Fig 1 - uploaded by Michael Hecht
Content may be subject to copyright.
Nuclear magnetic resonance spectroscopy solution structures of proteins S-824 and S-836. Top: Ribbon diagrams of four-helix bundle proteins. Bottom: Space-filling models of corresponding proteins with non-polar amino acids sequestered in the core (yellow) and polar amino acids on the exterior (red or blue). Figure adapted from Wei et al . (2003b) and Go et al . (2008). 

Nuclear magnetic resonance spectroscopy solution structures of proteins S-824 and S-836. Top: Ribbon diagrams of four-helix bundle proteins. Bottom: Space-filling models of corresponding proteins with non-polar amino acids sequestered in the core (yellow) and polar amino acids on the exterior (red or blue). Figure adapted from Wei et al . (2003b) and Go et al . (2008). 

Source publication
Article
Full-text available
Collections of de novo-designed proteins provide a unique opportunity to probe the functional potential of sequences that are stably folded, but were neither explicitly designed nor evolutionarily selected to perform any particular type of activity. A combinatorial library of folded proteins was designed previously using a strategy that exploits th...

Contexts in source publication

Context 1
... : binary code/directed evolution/peroxidase/protein design/synthetic biology Designing proteins de novo provides a unique opportunity to explore sequences, structures and functions that are chem- ically reasonable, but which have not been sampled by biological evolution. In the initial steps toward such explorations, we designed combinatorial libraries of a -helical sequences, determined the three-dimensional structures of several proteins from these libraries, and probed their potential for binding and catalytic function (Kamtekar et al ., 1993; Rojas et al ., 1997; Moffet et al ., 2000; Wei et al ., 2003a,b; Hecht et al ., 2004; Go et al ., 2008; Patel et al ., 2009; Fisher et al ., 2011; Arai et al ., 2012; Cherny et al ., 2012). Combinatorial libraries of a -helical proteins were constructed using the binary code strategy for protein design. A canonical alpha helix has 3.6 residues per turn. Accordingly, a non-polar amino acid is placed every 3 or 4 positions in the primary sequence, generating the following pattern: W † WW †† WW † WW †† W , where W and † represent arbitrarily chosen polar and non-polar residues, respectively (Kamtekar et al ., 1993; Hecht et al ., 2004). This pattern encodes an amphipathic helix, where all non-polar residues align on one face of the helix and all polar residues align on the opposite face. When four such helices are linked, the hydro- phobic effect drives them to pack against one another, thereby forming a four-helix bundle with non-polar residues pointing toward the protein core, and polar residues exposed to solvent. Several proteins from previously constructed libraries have been shown to form stable well-ordered structures (Wei et al ., 2003a). Furthermore, determination of the structures of two proteins (S-824 and S-836) provided evidence that these binary- patterned proteins folded into four-helix bundles as designed (Wei et al ., 2003b; Go et al ., 2008) (Fig. 1). Although the proteins in our binary-patterned libraries were not explicitly designed for any particular activities, characterization of several hundred arbitrarily chosen sequences showed that many of them bound the heme cofactor, and several of these novel heme proteins catalyzed peroxidase activity at levels substantially above background (Patel et al ., 2009). The observation of cofactor binding and catalytic activity in proteins from a library of a -helical bundles that had neither been designed nor selected ( in vivo or in vitro ) for function suggested that this superfamily of de novo proteins could be viewed as a novel ‘feedstock’ for evolution. Thus, by starting with a binary-patterned protein with relatively low-level activity, one can mimic natural selection by intro- ducing random mutations and screening for variants with improved activity. This process is reminiscent of numerous directed evolution experiments that have been conducted on various natural proteins (Joo et al ., 1999; Farinas et al ., 2001; Glieder et al ., 2002; Gould and Tawfik, 2005). However, there is a significant difference. Previous experiments in laboratory-based evolution started with sequences that were themselves products of billions of years of natural selection. In contrast, laboratory evolution of our binary- patterned sequences enables the exploration of evolutionary trajectories emanating from naive sequences that are not biased by natural selection for biological function in living organisms. Starting with the previously characterized binary-patterned proteins, S-824 and S-836, we demonstrate that mutagenesis of a de novo -designed protein, followed by screening of a relatively small number (hundreds or thousands) of variants can yield novel sequences with improved peroxidase activity. Error-prone polymerase chain reaction (PCR) using nucleotide analogs was used to introduce random mutations into genes encoding the novel proteins (Zaccolo et al ., 1996). The nucleotide analogs d-PTP (2 0 -deoxy-P-nucleoside-5 0 triphosphate) and 8-oxo-dGTP (8-oxo-2 0 -deoxyguanosine- 5 0 -triphosphate) were obtained from TriLink BioTechnologies. The products were gel purified (ZymoClean) and amplified using error-free PCR. The final PCR products were purified, concentrated (ZymoClean) and double-digested with NdeI and BamHI. Digestion mixtures were run on a low-melting 2% agarose gel, and the band at 300 bp was excised and purified. The mutation rate was three to four base mutations per sequence of 306 bp. Mutagenized gene sequences were cloned into the pET-3a vector. Previously, a stuffer fragment of lambda DNA had been cloned into the vector to ensure that only double-cut vector would be isolated following double di- gestion and gel purification. The purified double-digested vector was ligated with the library insert at 16 8 C for 12 h, desalted and electroporated into ElectroMAX competent cells (200 m l) at 2.3 kV (Transporator Plus) using a 2-mm gap cuvette. Transformation efficiency was 10 8 colony forming units/ m g DNA. After overnight growth, the colonies were scraped and pooled, and libraries of plasmid DNA were isolated for subsequent experiments. BL21(DE3) Escherichia coli cells were transformed via elec- troporation with plasmid library DNA, and the cells were grown overnight at 37 8 C on LB/amp (100 m g/ml) plates. The colonies were picked and placed in 96-deep well plates containing auto-induction media (600 m l) and grown overnight at 37 8 C on a plate shaker. Auto-induction media was prepared by combining glycerol (0.4% v/v), glucose (0.05% w/v) and a -lactose (0.2% w/v) in LB/amp (Studier, 2005). Auto-induction media facilitates overexpression of the target protein without the need to monitor cell density or add isopro- pyl b -D-1-thiogalactopyranoside (IPTG). Following protein expression, aliquots of 100 m l were saved and the remaining cells were harvested by centrifugation. The cells were dis- rupted by addition of BugBuster lysis solution (200 m l; Novagen) and activity buffer (300 m l, 50 mM Tris – HCl, pH 8). The lysed cells were centrifuged and the supernatant was assayed for peroxidase activity. Samples for the assay were prepared by mixing protein supernatant (50 m l), hemin chloride (5 m M final concentration; Sigma), ABTS (1 mg/ml final concentration; Sigma), H 2 O 2 (0.006% final concentration) in activity buffer (50 mM Tris –HCl, pH 8) for a final volume of 200 m l. After addition of H 2 O 2 , product formation was detected at 650 nm using a plate reader. BL21(DE3) cells were transformed and grown overnight on LB/amp (100 m g/ml) plates as described above. The colonies were replica plated using nitrocellulose filter paper (Millipore). The original cells were grown at 37 8 C until the colonies reformed. Duplicate colonies on the nitrocellulose filter paper were placed on LB/amp (100 m g/ml) IPTG (200 m g/ml) plates for 4 h at room temperature to induce protein expression. The cells were ...
Context 2
... purified double-digested vector was ligated with the library insert at 16 8 C for 12 h, desalted and electroporated into ElectroMAX competent cells (200 m l) at 2.3 kV (Transporator Plus) using a 2-mm gap cuvette. Transformation efficiency was 10 8 colony forming units/ m g DNA. After overnight growth, the colonies were scraped and pooled, and libraries of plasmid DNA were isolated for subsequent experiments. BL21(DE3) Escherichia coli cells were transformed via elec- troporation with plasmid library DNA, and the cells were grown overnight at 37 8 C on LB/amp (100 m g/ml) plates. The colonies were picked and placed in 96-deep well plates containing auto-induction media (600 m l) and grown overnight at 37 8 C on a plate shaker. Auto-induction media was prepared by combining glycerol (0.4% v/v), glucose (0.05% w/v) and a -lactose (0.2% w/v) in LB/amp (Studier, 2005). Auto-induction media facilitates overexpression of the target protein without the need to monitor cell density or add isopro- pyl b -D-1-thiogalactopyranoside (IPTG). Following protein expression, aliquots of 100 m l were saved and the remaining cells were harvested by centrifugation. The cells were dis- rupted by addition of BugBuster lysis solution (200 m l; Novagen) and activity buffer (300 m l, 50 mM Tris – HCl, pH 8). The lysed cells were centrifuged and the supernatant was assayed for peroxidase activity. Samples for the assay were prepared by mixing protein supernatant (50 m l), hemin chloride (5 m M final concentration; Sigma), ABTS (1 mg/ml final concentration; Sigma), H 2 O 2 (0.006% final concentration) in activity buffer (50 mM Tris –HCl, pH 8) for a final volume of 200 m l. After addition of H 2 O 2 , product formation was detected at 650 nm using a plate reader. BL21(DE3) cells were transformed and grown overnight on LB/amp (100 m g/ml) plates as described above. The colonies were replica plated using nitrocellulose filter paper (Millipore). The original cells were grown at 37 8 C until the colonies reformed. Duplicate colonies on the nitrocellulose filter paper were placed on LB/amp (100 m g/ml) IPTG (200 m g/ml) plates for 4 h at room temperature to induce protein expression. The cells were then lysed by placing the nitrocellulose paper on top of filter paper soaked in BugBuster lysis solution (Novagen). The nitrocellulose paper was then moved to filter paper soaked in hemin chloride solution (5 m M; Sigma) to allow proteins in the lysate to bind heme. Finally, the nitrocellulose paper was moved to a Petri dish containing ABTS (2 mg/ml, Sigma) and H 2 O 2 (0.06%) in activity buffer (50 mM sodium phosphate, pH 6). Colonies with peroxidase activity turned green and were marked. Replica colonies from the original plate were used to recon- firm activity, to make cell stocks and for plasmid preparations. A single colony was used to inoculate 10 ml of LB/amp (100 mg/l final concentration) which was grown overnight at 37 8 C. This liquid culture was used to inoculate 1 l of LB/ amp (100 mg/l final concentration) at a 1 : 1000 ratio. The 1 l culture was grown at 37 8 C until reaching an OD600 of 0.6 and IPTG (200 mg/l final concentration) was added to induce expression. Protein was released from cells using the freeze-thaw method (Johnson and Hecht, 1994). The protein was resuspended in MgCl 2 (100 mM; 10 ml per 1 l of culture) to extract the protein from the perforated cells, and cellular debris was removed by centrifugation. The supernatant was then acidified (1 M sodium acetate buffer, pH 4; 1 ml per 1 l of culture) and cellular contaminants were removed by acid precipitation and centrifugation at 6000 Â g for 10 min. The resulting supernatant was loaded onto a Poros HS cation exchange column and purified using a gradi- ent of NaCl. The protein was concentrated and exchanged into activity buffer using Centricon Plus-20 filters (Millipore). Samples for the peroxidase assay were prepared by mixing protein (20 m M) with hemin chloride (5 m M) and ABTS (1 mg/ml) in activity buffer (50 mM Tris – HCl, pH 8). A hemin chloride stock solution (1 mM) was prepared in dimethyl sulfoxide, and an ABTS stock solution (20 mg/ml) was prepared in activity buffer. As the heme concentration is much lower than protein concentration, we assume the maximal concentration of heme – protein complex is 5 m M. On addition of H 2 O 2 (0.086 m M – 11 m M), time points were recorded for 30 min at 650 nm. Kinetic data were recorded using the Varioskan Flash Spectral Scanning Multimode Reader and SkanIt software (Thermo Fisher Scientific, Waltham, MA). Peroxidase rate constants were analyzed using Michaelis – Menten kinetics: 1/ V 1⁄4 K M /( k cat [E 0 ][S]) þ 1/( k [E ]). The de novo- designed four-helix bundles, S-824 and S-836, were targeted for directed evolution. These sequences were chosen as test cases because both proteins had been characterized extensively, and their structures were solved previously by Nuclear magnetic resonance spectroscopy, as shown in Fig. 1 (Wei et al ., 2003a,b; Go et al ., 2008). Furthermore, as both proteins are very stable, it seemed likely that their scaf- folds would tolerate mutations. Mutations in both sequences were created by error-prone PCR using nucleotide analogs (Zaccolo et al ., 1996). Mutant libraries of both sequences were initially screened for peroxidase activity in 96-well plates. These screens provide clear results, but are difficult to scale up without robotic instrumentation. An example of a 96-well plate assay for peroxidase activity in cell lysates for a library of S-836 mutants is shown in Fig. 2. The range in the intensity of color indicates a range of activity in the mutant library. Not surprisingly, many mutant sequences display lower levels of activity than the parental S-836 sequence. However, several mutants appear to be more active than S-836. The sequences responsible for the darkest color were chosen for further rounds of mutagenesis and screening. To facilitate sampling of larger libraries, we also devel- oped a colony-based screen. For this screen, the cells were grown, expressed, lysed and assayed for activity in colonies growing on filter paper above agar Petri dishes. When the colonies are exposed to substrate, those that change color first are picked for further analysis. In our hands, the 96-well format was convenient for screening several hundred sequences, whereas the colony-based assay allowed screening of thousands of clones. Figure 3 shows a control experiment for a colony screen. Each plate contains colonies expressing a single known sequence. The negative control (left) shows colorless colonies expressing sequence WA9, which has no peroxidase activity. In contrast, the positive control (right) shows colored colonies for cells expressing the active sequence, WA63. Thus, colonies expressing sequences with different levels of peroxidase activity can be distinguished readily. Initially, we used 96-well plates to screen lysates for enhanced peroxidase activity amidst libraries of mutants of proteins S-824 and S-836. Figure 4a shows the results for a single 96-well plate of mutants of S-824. The activity of each mutant is plotted relative to the parental sequence. Not surprisingly, most mutations are deleterious, and most sequences are less active than the parental sequence (ratios , 1 in Fig. 4a). However, in this initial screen, four proteins ( 4% of the library) showed activity above the parent by 3 standard deviations (a value of 1.3). Next, we screened the S-836 library in four 96-well plates (Fig. 4b). In this case, 20% of the library showed activity at least 3 standard deviations above the parent. As a larger percentage of the S-836 library (compared with the S-824 library) showed improved activity, we chose this library for further studies. To examine a larger collection of sequences, we screened the library of S-836 mutants using the colony screen described above. Approximately 2 Â 10 3 colonies were screened. Those that turned green first were picked and reassayed in liquid media. The results of these assays are shown in Fig. 5. All of the picked clones showed greater activity than the parent. Next, we took the top hits from the 96-well plate screen (clones 1-2B11, 1-2G6) and from the colony screen (clones 1-C3, 1-C8) and mutagenized these to create a second- generation library of mutants. This library was produced in the same way as the first-generation library, using error-prone PCR and nucleotide analogs. The second-generation library of mutants was then screened in four 96-well plates, and the results are shown in Fig. 6. Two percent of the mutants showed activity significantly above the parent S-836. The top hits from the first-generation 96-well screen (1-2B11, 1-2G6), the first-generation colony screen (1-C3, 1-C8) and the second-generation 96-well screen (2-2H12, 2-3F1, 2-4C1) were then reassayed for peroxidase activity in cell lysates. This assay confirmed that all of the putative hits showed activity above the parental S-836 protein. These results are shown in Fig. 7 and the corresponding sequences are shown in Fig. 8. As the activity observed in a cell lysate depends on both the intrinsic activity of a protein and the level of its expression, we estimated expression levels by gel electrophoresis. As shown in Fig. 9, all sequences expressed at similar levels. To verify that the mutant sequences are indeed more active than the parental S-836 protein, we chose two proteins for purification and biochemical characterization. Sequences 1-2B11 and 2-2H12 were chosen because they displayed relatively high levels of activity in lysates (Fig. 7), and were straightforward to purify. Protein 1-2B11 had been isolated in the 96-well screen of the first-generation library, and protein 2-H12 was the top hit in the 96-well screen of the second-generation library. Sequence 2-H12 is a variant of 1-2B11 (Fig. 8). The kinetic profiles of the purified mutant proteins are compared with ...

Similar publications

Article
Full-text available
A global optimization algorithm i s introduced which generalizes Kushner's univariate search [1]. It aims to minimize the number o f probes (function evaluations) required for a g i v e n confidence in the results. All known p r o b e s contribute to a stochastic model of the underlying "score surface"; this model is interrogated for the location m...
Article
Full-text available
Circulating miRNAs are emerging as novel reliable biomarkers for early detection of cancer diseases. Through combining the advantages of electrochemical methods and nanomaterials with the selectivity of the oligo-hybridization-based biosensors, a novel electrochemical nanobiosensor for plasma miR-155 detection have demonstrated here, based on thiol...
Article
Lysine methyltransferase G9a regulates the transcription of multiple genes by primarily catalyzing mono- and di-methylation of histone H3 lysine 9, as well as several non-histone lysine sites. An attractive therapeutic target in treating leukemia, knockout studies of G9a in mice have found dramatically slowed proliferation and self-renewal of acute...
Article
Full-text available
Previous studies in speech production and acquisition have mainly focused on how feedback vs. goals and feedback vs. prediction regulate learning and speech control. The present study investigated the less studied mechanism-prediction vs. goals in the context of adult Mandarin speakers' acquisition of non-native sounds, using an auditory feedback m...

Citations

... [3] The latter has been utilized to tailor numerous enzyme properties, such as activity, stability, substrate specificity, enantioselectivity, and binding affinity. [4][5][6][7][8] Several recombination strategies have been developed to tailor proteins, ranging from recombining single-point mutations or DNA fragments stepwise to generating multiple points simultaneously ( Figure 1). In addition to the epPCR method, Stemmer and colleagues reported in 1994 that DNA shuffling is a rapid evolution method through reassembling random fragments with homologies. ...
Cover Page
Full-text available
Various substitution recombination methods in the field of protein engineering are described in this paper. The enzyme properties can be empowered by experimental and computer-assisted recombination methods. This cover feature shows the tree of substitution recombination in an impressionistic style, where the branches are made up of numerous protein sequences and the orbs on the branches symbolize beneficial substitutions; and the recombination tree can be ever-changing… More information can be found in the Review by X. Li, H. Cui and co-workers (DOI: 10.1002/chem.202303889).
... [3] The latter has been utilized to tailor numerous enzyme properties, such as activity, stability, substrate specificity, enantioselectivity, and binding affinity. [4][5][6][7][8] Several recombination strategies have been developed to tailor proteins, ranging from recombining single-point mutations or DNA fragments stepwise to generating multiple points simultaneously ( Figure 1). In addition to the epPCR method, Stemmer and colleagues reported in 1994 that DNA shuffling is a rapid evolution method through reassembling random fragments with homologies. ...
Article
Full-text available
Directed evolution stands as a seminal technology for generating novel protein functionalities, a cornerstone in biocatalysis, metabolic engineering, and synthetic biology. Today, with the development of various mutagenesis methods and advanced analytical machines, the challenge of diversity generation and high‐throughput screening platforms is largely solved, and one of the remaining challenges is: how to empower the potential of single beneficial substitutions with recombination to achieve the epistatic effect. This review overviews experimental and computer‐assisted recombination methods in protein engineering campaigns. In addition, integrated and machine learning‐guided strategies were highlighted to discuss how these recombination approaches contribute to generating the screening library with better diversity, coverage, and size. A decision tree was finally summarized to guide the further selection of proper recombination strategies in practice, which was beneficial for accelerating protein engineering.
... Importantly, these proteins were neither selected by nature nor do they have any relationship to natural biochemical pathways. Several proteins from this "alternate universe" of sequences have been shown to possess binding and/or catalytic activities both in vitro and in vivo (1,5,(14)(15)(16)(17)(18)(19)(20)(21)(22). The occurrence of enzymatic activity is especially surprising as these sequences have no relationship to natural proteins and were not explicitly designed to have catalytic activity. ...
Article
Full-text available
De novo proteins constructed from novel amino acid sequences are distinct from proteins that evolved in nature. Construct K (ConK) is a binary-patterned de novo designed protein that rescues Escherichia coli from otherwise toxic concentrations of copper. ConK was recently found to bind the cofactor PLP (pyridoxal phosphate, the active form of vitamin B6). Here, we show that ConK catalyzes the desulfurization of cysteine to H2S, which can be used to synthesize CdS nanocrystals in solution. The CdS nanocrystals are approximately 3 nm, as measured by transmission electron microscope, with optical properties similar to those seen in chemically synthesized quantum dots. The CdS nanocrystals synthesized using ConK have slower growth rates and a different growth mechanism than those synthesized using natural biomineralization pathways. The slower growth rate yields CdS nanocrystals with two desirable properties not observed during biomineralization using natural proteins. First, CdS nanocrystals are predominantly of the zinc blende crystal phase; this is in stark contrast to natural biomineralization routes that produce a mixture of zinc blende and wurtzite phase CdS. Second, in contrast to the growth and eventual precipitation observed in natural biomineralization systems, the CdS nanocrystals produced by ConK stabilize at a final size. Future optimization of CdS nanocrystal growth using ConK-or other de novo proteins-may help to overcome the limits on nanocrystal quality typically observed from natural biomineralization by enabling the synthesis of more stable, high-quality quantum dots at room temperature.
... 5,6 However, its use is limited to easy-to-screen phenotypes such as growth or color. [7][8][9] Targeted genome editing usually involves multiple rounds of genomic modification with one gene at a time, which is laborious and time consuming. The parallel disruption and/or integration of multicopy genes in the genome can not only speed up construction of strains, but also generate gene dosage diversity of multiple genes or pathways. ...
Article
Directed evolution and targeted genome editing have been deployed to create genetic variants with usefully altered phenotypes. However, these methods are limited to high-throughput screening methods or serial manipulation of single genes. In this study, we implemented multicopy chromosomal integration using CRISPR-associated transposases (MUCICAT) to simultaneously target up to 11 sites on the Escherichia coli chromosome for multiplex gene interruption and/or insertion, generating combinatorial genomic diversity. The MUCICAT system was improved by replacing the isopropyl-beta-D-thiogalactoside (IPTG)-dependent promoter to decouple gene editing and product synthesis and truncating the right end to reduce the leakage expression of cargo. We applied MUCICAT to engineer and optimize the N-acetylglucosamine (GlcNAc) biosynthesis pathway in E. coli to overproduce the industrially important GlcNAc in only 8 days. Two rounds of transformation, the first round for disruption of two degradation pathways related gene clusters and the second round for multiplex integration of the GlcNAc gene cassette, would generate a library with 1-11 copies of the GlcNAc cassette. We isolated a best variant with five copies of GlcNAc cassettes, producing 11.59 g/L GlcNAc, which was more than sixfold than that of the strain containing the pET-GNAc plasmid. Our multiplex approach MUCICAT has potential to become a powerful tool of cell programing and can be widely applied in many fields such as synthetic biology.
... D espite the significant advances in protein design, there still remain few examples of de novo enzymes constructed from bona fide, de novo protein scaffolds that both approach the catalytic efficiencies of their natural counterparts and are of potential use in an industrial or biological context (1)(2)(3)(4)(5)(6)(7)(8)(9)(10). This reflects the inherent complexities experienced in the biomolecular design process, where approaches are principally focused on either atomistically precise redesign of natural proteins to stabilize reaction transition states (1)(2)(3)(4) or imprinting the intrinsic chemical reactivity of cofactors or metal ions on simple, generic protein scaffolds (5)(6)(7)(8)(9)(10); both can be significantly enhanced by implementing powerful, yet randomized, directed evolution strategies to hone and optimize incipient function (11,12). While the latter approach often results in de novo proteins that lack a singular structure (6,13,14), the incorporation of functionally versatile cofactors, such as heme, is proven to facilitate the design and construction of de novo proteins and enzymes that recapitulate the function of natural heme-containing proteins in stable, simple, and highly mutable tetrahelical chassis (termed maquettes) (15)(16)(17). ...
Article
Full-text available
Significance While the bottom-up design of enzymes appears to be an intractably complex problem, a minimal approach that combines elementary, de novo-designed proteins with intrinsically reactive cofactors offers a simple means to rapidly access sophisticated catalytic mechanisms. Not only is this method proven in the reproduction of powerful oxidative chemistry of the natural peroxidase enzymes, but we show here that it extends to the efficient, abiological—and often asymmetric—formation of strained cyclopropane rings, nitrogen–carbon and carbon–carbon bonds, and the ring expansion of a simple cyclic molecule to form a precursor for NAD+, a fundamentally important biological cofactor. That the enzyme also functions in vivo paves the way for its incorporation into engineered biosynthetic pathways within living organisms.
... 10 Our motivation to explore repeat proteins that fold in helix− turn−helix structures as biocatalyst for new reactions lies in previous successful works of several groups on different enzymatic activities using alpha helices as the principal structural component and combining multiple helical elements. 29−36 The reported catalytic activities include hydrolytic activity (esterase activity), 32,34,36 transesterification activity, 32 acyl transfer activity, 29 peroxidase activity, 34,35 lipase activity, 34 and very recently, Diels−Alderase activity, 37 revealing the versatility of α-helical elements in catalyst design. Interestingly, helical based scaffolds led to catalytic activities through very different approaches. ...
... 38 Contrary to rational approaches, Hecht and coworkers identified several catalytic activities from unevolved libraries of helical proteins. 34,35 These studies revealed that proteins which have neither been designed for a specific function nor subjected to evolutionary pressure can provide enzymatic activities. Similarly, minimalist approaches to protein design, which emphasize the use of simple scaffolds and the minimal number of mutations required to achieve a given activity have shown to be successful. ...
Poster
In this Communication, our recent results on biomolecules with Huisgenase activity will be presented. The reaction tested consists of a formal (3+2) cycloaddition between in situ formed azomethine ylides and p-deficient alkenes to yield unnatural proline ester derivatives (Figure 1). The behavior of both Nmetallated and NH-azomethine ylides has been tested in the presence of DNA incorporating heterobimetallic [Pt(II),Cu(II)]-complexes1a and using Consensus Tetratricopeptide Repeats (CTPR) proteins, 1b respectively. The reactions have been carried out in water and in nonaqueous solvents.
... Heme-protein redesign, through scaffold engineering and/or cofactor replacement, afforded new enzymes with a variety of functionalities (Garner et al., 2011;Cai et al., 2013;Oohora et al., 2017). Further, de novo design approaches afforded the construction of artificial peroxidases with impressive enzymatic rate constants (Patel and Hecht, 2012;D'Souza et al., 2017;Watkins et al., 2017). ...
Article
Full-text available
Manganese-porphyrins are important tools in catalysis, due to their capability to promote a wide variety of synthetically valuable transformations. Despite their great reactivity, the difficulties to control the reaction selectivity and to protect the catalyst from self-degradation hamper their practical application. Compared to small-molecule porphyrin complexes, metalloenzymes display remarkable features, because the reactivity of the metal center is finely modulated by a complex interplay of interactions within the protein matrix. In the effort to combine the catalytic potential of manganese porphyrins with the unique properties of biological catalysts, artificial metalloenzymes have been reported, mainly by incorporation of manganese-porphyrins into native protein scaffolds. Here we describe the spectroscopic and catalytic properties of Mn-Mimochrome VI*a (Mn-MC6*a), a mini-protein with a manganese deuteroporphyrin active site within a scaffold of two synthetic peptides covalently bound to the porphyrin. Mn-MC6*a is an efficient catalyst endowed with peroxygenase activity. The UV-vis absorption spectrum of Mn-MC6*a resembles that of Mn-reconstituted horseradish peroxidase (Mn-HRP), both in the resting and high-valent oxidized states. Remarkably, Mn-MC6*a shows a higher reactivity compared to Mn-HRP, because higher yields and chemoselectivity were observed in thioether oxidation. Experimental evidences also provided indications on the nature of the high-valent reactive intermediate and on the sulfoxidation mechanism.
... Despite the significant advances in protein design, there still remain few examples of de novo enzymes that both approach the catalytic efficiencies of their natural counterparts and are of potential use in an industrial or biological context 1-7 . This reflects the inherent complexities experienced in the biomolecular design process, where approaches are principally focussed on either atomistically-precise redesign of natural proteins to stabilise reaction transition states 1-3 or imprinting the intrinsic chemical reactivity of cofactors or metal ions on simple, generic protein scaffolds 4-7 ; both can be significantly enhanced by implementing powerful, yet randomised directed evolution strategies to hone and optimise incipient function 8,9 . While the latter approach often results in de novo proteins that lack a singular structure 4,10,11 , the incorporation of functionally versatile cofactors, such as heme, is proven to facilitate the design and construction of de novo proteins and enzymes that recapitulate the function of natural heme-containing proteins in stable, simple and highly mutable tetrahelical chassis (termed maquettes) 12-14 . ...
Preprint
Full-text available
By constructing an in vivo assembled, catalytically proficient peroxidase, C45, we have recently demonstrated the catalytic potential of simple, de novo-designed heme proteins. Here we show that C45s enzymatic activity extends to the efficient and stereoselective intermolecular transfer of carbenes to olefins, heterocycles, aldehydes and amines. Not only is this the first report of carbene transferase activity in a de novo protein, but also of enzyme-catalyzed ring expansion of aromatic heterocycles via carbene transfer by any enzyme.
... The WA20 structure was an unusual dimeric structure with an intermolecularly folded (domainswapped) 4-helix bundle: each monomer (Bnunchaku^like structure) comprises two long α-helices that are intertwined with the helices of another monomer (Fig. 3c). From the third-generation combinatorial library of 4-helix bundle de novo proteins with the binary patterns, a variety of functional de novo proteins with cofactor binding, drug binding, and enzyme-like functions in vitro (Patel et al. 2009;Cherny et al. 2012;Patel and Hecht 2012) and life-sustaining functions in vivo (Fisher et al. 2011;Smith et al. 2015; Digianantonio a n d H e c h t 2 0 1 6 ; H o e g l e r a n d H e c h t 2 0 1 6 ; Digianantonio et al. 2017) have been successfully produced. Notably, the several binary-patterned de novo proteins which can rescue conditionally lethal gene deletions in E. coli auxotrophs (ΔserB, ΔgltA, ΔilvA, and Δfes) were reported for the first time (Fisher et al. 2011) (Fig. 3d). ...
Article
In multiscale structural biology, synthetic approaches are important to demonstrate biophysical principles and mechanisms underlying the structure, function, and action of bio-nanomachines. A central goal of “synthetic structural biology” is the design and construction of artificial proteins and protein complexes as desired. In this paper, I review recent remarkable progress of an array of approaches for hierarchical design of artificial proteins and complexes that signpost the path forward toward synthetic structural biology as an emerging interdisciplinary field. Topics covered include combinatorial and protein-engineering approaches for directed evolution of artificial binding proteins and membrane proteins, binary code strategy for structural and functional de novo proteins, protein nanobuilding block strategy for constructing nano-architectures, protein–metal–organic frameworks for 3D protein complex crystals, and rational and computational approaches for design/creation of artificial proteins and complexes, novel protein folds, ideal/optimized protein structures, novel binding proteins for targeted therapeutics, and self-assembling nanomaterials. Protein designers and engineers look toward a bright future in synthetic structural biology for the next generation of biophysics and biotechnology.
... Several of these systems exhibited properties highly desirable at the origins of life, such as autocatalysis, cross-catalysis, nearly exponential growth, the potential for network formation [44][45][46][47][48], and even chiral selectivity [49]. Hecht and co-workers investigated a number of catalytic peptides based on a four-helical bundle structure [50][51][52], including directed evolution studies [53] and divergent evolution from a generalist to two different specialists [54]. In their recent study [55], with the aid of circular dichroism size-exclusion chromatography and NMR, they found that some of their active, α-helical proteins formed structures that fluctuated between monomeric and dimeric states. ...
Article
Full-text available
Almost all modern proteins possess well-defined, relatively rigid scaffolds that provide structural preorganization for desired functions. Such scaffolds require the sufficient length of a polypeptide chain and extensive evolutionary optimization. How ancestral proteins attained functionality, even though they were most likely markedly smaller than their contemporary descendants, remains a major, unresolved question in the origin of life. On the basis of evidence from experiments and computer simulations, we argue that at least some of the earliest watersoluble and membrane proteins were markedly more flexible than their modern counterparts. As an example, we consider a small, evolved in vitro ligase, based on a novel architecture that may be the archetype of primordial enzymes. The protein does not contain a hydrophobic core or conventional elements of the secondary structure characteristic of modern water-soluble proteins, but instead is built of a flexible, catalytic loop supported by a small hydrophilic core containing zinc atoms. It appears that disorder in the polypeptide chain imparts robustness to mutations in the protein core. Simple ion channels, likely the earliest membrane protein assemblies, could also be quite flexible, but still retain their functionality, again in contrast to their modern descendants. This is demonstrated in the example of antiamoebin, which can serve as a useful model of small peptides forming ancestral ion channels. Common features of the earliest, functional protein architectures discussed here include not only their flexibility, but also a low level of evolutionary optimization and heterogeneity in amino acid composition and, possibly, the type of peptide bonds in the protein backbone.