowry who developed the reagent in the 1940s. His 1951 paper describing the technique is the most-highly cited paper ever in the scientific literature, cited over 300,000 times.

The Negishi coupling is a widely employed transition metal catalyzed cross-coupling reaction. The reaction couples organic halides or triflates with organozinc compounds, forming carbon-carbon bonds (C-C) in the process. A palladium (0) species is generally utilized as the metal catalyst, though nickel is sometimes used. A variety of nickel catalysts in either Ni0 or NiII oxidation state can be employed in Negishi cross couplings such as Ni(PPh3)4, Ni(acac)2, Ni(COD)2 etc.

Glycopeptides are peptides that contain carbohydrate moieties (glycans) covalently attached to the side chains of the amino acid residues that constitute the peptide.

<span class="mw-page-title-main">PEPPSI</span> Group of chemical compounds

PEPPSI is an abbreviation for pyridine-enhanced precatalyst preparation stabilization and initiation. It refers to a family of commercially available palladium catalysts developed around 2005 by Prof. Michael G. Organ and co-workers at York University, which can accelerate various carbon-carbon and carbon-heteroatom bond forming cross-coupling reactions. In comparison to many alternative palladium catalysts, Pd-PEPPSI-type complexes are stable to air and moisture and are relatively easy to synthesize and handle.

The term bioorthogonal chemistry refers to any chemical reaction that can occur inside of living systems without interfering with native biochemical processes. The term was coined by Carolyn R. Bertozzi in 2003. Since its introduction, the concept of the bioorthogonal reaction has enabled the study of biomolecules such as glycans, proteins, and lipids in real time in living systems without cellular toxicity. A number of chemical ligation strategies have been developed that fulfill the requirements of bioorthogonality, including the 1,3-dipolar cycloaddition between azides and cyclooctynes, between nitrones and cyclooctynes, oxime/hydrazone formation from aldehydes and ketones, the tetrazine ligation, the isocyanide-based click reaction, and most recently, the quadricyclane ligation.

An aldehyde tag is a short peptide tag that can be further modified to add fluorophores, glycans, PEG chains, or reactive groups for further synthesis. A short, genetically-encoded peptide with a consensus sequence LCxPxR is introduced into fusion proteins, and by subsequent treatment with the formylglycine-generating enzyme (FGE), the cysteine of the tag is converted to a reactive aldehyde group. This electrophilic group can be targeted by an array of aldehyde-specific reagents, such as aminooxy- or hydrazide-functionalized compounds.

<span class="mw-page-title-main">Dehydroamino acid</span>

In biochemistry, a dehydroamino acid or α,β-dehydroamino acid is an amino acids, usually with a C=C double bond in its side chain. Dehydroamino acids are not coded by DNA, but arise via post-translational modification.

Dialkylbiaryl phosphine ligands are phosphine ligands that are used in homogeneous catalysis. They have proved useful in Buchwald-Hartwig amination and etherification reactions as well as Negishi cross-coupling, Suzuki-Miyaura cross-coupling, and related reactions. In addition to these Pd-based processes, their use has also been extended to transformations catalyzed by nickel, gold, silver, copper, rhodium, and ruthenium, among other transition metals.

<span class="mw-page-title-main">3-Arylpropiolonitriles</span> Chemical compound

3-Arylpropiolonitriles (APN) belong to a class of electron-deficient alkyne derivatives substituted by two electron-withdrawing groups – a nitrile and an aryl moieties. Such activation results in improved selectivity towards highly reactive thiol-containing molecules, namely cysteine residues in proteins. APN-based modification of proteins was reported to surpass several important drawbacks of existing strategies in bioconjugation, notably the presence of side reactions with other nucleophilic amino acid residues and the relative instability of the resulting bioconjugates in the blood stream. The latter drawback is especially important for the preparation of targeted therapies, such as antibody-drug conjugates.

Bradley Lether Pentelute is currently a professor of chemistry at the Massachusetts Institute of Technology (MIT). His research program lies at the intersection of chemistry and biology and develops bioconjugation strategies, cytosolic delivery platforms, and rapid flow synthesis technologies to optimize the production, achieve site-specific modification, enhance stability, and modulate function of a variety of bioactive agents. His laboratory successfully modified proteins via cysteine-containing “pi-clamps” made up of a short sequence of amino acids, and delivered large biomolecules, such as various proteins and drugs, into cells via the anthrax delivery vehicle. Pentelute has also made several key contributions to automated synthesis technologies in flow. These advances includes the invention of the world's fastest polypeptide synthesizer. This system is able to form amide bonds at a more efficient rate than standard commercial equipment and has helped in the process of understanding protein folding and its mechanisms. This automated flow technology was recently used to achieve total chemical synthesis of protein chains up to 164 amino acids in length that retained the structure and function of native variants obtained by recombinant expression. The primary goal of his endeavor is to use these processes to create designer biologics that can be used to treat diseases and solve the manufacturing problem for on-demand personalized therapies, such as cancer vaccines.

An artificial metalloenzyme (ArM) is a metalloprotein made in the laboratory which cannot be found in the nature and can catalyze certain desired chemical reactions. Despite fitting into classical enzyme categories, ArMs also have potential in chemical reactivity like catalyzing Suzuki coupling, metathesis and so on, which are never reported in natural enzymatic reaction. With the progress in organometallic synthesis and protein engineering, more and more new kind of design of ArMs came out, showing promising future in both academia and industrial aspects.

References

  1. 1 2 3 4 5 6 7 Stephanopoulos N, Francis MB (November 2011). "Choosing an effective protein bioconjugation strategy". Nature Chemical Biology. 7 (12): 876–884. doi:10.1038/nchembio.720. PMID   22086289.
  2. 1 2 3 Tilley SD, Joshi NS, Francis MB (2008). "Proteins: Chemistry and Chemical Reactivity". Wiley Encyclopedia of Chemical Biology. pp. 1–16. doi:10.1002/9780470048672.wecb493. ISBN   978-0470048672.
  3. 1 2 3 4 5 Francis MB, Carrico IS (December 2010). "New frontiers in protein bioconjugation". Current Opinion in Chemical Biology. 14 (6): 771–773. doi:10.1016/j.cbpa.2010.11.006. PMID   21112236.
  4. 1 2 Kalia J, Raines RT (January 2010). "Advances in Bioconjugation". Current Organic Chemistry. 14 (2): 138–147. doi:10.2174/138527210790069839. PMC   2901115 . PMID   20622973.
  5. Gerber HP, Senter PD, Grewal IS (2009). "Antibody drug-conjugates targeting the tumor vasculature: Current and future developments". mAbs. 1 (3): 247–253. doi:10.4161/mabs.1.3.8515. PMC   2726597 . PMID   20069754. Archived from the original on February 2, 2014.
  6. Koniev O, Wagner A (August 2015). "Developments and recent advancements in the field of endogenous amino acid selective bond forming reactions for bioconjugation". Chemical Society Reviews. 44 (15): 5495–5551. doi: 10.1039/C5CS00048C . PMID   26000775.
  7. Hutchins GH, Kiehstaller S, Poc P, Lewis AH, Oh J, Sadighi R, et al. (February 2024). "Covalent bicyclization of protein complexes yields durable quaternary structures". Chem. 10 (2): 615–627. doi:10.1016/j.chempr.2023.10.003. PMC   10857811 . PMID   38344167.
  8. Thordarson P, Le Droumaguet B, Velonia K (November 2006). "Well-defined protein-polymer conjugates--synthesis and potential applications". Applied Microbiology and Biotechnology. 73 (2): 243–254. doi:10.1007/s00253-006-0574-4. PMID   17061132. S2CID   23657616.
  9. Yang W, Thordarson P (2007). "Carbon nanotubes for biological and biomedical applications". Nanotechnology . 18 (41): 412001. Bibcode:2007Nanot..18O2001Y. doi:10.1088/0957-4484/18/41/412001. S2CID   137867074.
  10. 1 2 Carrico IS, Carlson BL, Bertozzi CR (June 2007). "Introducing genetically encoded aldehydes into proteins". Nature Chemical Biology. 3 (6): 321–322. doi:10.1038/nchembio878. PMID   17450134.
  11. 1 2 Agard NJ, Prescher JA, Bertozzi CR (November 2004). "A strain-promoted [3 + 2] azide-alkyne cycloaddition for covalent modification of biomolecules in living systems". Journal of the American Chemical Society. 126 (46): 15046–15047. doi:10.1021/ja044996f. PMID   15547999.
  12. Kolb HC, Finn MG, Sharpless KB (June 2001). "Click Chemistry: Diverse Chemical Function from a Few Good Reactions". Angewandte Chemie. 40 (11): 2004–2021. doi:10.1002/1521-3773(20010601)40:11<2004::AID-ANIE2004>3.0.CO;2-5. PMID   11433435.
  13. 1 2 3 Rostovtsev VV, Green LG, Fokin VV, Sharpless KB (July 2002). "A stepwise huisgen cycloaddition process: copper(I)-catalyzed regioselective "ligation" of azides and terminal alkynes". Angewandte Chemie. 41 (14): 2596–2599. doi:10.1002/1521-3773(20020715)41:14<2596::AID-ANIE2596>3.0.CO;2-4. PMID   12203546.
  14. Dovgan I, Ursuegui S, Erb S, Michel C, Kolodych S, Cianférani S, et al. (May 2017). "Acyl Fluorides: Fast, Efficient, and Versatile Lysine-Based Protein Conjugation via Plug-and-Play Strategy". Bioconjugate Chemistry. 28 (5): 1452–1457. doi:10.1021/acs.bioconjchem.7b00141. PMID   28443656.
  15. Fodje MN, Al-Karadaghi S (May 2002). "Occurrence, conformational features and amino acid propensities for the pi-helix". Protein Engineering. 15 (5): 353–358. doi: 10.1093/protein/15.5.353 . PMID   12034854.
  16. Shiu HY, Chan TC, Ho CM, Liu Y, Wong MK, Che CM (2009). "Electron-deficient alkynes as cleavable reagents for the modification of cysteine-containing peptides in aqueous medium". Chemistry: A European Journal. 15 (15): 3839–3850. doi:10.1002/chem.200800669. PMID   19229937.
  17. Koniev O, Leriche G, Nothisen M, Remy JS, Strub JM, Schaeffer-Reiss C, et al. (February 2014). "Selective irreversible chemical tagging of cysteine with 3-arylpropiolonitriles". Bioconjugate Chemistry. 25 (2): 202–206. doi:10.1021/bc400469d. PMID   24410136.
  18. Ban H, Nagano M, Gavrilyuk J, Hakamata W, Inokuma T, Barbas CF (April 2013). "Facile and stabile linkages through tyrosine: bioconjugation strategies with the tyrosine-click reaction". Bioconjugate Chemistry. 24 (4): 520–532. doi:10.1021/bc300665t. PMC   3658467 . PMID   23534985.
  19. Joshi NS, Whitaker LR, Francis MB (December 2004). "A three-component Mannich-type reaction for selective tyrosine bioconjugation". Journal of the American Chemical Society. 126 (49): 15942–15943. doi:10.1021/ja0439017. PMID   15584710.
  20. 1 2 Dirksen A, Hackeng TM, Dawson PE (November 2006). "Nucleophilic catalysis of oxime ligation". Angewandte Chemie. 45 (45): 7581–7584. doi: 10.1002/anie.200602877 . PMID   17051631.
  21. Kool ET, Park DH, Crisalli P (November 2013). "Fast hydrazone reactants: electronic and acid/base effects strongly influence rate at biological pH". Journal of the American Chemical Society. 135 (47): 17663–17666. doi:10.1021/ja407407h. PMC   3874453 . PMID   24224646.
  22. Schmidt P, Zhou L, Tishinov K, Zimmermann K, Gillingham D (October 2014). "Dialdehydes lead to exceptionally fast bioconjugations at neutral pH by virtue of a cyclic intermediate". Angewandte Chemie. 53 (41): 10928–10931. doi:10.1002/anie.201406132. PMID   25164607.
  23. Schmidt P, Stress C, Gillingham D (June 2015). "Boronic acids facilitate rapid oxime condensations at neutral pH". Chemical Science. 6 (6): 3329–3333. doi:10.1039/C5SC00921A. PMC   5656983 . PMID   29142692.
  24. Lemieux GA, De Graffenried CL, Bertozzi CR (April 2003). "A fluorogenic dye activated by the staudinger ligation". Journal of the American Chemical Society. 125 (16): 4708–4709. doi:10.1021/ja029013y. PMID   12696879.
  25. Laughlin ST, Baskin JM, Amacher SL, Bertozzi CR (May 2008). "In vivo imaging of membrane-associated glycans in developing zebrafish". Science. 320 (5876): 664–667. Bibcode:2008Sci...320..664L. doi:10.1126/science.1155106. PMC   2701225 . PMID   18451302.
  26. Saxon E, Bertozzi CR (March 2000). "Cell surface engineering by a modified Staudinger reaction". Science. 287 (5460): 2007–2010. Bibcode:2000Sci...287.2007S. doi:10.1126/science.287.5460.2007. PMID   10720325. S2CID   19720277.
  27. Prescher JA, Dube DH, Bertozzi CR (August 2004). "Chemical remodelling of cell surfaces in living animals". Nature. 430 (7002): 873–877. Bibcode:2004Natur.430..873P. doi:10.1038/nature02791. PMID   15318217. S2CID   4371934.
  28. Lin FL, Hoyt HM, van Halbeek H, Bergman RG, Bertozzi CR (March 2005). "Mechanistic investigation of the staudinger ligation". Journal of the American Chemical Society. 127 (8): 2686–2695. doi:10.1021/ja044461m. PMID   15725026.
  29. Chang PV, Prescher JA, Sletten EM, Baskin JM, Miller IA, Agard NJ, et al. (February 2010). "Copper-free click chemistry in living animals". Proceedings of the National Academy of Sciences of the United States of America. 107 (5): 1821–1826. doi: 10.1073/pnas.0911116107 . PMC   2836626 . PMID   20080615.
  30. Rodríguez J, Martínez-Calvo M (August 2020). "Transition-Metal-Mediated Modification of Biomolecules". Chemistry: A European Journal. 26 (44): 9792–9813. doi:10.1002/chem.202001287. PMID   32602145. S2CID   220272489.
  31. Vinogradova EV (2017-11-01). "Organometallic chemical biology: an organometallic approach to bioconjugation". Pure and Applied Chemistry. 89 (11): 1619–1640. doi: 10.1515/pac-2017-0207 . ISSN   1365-3075. S2CID   103469980.
  32. Antos JM, Francis MB (August 2004). "Selective tryptophan modification with rhodium carbenoids in aqueous solution". Journal of the American Chemical Society. 126 (33): 10256–10257. doi:10.1021/ja047272c. PMID   15315433.
  33. Kundu R, Ball ZT (May 2013). "Rhodium-catalyzed cysteine modification with diazo reagents". Chemical Communications. 49 (39): 4166–4168. doi:10.1039/c2cc37323h. PMID   23175246.
  34. Sauthoff G (1996-01-01). "Intermetallics J. Am. Chem. Soc. 1996, 118, 934". Journal of the American Chemical Society. 118 (43): 10678. doi:10.1021/ja965445v. ISSN   0002-7863.
  35. McFarland JM, Francis MB (October 2005). "Reductive alkylation of proteins using iridium catalyzed transfer hydrogenation". Journal of the American Chemical Society. 127 (39): 13490–13491. doi:10.1021/ja054686c. PMID   16190700.
  36. Tilley SD, Francis MB (February 2006). "Tyrosine-selective protein alkylation using pi-allylpalladium complexes". Journal of the American Chemical Society. 128 (4): 1080–1081. doi:10.1021/ja057106k. PMID   16433516.
  37. Chan AO, Tsai JL, Lo VK, Li GL, Wong MK, Che CM (February 2013). "Gold-mediated selective cysteine modification of peptides using allenes". Chemical Communications. 49 (14): 1428–1430. doi:10.1039/c2cc38214h. PMID   23322001.
  38. Ruiz-Rodríguez J, Albericio F, Lavilla R (January 2010). "Postsynthetic modification of peptides: chemoselective C-arylation of tryptophan residues". Chemistry: A European Journal. 16 (4): 1124–1127. doi:10.1002/chem.200902676. PMID   20013969.
  39. Schischko A, Ren H, Kaplaneris N, Ackermann L (February 2017). "Bioorthogonal Diversification of Peptides through Selective Ruthenium(II)-Catalyzed C-H Activation". Angewandte Chemie. 56 (6): 1576–1580. doi:10.1002/anie.201609631. PMID   28074503.
  40. Williams TJ, Reay AJ, Whitwood AC, Fairlamb IJ (March 2014). "A mild and selective Pd-mediated methodology for the synthesis of highly fluorescent 2-arylated tryptophans and tryptophan-containing peptides: a catalytic role for Pd(0) nanoparticles?". Chemical Communications. 50 (23): 3052–3054. doi: 10.1039/C3CC48481E . PMID   24516861.
  41. Bong DT, Ghadiri MR (August 2001). "Chemoselective Pd(0)-catalyzed peptide coupling in water". Organic Letters. 3 (16): 2509–2511. doi:10.1021/ol016169e. PMID   11483047.
  42. Vinogradova EV, Zhang C, Spokoyny AM, Pentelute BL, Buchwald SL (October 2015). "Organometallic palladium reagents for cysteine bioconjugation". Nature. 526 (7575): 687–691. doi:10.1038/nature15739. PMC   4809359 . PMID   26511579.
  43. Rojas AJ, Zhang C, Vinogradova EV, Buchwald NH, Reilly J, Pentelute BL, et al. (June 2017). "Divergent unprotected peptide macrocyclisation by palladium-mediated cysteine arylation". Chemical Science. 8 (6): 4257–4263. doi:10.1039/C6SC05454D. PMC   5635729 . PMID   29081961.
  44. Rojas AJ, Pentelute BL, Buchwald SL (August 2017). "Water-Soluble Palladium Reagents for Cysteine S-Arylation under Ambient Aqueous Conditions". Organic Letters. 19 (16): 4263–4266. doi:10.1021/acs.orglett.7b01911. PMC   5818991 . PMID   28777001.
  45. Rojas AJ, Wolfe JM, Dhanjee HH, Buslov I, Truex NL, Liu RY, et al. (October 2022). "Palladium-peptide oxidative addition complexes for bioconjugation". Chemical Science. 13 (40): 11891–11895. doi:10.1039/D2SC04074C. PMC   9580489 . PMID   36320916.
  46. Jbara M, Rodriguez J, Dhanjee HH, Loas A, Buchwald SL, Pentelute BL (May 2021). "Oligonucleotide Bioconjugation with Bifunctional Palladium Reagents". Angewandte Chemie. 60 (21): 12109–12115. doi:10.1002/anie.202103180. PMC   8143041 . PMID   33730425.
  47. Dhanjee HH, Saebi A, Buslov I, Loftis AR, Buchwald SL, Pentelute BL (May 2020). "Protein-Protein Cross-Coupling via Palladium-Protein Oxidative Addition Complexes from Cysteine Residues". Journal of the American Chemical Society. 142 (20): 9124–9129. doi:10.1021/jacs.0c03143. PMC   7586714 . PMID   32364380.
  48. Lee HG, Lautrette G, Pentelute BL, Buchwald SL (March 2017). "Palladium-Mediated Arylation of Lysine in Unprotected Peptides". Angewandte Chemie. 56 (12): 3177–3181. doi:10.1002/anie.201611202. PMC   5741856 . PMID   28206688.
  49. Li J, Chen PR (August 2012). "Moving Pd-mediated protein cross coupling to living systems". ChemBioChem. 13 (12): 1728–1731. doi:10.1002/cbic.201200353. PMID   22764130. S2CID   19601358.
  50. Li J, Lin S, Wang J, Jia S, Yang M, Hao Z, et al. (May 2013). "Ligand-free palladium-mediated site-specific protein labeling inside gram-negative bacterial pathogens". Journal of the American Chemical Society. 135 (19): 7330–7338. doi:10.1021/ja402424j. PMID   23641876.
  51. Li N, Lim RK, Edwardraja S, Lin Q (October 2011). "Copper-free Sonogashira cross-coupling for functionalization of alkyne-encoded proteins in aqueous medium and in bacterial cells". Journal of the American Chemical Society. 133 (39): 15316–15319. doi:10.1021/ja2066913. PMC   3184007 . PMID   21899368.
  52. Li N, Ramil CP, Lim RK, Lin Q (February 2015). "A genetically encoded alkyne directs palladium-mediated protein labeling on live mammalian cell surface". ACS Chemical Biology. 10 (2): 379–384. doi:10.1021/cb500649q. PMC   4340352 . PMID   25347611.
  53. Chalker JM, Wood CS, Davis BG (November 2009). "A convenient catalyst for aqueous and protein Suzuki-Miyaura cross-coupling". Journal of the American Chemical Society. 131 (45): 16346–16347. doi:10.1021/ja907150m. PMID   19852502.
  54. Spicer CD, Davis BG (February 2011). "Palladium-mediated site-selective Suzuki-Miyaura protein modification at genetically encoded aryl halides". Chemical Communications. 47 (6): 1698–1700. doi:10.1039/C0CC04970K. PMID   21206952.
  55. Spicer CD, Triemer T, Davis BG (January 2012). "Palladium-mediated cell-surface labeling". Journal of the American Chemical Society. 134 (2): 800–803. doi:10.1021/ja209352s. PMID   22175226.
  56. Dumas A, Spicer CD, Gao Z, Takehana T, Lin YA, Yasukohchi T, et al. (April 2013). "Self-liganded Suzuki-Miyaura coupling for site-selective protein PEGylation". Angewandte Chemie. 52 (14): 3916–3921. doi:10.1002/anie.201208626. PMID   23440916.
  57. Brustad E, Bushey ML, Lee JW, Groff D, Liu W, Schultz PG (2008-10-13). "A genetically encoded boronate-containing amino acid". Angewandte Chemie. 47 (43): 8220–8223. doi:10.1002/anie.200803240. PMC   2873848 . PMID   18816552.
  58. Azie O, Greenberg ZF, Batich CD, Dobson JP (June 2019). "Carbodiimide Conjugation of Latent Transforming Growth Factor β1 to Superparamagnetic Iron Oxide Nanoparticles for Remote Activation". International Journal of Molecular Sciences. 20 (13): 3190. doi: 10.3390/ijms20133190 . PMC   6651417 . PMID   31261853.
  59. Lin L, Liu L, Zhao B, Xie R, Lin W, Li H, et al. (May 2015). "Carbon nanotube-assisted optical activation of TGF-β signalling by near-infrared light". Nature Nanotechnology. 10 (5): 465–471. Bibcode:2015NatNa..10..465L. doi:10.1038/nnano.2015.28. PMID   25775150.
  60. Lalli E, Sarti G, Boi C (2018). "Effect of the spacer arm on non-specific binding in membrane affinity chromatography". MRS Communications. 8 (1): 65–70. doi:10.1557/mrc.2018.4. S2CID   103064199.