Open Access Article
This Open Access Article is licensed under a
Creative Commons Attribution 3.0 Unported Licence

Synthetic biology for the directed evolution of protein biocatalysts: navigating sequence space intelligently

Andrew Currin abc, Neil Swainston acd, Philip J. Day ace and Douglas B. Kell *abc
aManchester Institute of Biotechnology, The University of Manchester, 131, Princess St, Manchester M1 7DN, UK. E-mail: dbk@manchester.ac.uk; Web: http://dbkgroup.org/@dbkell Tel: +44 (0)161 306 4492
bSchool of Chemistry, The University of Manchester, Manchester M13 9PL, UK
cCentre for Synthetic Biology of Fine and Speciality Chemicals (SYNBIOCHEM), The University of Manchester, 131, Princess St, Manchester M1 7DN, UK
dSchool of Computer Science, The University of Manchester, Manchester M13 9PL, UK
eFaculty of Medical and Human Sciences, The University of Manchester, Manchester M13 9PT, UK

Received 20th October 2014

First published on 15th December 2014


Abstract

The amino acid sequence of a protein affects both its structure and its function. Thus, the ability to modify the sequence, and hence the structure and activity, of individual proteins in a systematic way, opens up many opportunities, both scientifically and (as we focus on here) for exploitation in biocatalysis. Modern methods of synthetic biology, whereby increasingly large sequences of DNA can be synthesised de novo, allow an unprecedented ability to engineer proteins with novel functions. However, the number of possible proteins is far too large to test individually, so we need means for navigating the ‘search space’ of possible protein sequences efficiently and reliably in order to find desirable activities and other properties. Enzymologists distinguish binding (Kd) and catalytic (kcat) steps. In a similar way, judicious strategies have blended design (for binding, specificity and active site modelling) with the more empirical methods of classical directed evolution (DE) for improving kcat (where natural evolution rarely seeks the highest values), especially with regard to residues distant from the active site and where the functional linkages underpinning enzyme dynamics are both unknown and hard to predict. Epistasis (where the ‘best’ amino acid at one site depends on that or those at others) is a notable feature of directed evolution. The aim of this review is to highlight some of the approaches that are being developed to allow us to use directed evolution to improve enzyme properties, often dramatically. We note that directed evolution differs in a number of ways from natural evolution, including in particular the available mechanisms and the likely selection pressures. Thus, we stress the opportunities afforded by techniques that enable one to map sequence to (structure and) activity in silico, as an effective means of modelling and exploring protein landscapes. Because known landscapes may be assessed and reasoned about as a whole, simultaneously, this offers opportunities for protein improvement not readily available to natural evolution on rapid timescales. Intelligent landscape navigation, informed by sequence-activity relationships and coupled to the emerging methods of synthetic biology, offers scope for the development of novel biocatalysts that are both highly active and robust.


image file: c4cs00351a-p1.tif

Andrew Currin

Andrew Currin is a research associate at the Manchester Institute of Biotechnology, University of Manchester. He received his undergraduate degree (First Class Honours) in Biomedical Science in 2008 from the University of Birmingham, followed by a PhD in Biochemistry in 2012 at the University of Manchester. His work now focuses on the engineering of biocatalysts and developing DNA technology as a tool for synthetic biology. His interests lie in protein engineering, molecular biology, synthetic biology and drug discovery, with a particular focus on protein structure–function relationships and developing improved methodologies to investigate them.

image file: c4cs00351a-p2.tif

Neil Swainston

Neil Swainston is a Research Fellow at the Manchester Institute of Biotechnology, University of Manchester. Following several years of industrial experience in proteomics bioinformatics software development with the Waters Corporation, he began his research career 8 years ago in the Manchester Centre for Integrative Systems Biology. His research interests span 'omics data analysis and management, genome-scale metabolic modelling, and enzyme optimisation through synthetic biology, and he has published over 25 papers covering these subjects. Driving all of these interests is a continued commitment to software development, data standardisation and reusability, and the development of novel informatics approaches.

image file: c4cs00351a-p3.tif

Philip J. Day

Philip Day is Reader in Quantitative Analytical Genomics and Synthetic Biology, Manchester University, Philip leads interdisciplinary research for developing innovative tools in genomics and for pediatric cancer studies. Current research focuses on closed loop strategies for directed evolution gene synthesis and aptamer developments, and the development of active drug uptake using membrane transporters. Philip applies miniaturization for single cell analyses to decipher molecules per cell activities across heterogeneous cell populations. His research aims to providing exquisite quantitative data for systems biology applications and pathway analysis as a central theme for enabling personalised healthcare.

image file: c4cs00351a-p4.tif

Douglas B. Kell

Douglas Kell is Research Professor in Bioanalytical Science at the University of Manchester, UK. His interests lie in systems biology, iron metabolism and dysregulation, cellular drug transporters, synthetic biology, e-science, chemometrics and cheminformatics. He was Director of the Manchester Centre for Integrative Systems Biology prior to a 5 year secondment (2008–2013) as Chief Executive of the UK Biotechnology and Biological Sciences Research Council. He is a Fellow of the Learned Society of Wales and of the American Association for the Advancement of Science, and was awarded a CBE for services to Science and Research in the New Year 2014 Honours list.


Introduction

Much of science and technology consists of the search for desirable solutions, whether theoretical or realised, from an enormously larger set of possible candidates. The design, selection and/or improvement of biomacromolecules such as proteins represents a particularly clear example.1 This is because natural molecular evolution is caused by changes in protein primary sequence that (leaving aside other factors such as chaperones and post-translational modifications) can then fold to form higher-order structures with altered function or activity; the protein then undergoes selection (positive or negative) based on its new function (Fig. 1). Bioinformatic analyses can trace the path of protein evolution at the sequence level2–4 and match this to the corresponding change in function.
image file: c4cs00351a-f1.tif
Fig. 1 Relationship between amino acid sequence, 3D structure (and dynamics) and biocatalytic activity. Implicitly, there is a host in which these manipulations take place (or they may be done entirely in vitro). This is not a major focus of the review. Typically, a directed evolution study concentrates on the relationships between protein sequence, structure and activity, and the usual means for assessing these are outlined (within the boxes). Many methods are available to connect and rationalise these relationships and some examples are shown (grey boxes). Thorough directed evolution studies require understanding of each of these parameters so that the changes in protein function can be rationalised, thereby to allow effective search of the sequence space. The key is to use emerging knowledge from multiple sources to navigate the search spaces that these represent. Although the same principles apply to multi-subunit proteins and protein complexes, most of what is written focuses on single-domain proteins that, like ribonuclease,1342,1343 can fold spontaneously into their tertiary structures without the involvement of other proteins, chaperones, etc.

Proteins are nature's primary catalysts, and as the unsustainability of the present-day hydrocarbon-based petrochemicals industry becomes ever more apparent, there is a move towards carbohydrate feedstocks and a parallel and burgeoning interest in the use of proteins to catalyse reactions of non-natural as well as of natural chemicals. Thus, as well as observing the products of natural evolution we can now also initiate changes, whether in vivo or in vitro, for any target sequence. When the experimenter has some level of control over what sequence is made, variations can be introduced, screened and selected over several iterative cycles (‘generations’), in the hope that improved variants can be created for a particular target molecule, in a process usually referred to as directed evolution (Fig. 2) or DE. Classically this is achieved in a more or less random manner or by making a small number of specific changes to an existing sequence (see below); however, with the emergence of ‘synthetic biology’ a greater diversity of sequences can be created by assembling the desired sequence de novo (without a starting template to amplify from). Hence, almost any bespoke DNA sequence can be created, thus permitting the engineering of biological molecules and systems with novel functions. This is possible largely due to the reducing cost of DNA oligonucleotide synthesis and improvements in the methods that assemble these into larger fragments and even genomes.5,6 Therefore, the question arises as to what sequences one should make for a particular purpose, and on what basis one might decide these sequences.


image file: c4cs00351a-f2.tif
Fig. 2 The essential components of an evolutionary system. At the outset, a starting individual or population is selected, and one or more fitness criteria that reflect the objective of the process are determined. Next, the ability to rank these fitnesses and to select for diversity is created (by breeding individuals with variant sequences, introduced typically by mutation and/or recombination) in a way that tends to favour fitter individuals, this is repeated iteratively until a desired criterion is met.

In this intentionally wide-ranging review, we introduce the basis of protein evolution (sequence spaces, constraints and conservation), discuss the methodologies and strategies that can be utilised for the directed evolution of individual biocatalysts, and reflect on their applications in the recent literature. To restrict our scope somewhat, we largely discount questions of the directed evolution of pathways (i.e. series of reactions) or gene clusters (e.g.ref. 7 and 8) and of the choice9 or optimization of the host organism or expression conditions in which such directed evolution might be performed or its protein products expressed, nor the process aspects of any fermentation or biotransformation. We also focus on catalytic rate constants, albeit we recognize the importance of enzyme stability as well. Most of the strategies we describe can equally well be applied to proteins whose function is not directly catalytic, such as vaccines, binding agents, and the like. Consequently we intend this review to be a broadly useful resource or portal for the entire community that has an interest in the directed evolution of protein function. A broad summary is given as a mind map in Fig. 3, while the various general elements of a modern directed evolution program, on which we base our development of the main ideas, appears as Fig. 4.


image file: c4cs00351a-f3.tif
Fig. 3 A ‘mind map’1344 of the contents of this paper; to read this start at “twelve o'clock” and read clockwise.

image file: c4cs00351a-f4.tif
Fig. 4 An example of the basic elements of a mixed computational and experimental programme in directed evolution. Implicit are the choice of objective function (e.g. a particular catalytic activity with a certain turnover number) and the starting sequences that might be used with an initial or ‘wild type’ activity from which one can evolve improved variants. The core experimental (blue) and computational (red) aspects are shown as seven steps of an iterative cycle involving the creation and analysis of appropriate protein sequences and their attendant activities. Additional facets that can contribute to the programme are also shown (connected using dotted lines).

The size of sequence space

An important concept when considering a protein's amino acid sequence is that of (its) sequence space, i.e. the number of variations of that sequence that can possibly exist. Straightforwardly, for a protein that contains just the 20 main natural amino acids, a sequence length of N residues has a total number of possible sequences of 20N. For N = 100 (a rather small protein) the number 20100 (∼1.3 × 10[thin space (1/6-em)]130) is already far greater than the number of atoms in the known universe. Even a library with the mass of the Earth itself – 5.98 × 1027 g – would comprise at most 3.3 × 1047 different sequences, or a miniscule fraction of such diversity.10 Extra complexity, even for single-subunit proteins, also comes with incorporation of additional structural features beyond the primary sequence, like disulphide linkages, metal ions,11 cofactors and post-translational modifications, and the use of non-standard amino acids (outwith the main 20). Beyond this, there may be ‘moonlighting’ activities12 by which function is modified via interaction with other binding partners.

Considering sequence variation, using only the 20 ‘common’ amino acids, the number of sequence variants for M substitutions in a given protein of N amino acids is image file: c4cs00351a-t1.tif.13 For a protein of 300 amino acids with random changes in just 1, 2 or 3 amino acids in the whole protein this is 5700, ca. 16 million and ca. 30 billion, while even for a comparatively small protein of N = 100 amino acids, the number of variants exceeds 1015 when M = 10. Insertions can be considered as simply increasing the length of N and the number of variants to 21 (a ‘gap’ being coded as a 21st amino acid), respectively.

Consequently, the search for variants with improved function in these large sequence spaces is best treated as a combinatorial optimization problem,1 in which a number of parameters must be optimised simultaneously to achieve a successful outcome. To do this, heuristic strategies (that find good but not provably optimal solutions) are appropriate; these include algorithms based on evolutionary principles.

The ‘curse of dimensionality’ and the sparseness or ‘closeness’ of strings in sequence space

One way to consider protein sequences (or any other strings of this type) is to treat each position in the string as a dimension in a discrete and finite space. In an elementary way, an amino acid X has one of 20 positions in 1-dimensional space, an individual dimer XkYl has a specified position or represents a point (from 400 discrete possibilities) in 2D space, a trimer XkYlZm a specified location (from 8000) in 3D space, and so on. Various difficulties arise, however (‘the curse of dimensionality’14,15) as the number of dimensions increases, even for quite small numbers of dimensions or string length, since the dimensionality increases exponentially with the number of residues being changed. One in particular is the potential ‘closeness’ to each other of various randomly selected sequences, and how this effectively diverges extremely rapidly as their length is increased.

Imagine (as in ref. 16) that we have examples uniformly distributed in a p-dimensional hypercube, and wish to surround a target point with a hypercubical ‘neighbourhood’ to capture a fraction r of all the samples. The edge length of the (hyper)cube will be ep(r) = r(1/p). In just 10 dimensions e10(0.01) = 0.63 and e10(0.1) = 0.79 while the range (of a unit hypercube) for each dimension is just 1. Thus to capture even just 1% or 10% of the observations we need to cover 63% or 80% of the range (i.e. values) of each individual dimension. Two consequences for any significant dimensionality are that even large numbers of samples cover the space only very sparsely indeed, and that most samples are actually close to the edge of the n-dimensional hypercube. We shall return later to the question of metrics for the effective distance between protein strings and for the effectiveness of protein catalysts; for the latter we shall assume (and discuss below) that the enzyme catalytic rate constant or turnover number (with units of s−1, or in less favourable cases min−1, h−1, or d−1) is a reasonable surrogate for most functional purposes.

Overall, it is genuinely difficult to grasp or to visualise the vastness of these search spaces,17 and the manner in which even very large numbers of examples populate them only extremely sparsely. One way to visualise them18–22 is to project them into two dimensions. Thus, if we consider just 30mers of nucleic acid sequences, and in which each position can be A, T, G or C, the number of possible variants is 430, which is ∼1018, and even if arrayed as 5 μm spots the array would occupy 29 km2!23 The equivalent array for proteins would contain only 14mers, in that there are more than 1018 possible proteins containing the 20 natural amino acids when their length is just 14 amino acids.

The nature of sequence space

Sequence, structure and function

One of the fundamental issues in the biosciences is the elucidation of the relationship between a protein's primary sequence, its structure and its function. Difficulties arise because the relationship between a protein's sequence and structure is highly complex, as is the relationship between structure and function. Even single mutations at an individual residue can change a protein's activity completely – hence the discovery of ‘inborn errors of metabolism’.24,25 (The same is true in pharmaceutical drug discovery, with quite small changes in small molecule structure often leading to a dramatic change in activity – so-called ‘activity cliffs’26–33 – and with similar metaphors of structure–activity relationships, rather than those of sequence-activity, being equally explicit.34–37) Annotation of putative function from unknown sequences is largely based upon sequence homology (similarity) to proteins of known characterised function and particularly the presence of specific sequence/structure motifs (such as the Rossmann fold38 or the P-loop motif39). While there have been great advances in predicting protein structure from primary sequence (see later), the prediction of function from structure (let alone sequence) remains an important (if largely unattained) aim.40–54

How much of sequence space is ‘functional’?

The relationship between sequence and function is often considered in terms of a metaphor in which their evolution is seen as akin to traversing a ‘landscape’,55 that may be visualised in the same way as one considers the topology of a natural landscape,56,57 with the ‘position’ reflecting the sequence and the desirable function(s) or fitness reflected in the ‘height’ at that position in the landscape (Fig. 5).
image file: c4cs00351a-f5.tif
Fig. 5 A fitness landscape and its navigation. The initial or wild-type activity denotes the starting point (initialisation) for a directed evolution study (red circle). Accumulation of mutations that increase activity is represented by four routes to different positions in the landscape. Route 1 successfully increases activity through a series of additive mutations, but becomes stuck in a local optimum. Due to the nature of rugged fitness landscapes some of the shorter paths to a maximum possible (global optimum) fitness (activity) can require movement into troughs before navigating a new higher peak (route 2). Alternatively, one can arrive at the global optimum using longer but typically less steep routes without deep valleys (equivalent over flat ground to neutral mutations – routes 3 and 4).

Given the enormous numbers for populating sequence space, and the present impossibility of computing or sampling function from sequence alone, it is clear that natural evolution cannot possibly have sampled all possible sequences that might have biological function.58 Hence, the strategy of a DE project faces the same questions as those faced in nature: how to navigate sequence space effectively while maintaining at least some function, but introducing sufficient variation that is required to improve that function. For DE there are also the practical considerations: how many variants can be screened (and/or selected for) and analysed with our current capabilities?

The first general point to be made is that most completely random proteins are practically non-functional.10,56,59–66 Indeed, many are not even soluble,67,68 although they may be evolved to become so.69 Keefe and Szostak noted that ca. 1 in 1011 of random sequences have measurable ATP-binding affinity.70 Consistent with this relative sparseness of functional protein space is the fact that even if one does have a starting structure(/function), one typically need not go ‘far’ from such a structure to lose structure quite badly,71 albeit that with a ‘density’ of only 1 in 1011 proteins being functional this implies that all such functional sequences are connected by trajectories involving changes in only a single amino acid72 (and see ref. 58). This is also consistent with the fact that sequence space is vast, and only a tiny fraction of possible sequences tend to be useful and hence selected for by natural evolution. One may note70,73 that at least some degree of randomness will be accompanied by some structure,74,75 functionality or activity. For proteins, secondary structure is understood to be a strong evolutionary driver,76 particularly through the binary-patterning (arrangement of hydrophilic/hydrophobic residues),64,77–84 and so is the (somewhat related) packing density.85–89 In a certain sense, proteins must at some point have begun their evolution as more or less random sequences.90 Indeed “Folded proteins occur frequently in libraries of random amino acid sequences”,91 but quite small changes can have significantly negative effects.92 Harms and Thornton give a very thoughtful account of evolutionary biochemistry,4 recognizing that the “physical architecture {of proteins both} facilitates and constrains their evolution”. This means that it will be hard (but not impossible), especially without plenty of empirical data,93 to make predictions about the best trajectories. Fortunately, such data are now beginning to appear.57,94 Indeed, the leitmotiv of this review is that understanding such (sequence-structure–activity) landscapes better will assist us considerably in navigating them.

What is evolving and for what purpose?

In a simplistic way, it is easy to assume that protein sequences are being selected for on the basis of their contribution to the host organism's fitness, without normally having any real knowledge of what is in fact being implied or selected for. However, a profound and interesting point has been made by Keiser et al.95 to the effect that once a metabolite has been ‘chosen’ (selected) to be part of a metabolic or biochemical network, proteins are somewhat constrained to evolve as ‘slaves’, to learn to bind and react with the metabolites that exist. Thus, in evolution, the proteins follow the metabolites as much as vice versa, making knowledge of ligand binding96,97 and affinity98 to protein binding sites a matter of primary interest, especially if (as in the DE of biocatalysts) we wish to bind or evolve catalysts for novel (and xenobiotic) small molecule substrates. In DE we largely assume that the experimenter has determined what should be the objective function(s) or fitness(es), and we shall indicate the nature of some of the choices later; notwithstanding, several aspects of DE do tend to differ from those selected by natural evolution (Table 1). Thus, most mutations are pleiotropic in vivo,99,100 for instance. As DNA sequencing becomes increasingly economical and of higher throughput101,102 a greater provenance of sequence data enables a more thorough knowledge of the entire evolutionary landscape to be obtained. In the case of short sequences most103 or all104 of the entire genotype-fitness landscape may be measured experimentally. We note too (and see later) that there are equivalent issues in the optimization and algorithms of evolutionary computing (e.g.ref. 105–107), where strategies such as uniform cross-over,108 with no real counterpart in natural or experimental evolution, have been shown to be very effective.
Table 1 Some features by which natural evolution, classical DE of biocatalysts, and directed evolution of biocatalysts using synthetic biology differ from each other. Population structures also differ in natural evolution vs. DE, but in the various strategies for DE they follow from the imposed selection in ways that are difficult to generalize
Feature Natural evolution Classical DE DE with synthetic biology
Objective function and selection pressure Unclear; there is only a weak relation of a protein's function with organismal fitness;117kcat is not strongly selected for. Although presumably multi-objective, actual selection and fitness are ‘composites’. If there is no redundancy, organisms must retain function during evolution.58,118 Typically strong selection weak mutation (rarely was sequencing done so selection was based on fitness only). Can select explicitly for multiple outputs (e.g. kcat, thermostability). Much as with classical DE, but diversity maintenance can be much enhanced via high-throughput methods of DNA synthesis and sequencing.
Mutation rates Varies with genome size over orders of magnitude,119 but typically (for organisms from bacteria to humans) <10−8 per base per generation.120,121 Can itself be selected for.122 Mutation rates are controlled but often limited to only a few residues per generation, e.g. to 1/L where L is the aa length of the protein; much more can lead to too many stop codons. Library design schemes that permit stop codons only where required mean that mutation rates can be almost arbitrarily high.
Recombination rates Very low in most organisms (though must have occurred in cases of ‘horizontal gene transfer’); in some cases almost non-existent.123 Could be extremely high in the various schemes of DNA shuffling, including the creation of chimaeras from different parents. Again it can be as high or low as desired; the experimenter has (statistically) full control.
Randomness of mutation Although there are ‘hot spots’, mutations in natural evolution are considered to be random and not ‘directed’.124 In error-prone PCR, mutations are seen as essentially random. Site-directed methods offer control over mutations at a small number of specified positions. As much or as little randomness may be introduced as the experimenter desires by using defined mixtures of bases for each codon, e.g. NNN or NNK as alternatives to specific subsets such as polar or apolar.
Evolutionary ‘memory’ For individuals (cf. populations125) there is no ‘memory’ as such, although the sequence reflects the evolutionary ‘trace’ (but not normally the pathway – cf.ref. 126 and 127). Again, there is no real ‘memory’ in the absence of large-scale sequencing, but there is potential for it.56 With higher-throughput sequencing we can create an entire map of the landscape as sampled to date, to help guide the informed assessment of which sequences to try next.
Degree of epistasis It exists, but only when there is a more or less neutral pathway joining the epistatic sites. It is comparatively hard to detect at low mutation rates. Potentially epistasis is much more obvious as sites can be mutated pairwise or in more complex designed patterns.
Maintenance of individuals of lower or similar fitness in population They are soon selected out in a ‘strong selection, weak mutation’ regime; this limits jumps via lower fitness, and enforces at least neutral mutations. It is in the hands of the experimenter, and usually not done when only fitnesses are measured. Again it is entirely up to the experimenter; diversity may be maintained to trade exploration against exploitation.


However, in the case of multi-objective optimisation (e.g. seeking to optimise two objectives such as both kcat and thermostability, or activity vs. immunogenicity109), there is normally no individual preferred solution that is optimal for all objectives,110 but a set of them, known as the Pareto front (Fig. 6), whose members are optimal in at least one objective while not being bettered (not ‘dominated’) in any other property by any other individual. The Pareto front is thus also known as the non-dominated front or ‘set’ of solutions. A variety of algorithms in multi-objective evolutionary optimisation (e.g.ref. 111–116) use members of the Pareto front as the choice of which ‘parents’ to use for mutation and recombination in subsequent rounds.


image file: c4cs00351a-f6.tif
Fig. 6 A two-objective optimisation problem, illustrating the non-dominated or Pareto front. In this case we wish to maximise both objectives. Each individual symbol is a candidate solution (i.e. protein sequence), with the filled ones denoting an approximation to the Pareto front.

Protein folds and convergent and divergent evolution

What is certain, given that form follows function, is that natural evolution has selected repeatedly for particular kinds of secondary and tertiary structure ‘domains’ and ‘folds’.128,129 It is uncertain as to how many more are ‘common’ and are to be found via the methods of structural genomics,130 but many have been expertly classified,131e.g. in the CATH,132–134 SCOP135–137 or InterPro138,139 databases, and do occur repeatedly.

Given that structural conservation of protein folds can occur for sequences that differ markedly from each other, it is desirable that these analyses are done at the structural (rather than sequence) level (although there is a certain arbitrariness about where one fold ends and another begins140,141). Some folds have occurred and been selected via divergent evolution (similar sequences with different functions)142 and some via convergent evolution (different sequences with similar functions).143,144 This latter in particular makes the nonlinear mapping of sequence to function extremely difficult, and there are roughly two unrelated sequences for each E.C. (Enzyme Commission classification) number.145 As phrased by Ferrada and colleagues,146 “two proteins with the same structure and/or function in our data…{have} a median amino acid divergence of no less than 55 percent”. However, normally information is available only for extant molecules but not their history and precise evolutionary path (in contrast to DE). One conclusion might be that conventional means of phylogenetic analysis are not necessarily best placed to assist the processes of directed evolution, and we argue later (because a protein has no real ‘memory’ of its full evolutionary pathway) that modern methods of machine learning that can take into account ensembles of sequences and activities may prove more suitable. However, we shall first look at natural evolution.

Constraints on globular protein evolution structure in natural evolution

In gross terms, a major constraint on protein evolution is provided by thermodynamics, in that proteins will have a tendency to fold up to a state of minimum free energy.147–149 Consequently, the composition of the amino acids has a major influence over protein folding because this means satisfying, so far as is possible, the preference of hydrophilic or polar amino acids to bind to each other and the equivalent tendency of hydrophobic residues to do so.150–152 Alteration of residues, especially non-conservatively, often leads to a lowering of thermodynamic folding stability,153 which may of course be compensated by changes in other locations. Naturally, at one level proteins need to have a certain stability to function, but they also need to be flexible to effect catalysis. This is coupled to the idea that proteins are marginally stable objects in the face of evolution.154–159 Overall, this is equivalent to ‘evolution to the edge of chaos’,160,161 a phenomenon recognizing the importance of trading off robustness with evolvability that can also be applied162,163 to biochemical networks.164–170 Thermostability (see later) may also sometimes (but not always171–173) correlate with evolvability.174,175

Given the thermodynamic and biophysical157,176,177 constraints, that are related to structural contacts, various models (e.g.ref. 147 and 178) have been used to predict the distribution of amino acids in known proteins. As regards to specific mechanisms, it has been stated that “solvent accessibility is the primary structural constraint on amino acid substitutions and mutation rates during protein evolution.”,148 while “satisfaction of hydrogen bonding potential influences the conservation of polar sidechains”.179 Overall, given the tendency in natural evolution for strong selection, it is recognized that a major role is played by neutral mutations180–182 or neutral evolution183–188 (see Fig. 5 and 7). Gene duplication provides another strategy, allowing redundancy followed by evolution to new functions.189


image file: c4cs00351a-f7.tif
Fig. 7 Some evolutionary trajectories of a peptide sequence undergoing mutation. Mutations in the peptide sequence can cause an increase in fitness (e.g. enzyme activity, green), loss of fitness (salmon pink) or no change in fitness (grey). Typically, improved fitness mutations are selected for and subjected to further modification and selection. Neutral mutations keep sequences ‘alive’ in the series, and these can often be required for further improvements in fitness, as shown in steps 2 and 3 of this trajectory.

Coevolution of residues

Thus far, we have possibly implied that residues evolve (i.e. are selected for) independently, but that is not the case at all.190–192 There can be a variety of reasons for the conservation of sequence (including correlations between ‘distant’ regions193), but the importance to structure and function, and functional linkage between them, underlie such correlations.194–209 Covariation in natural evolution reflects the fact that, although not close in primary sequence, distal residues can be adjacent in the tertiary structure and may represent an interaction favourable to protein function. Covariation also provides an important computational approach to protein folding more generally (see below).

The nature, means of analysis and traversal of protein fitness landscapes

Since John Holland's brilliant and pioneering work in the 1970s (reprinted as ref. 210), it has been recognized that one can search large search spaces very effectively using algorithms that have a more or less close analogy to that of natural evolution. Such algorithms are typically known as genetic or evolutionary algorithms (e.g.ref. 106 and 211–213, and their implementation is referred to as evolutionary computing.106,214–216 The algorithms can be classified according to whether one knows only the fitnesses (phenotypes) of the population or also the genotypes (sequences).107

Since we cannot review the very large literature, essentially amounting to that of the whole of molecular protein evolution, on the nature of (natural) protein landscapes, we shall therefore seek to concentrate on a few areas where an improved understanding of the nature of the landscape may reasonably be expected to help us traverse it. Importantly, even for single objectives or fitnesses, a number of important concepts of ruggedness, additivity, promiscuity and epistasis are inextricably intertwined; they become more so where multiple and often incommensurate objectives are considered.

Additivity. Additivity implies simple continuing fixing of improved mutations,217–220 and follows from a model in which selection in natural evolution quite badly disfavours lower fitnesses,221 a circumstance known from Gillespie222,223 as ‘strong selection, weak mutation’ (SSWM, see also ref. 224–229). For small changes (close to neutral in a fitness or free energy sense), additivity may indeed be observed,230,231 and has been exploited extensively in DE.232–236 If additivity alone were true, however (and thus there is no epistasis for a given protein at all) then a rapid strategy for DE would be to synthesise all 20L amino acid variants at each position (of a starting protein of length L) and pick the best amino acid at each position. However, the very existence of convergent and divergent evolution implies that landscapes are rugged237 (and hence epistatic), so at the very least additivity and epistasis must coexist.236,238
Epistasis. The term ‘epistasis’ in DE covers a concept in which the ‘best’ amino acid at a given position depends on the amino acid at one or more other positions. In fact, we believe that one should start with an assumption of rather strong epistasis,238–248 as did Wright.55 Indeed the rugged fitness landscape is itself a necessary reflection of epistasis and vice versa. Thus, epistasis may be both cryptic and pervasive,249 the demonstrable coevolution goes hand in hand with epistasis, and “to understand evolution and selection in proteins, knowledge of coevolution and structural change must be integrated”.250
Promiscuity. The concept of enzyme promiscuity mainly implies that some enzymes may bind, or catalyse reactions with, more than one substrate, and this is inextricably linked to how one can traverse evolutionary landscapes.251–270 It clearly bears strongly on how we might seek to effect the directed evolution of biocatalysts.

NK landscapes as models for sequence-activity landscapes

A very important class of conceptual (and tunable) landscapes are the so-called NK landscapes devised by Kauffman161,271 and developed by many other workers (e.g.ref. 220, 221, 237 and 272–278). The ‘ruggedness’ of a given landscape is a slightly elusive concept,279 but can be conceptualized56,220 in a manner that implies that for a smooth landscape (like Mt Fuji280,281) fitness and distance tend to be correlated, while for a very ‘rugged’ landscape the correlation is much weaker (since as one moves away from a starting sequence one may pass through many peaks and troughs of fitness). In NK landscapes, K is the parameter that tunes the extent of ruggedness, and it is possible to seek landscapes whose ruggedness can be approximated by a particular value of K, since one of the attractions of NK is that they can reproduce (in a statistical sense) any kind of landscape.282 Indeed, we can use the comparatively sparse data presently available to determine that experimental sequence-fitness landscapes reflect NK landscapes that are fairly considerably (but not pathologically) rugged,23,57,104,241,251,274,276,283 and that there is likely to be one or more optimal mutation rates that themselves depend on the ruggedness (see later). Note too that the landscapes for individual proteins, as discussed here, are necessarily more rugged than are those of pathways or organisms, due to the more profound structural constraints in the former.57,157 (Parenthetically, NK-type landscapes and the evolutionary metaphor have also proved useful in a variety of other ‘complex’ spheres, such as business, innovation and economics (e.g.ref. 278 and 284–295, though a disattraction of NK landscapes in evolutionary biology itself is that they do not obey evolutionary rules.224)

Experimental directed protein evolution

A number of excellent books and review articles have been devoted to DE, and a sampling with a focus on biocatalysis includes.296–334 As indicated above, DE begins with a population that we hope contains at least one member that displays some kind of activity of interest, and progresses through multiple rounds of mutation, selection and analysis (as per the steps in Fig. 4).

Initialisation; the first generation

During the preliminary design of a DE project the main objective and required fitness criteria must be defined and these criteria influence the experimental design and screening strategy.

We consider in this review that a typical scenario is that one has a particular substrate or substrate class in mind, as well as the chemical reaction type (oxidation, hydroxylation, amination and so on) that one wishes to catalyse. If any activity at all can be detected then this can be a starting point. In some cases one does not know where to start at all because there are no proteins known either to catalyse a relevant reaction or to bind the substrate of interest. For pharmaceutical intermediates, it can still be useful to look for reactions involving metabolites, as most drugs do bear significant structural similarities to known metabolites,335,336 and it is possible to look for reactions involving the latter. A very useful starting point may be the structure-function linkage database http://sfld.rbvi.ucsf.edu/django/.337 There are also ‘hub’ sequences that can provide useful starting points,338 while Verma,330 Nov339 and Zaugg340 list various computational approaches. If one has a structure in the form of a PDB file one can try HotSpotWizard http://loschmidt.chemi.muni.cz/hotspotwizard/.341 Analysing the diversity of known enzyme sequences is also a very sensible strategy.342,343 Nowadays, an increasing trend is to seek relevant diversity, aligned using tools such as Clustal Omega,344,345 MUSCLE,346 PROMALS,347,348 or other methods based on polypharmacology,141,349,350 that one may hope contains enzymes capable of effecting the desired reaction. Another strategy is to select DNA from environments that have been exposed to the substrate of interest, using the methods of functional metagenomics.351,352 More commonly, however, one does have a very poor protein (clone) with at least some measurable activity, and the aim is to evolve this into a much more active variant.

In general, scientific advance is seen in a Popperian view (see e.g.ref. 353–357) as an iterative series of ‘conjectures’ and ‘refutations’ by which the search for scientific truth is ‘narrowed’ by finding what is not true (may be falsified) via predictions based on hypothetico-deductive reasoning and their anticipated and experimental outcomes. However, Popper was purposely coy about where hypotheses actually came from, and we prefer a variant358–362 (see also ref. 363 and 364) that recognises the equal contribution of a more empirical ‘data-driven’ arc to the ‘cycle of knowledge’ (Fig. 8).


image file: c4cs00351a-f8.tif
Fig. 8 The ‘cycle of knowledge’ in modern directed evolution. Both structure-based design and a more empirical data-driven approach can contribute to the evolution of a protein with improved properties, in a series of iterative cycles.

In a similar vein, many commentators (e.g.ref. 365–368) consider the best strategy for both the starting population and the subsequent steps to be a judicious blend between the more empirical approaches of (semi-)directed evolution and strategies more formally based on attempts to design369 (somewhat in the absence of fully established principles) sequences or structures based on what is known of molecular interactions. We concur with this, since at the present time it is simply not possible to design enzymes with high activities de novo (from scratch, or from sequence alone), despite progress in simple 4-helix-bundle and related ‘maquettes’.370–373 David Baker, probably the leading expert in protein design, considers that design is still incapable of predicting active enzymes even when the chemistry and active sites appear good.374,375 Several reviews attest to this,329,376–379 but crowdsourcing approaches have been shown to help,380 and computational design (and see below) certainly beats random sequences.381 Overall, the fairest comment is probably that we can benefit from design for binding, specificity and active site modelling, but that for improving kcat we need the more empirical methods of DE, especially (see below) of residues distant from the active site.

Scaffolds

Because natural evolution has selected for a variety of motifs that have been shown in general terms to admit a wide range of possible enzyme activities, a number of approaches have exploited these motifs or ‘scaffolds’.382 Triose phosphate isomerase (TIM) has proved a popular enzyme since the pioneering work of Albery and Knowles383 and more recent work on TIM energetics,384 and TIM (βα)8 barrels can be found in 5 of 6 EC classes.146 TIM and many (but not all) such natural enzymes are most active as dimers,385,386 caused by a tight interaction of 32 residues of each subunit in the wild type, though functional monomers can be created.387,388

Thus, (βα)8 barrel enzymes389–402 have proven particularly attractive as scaffolds for DE.403–407 Some use or need cofactors like PLP, FMN, etc.,392,408 and their folding mechanisms are to some degree known.385,409–411 We note, however, that virtual screening of substrates against these412 has shown a relative lack of effectiveness of consensus design because of the importance of correlations (i.e. epistasis).386

α/β and (α/β)2 barrels have also been favoured as scaffolds,395,413–417 while attempts at automated scaffold selection can also be found.374,418,419 A very interesting suggestion420 is that the polarity of a fold may determine its evolvability.

Although not focused on biocatalysis, other scaffolds such as lipocalins421–426 and affibodies427–437 have proved useful for combinatorial biosynthesis and directed evolution.

Computational protein design

While computational protein design completely from scratch (in silico) is not presently seen as reasonable, probably (as we stress later) because we cannot yet use it to predict dynamics effectively, significant progress continues to be made in a number of areas,373,438–456 including ‘fold to function’,457 combinatorial design,458 and a maximum likelihood framework for protein design.459 Notable examples include a metalloenzyme for organophosphate hydrolysis,460,461 aldolase462,463 and others.464–468 Theozymes469–472 (theoretical catalysts, constructed by computing the optimal geometry for transition state stabilization by model functional) groups represent another approach.

Arguably the most advanced strategies for protein design and manipulation in silico are Rosetta374,473–483 and RosettaBackrub,484,485 while more ‘bottom-up’ approaches, based on some of the ideas of synthetic biology, are beginning to appear.486–492 It is an easy prediction that developments in synthetic biology will have highly beneficial effects on de novo design, and vice versa.

Docking

If one is to find an enzyme that catalyses a reaction, one might hope to be able to predict that it can at least bind that substrate using the methods of in silico docking.493 To date, methods based on Autodock,494–499 APoC,500 Glide501–503 or other programs504–511 have been proposed, but this strategy is not yet considered mainstream for the DE of a first generation of biocatalysts (and indeed is subject to considerable uncertainty512). Our experience is that one must have considerable knowledge of the approximate answer (the binding site or pocket) before one tries these methods for DE of a biocatalyst.

Having chosen a member (or a population) as a starting point, the next step in any DE program is the important one of diversity creation. Indeed, the means of creating and exploiting suitable libraries that focus on appropriate parts of the protein landscape lies at the heart of any intelligent search method.513

Diversity creation and library design

A diversity of sequences can be created in many ways,514 but mutation or recombination methods are most commonly used in DE. Some are purely empirical and statistical (e.g. N mutations per sequence), while others are more focused to a specific part of the sequence (Fig. 9). Strategies may also be discriminated in terms of the degree of randomness of the changes and their extensiveness (Fig. 10). Two useful reviews include515 and,516 while others334,517–519 cover computational approaches. A DE library creation bibliography is maintained at http://openwetware.org/wiki/Reviews:Directed_evolution/Library_construction/bibliography.
image file: c4cs00351a-f9.tif
Fig. 9 Overview of the different mutagenesis strategies commonly employed to create variant protein libraries. Random methods (pink background) can create the greatest diversity of sequences in an uncontrolled manner. Mutations during error-prone PCR (A) are typically introduced by a polymerase amplifying sequences imperfectly (by being used under non-optimal conditions). In contrast, directed mutagenesis methods (blue background) introduce mutations at defined positions and with a controlled outcome. Site-directed mutagenesis (B) introduces a mutation, encoded by oligonucleotides, onto a template gene sequence in a plasmid. However, gene synthesis (C) can encode mutations on the oligonucleotides used to synthesise the sequence de novo, hence multiple mutations can be introduced simultaneously. X = random mutation, N = controlled mutation. →= PCR primer.

image file: c4cs00351a-f10.tif
Fig. 10 A Boston matrix of the different strategies for variant libraries. Methods are identified in terms of the randomness of the mutations they create and the number of residues that can be targeted.

Effect of mutation rates, implying that higher can be better

In classical evolutionary computing, the recognition that most mutations were or are deleterious meant that mutation rates were kept low. If only one in 103 sequences is an improvement when the mutation rate is 1/L per position (L being the length of the string), then (in the absence of epistasis) only 1 in 106 is at 2/L. (Of course 1/L is far greater than the mutation rates common in natural evolution, which scales inversely with genome size,119 may depend on cell–cell interactions,520 and is normally below 10−8 per base per generation for organisms from bacteria to humans.119–121) This logic is persuasive but limited, since it takes into account only the frequency but not the quality of the improvement (and as mentioned essentially does not consider epistasis). Indeed there is evidence that higher mutation rates are favoured both in silico220,521–524 and experimentally.525–528 This is especially the case for directed mutagenesis methods (especially those of synthetic biology), where stop codons can be avoided completely. We first discuss the more classical methods.

Random mutagenesis methods

Error-prone PCR (epPCR) is probably the most commonly used method for introducing random mutations. PCR amplification using Taq polymerase is performed under suboptimal conditions by altering the components of the reaction (in particular polymerase concentration, MgCl2 and dNTP concentration, or supplementation with MnCl2 (ref. 529)) or cycling conditions (increased extension times).530 Although epPCR is the simplest to implement and most commonly used method for library creation, it is limited by its failure to access all possible amino acid changes with just one mutation,339,531–533 a strong bias towards transition mutations (AT to GC mutations),531 and an aversion to consecutive nucleotide mutations.532,534

Refinement of these methods has allowed greater control over the mutation bias, rate of mutations530,535–537 and the development of alternative methodologies like Mutagenic Plasmid Amplification,538 replication,539 error-prone rolling circle540 and indel541–543 mutagenesis. Typically, for reasons indicated above, the epPCR mutation rate is tuned to produce a small number of mutations per gene copy (although orthogonal replication in vivo may improve this544), since entirely random epPCR produces multiple stop codons (3 in every 64 mutations) and a large proportion of non-functional, truncated or insoluble proteins.545 The library size also dictates that a large number of mutants must be screened to test for all possibilities, which may also be impractical depending on the screening strategy available. While random methods for library design can be successful, intelligent searching of the sequence space, as per the title of this review, does not include purely random methods.546 In particular, these methods do not allow information about which parts of the sequence have been mutated or whether all possible mutations for a particular region of interest have been screened.

Site-directed mutagenesis to target specific residues

Since the combinatorial explosion means that one cannot try every amino acid at every residue, one obvious approach is to restrict the number of target residues (in the following sections we will discuss why we do not think this is the best strategy for making faster biocatalysts). Indeed, mutagenesis directed at specific residues, usually referred to as site-directed mutagenesis,547,548 dates from the origins of modern protein engineering itself.549

In site-directed mutagenesis, an oligonucleotide encoding the desired mutation is designed with flanking sequences either side that are complementary to the target sequence and these direct its binding to the desired sequence on a template. This oligomer is used as a PCR primer to amplify the template sequence, hence all amplicons encode the desired mutation. This control over the mutation enables particular types of mutation to be made by using mixed base codons, i.e. codons that contain a mixture of bases at a specified position (e.g. N denotes an equal mixture of A, T, G or C at a single position). Fig. 11 shows a compilation of the more common types of mixed codons used. These range from those capable of encoding all 20 amino acids (e.g. NNK) to a small subset of residues with a particular physicochemical property (e.g. NTN for nonpolar residues only).


image file: c4cs00351a-f11.tif
Fig. 11 Examples of some of the common degenerate codons used in DE studies. A codon containing specific mixed bases is used to encode a particular set of amino acids, ranging from all twenty amino acids (NNN or NNK) to those with particular properties. Hence, choice of degenerate codons to use depends on the design and objective of the study. In the IUPAC terminology590 K = G/T, M = A/C, R = A/G, S = C/G, W = A/T, Y = C/T, B = C/G/T, D = A/G/T, H = A/C/T, V = A/C/G, N = A/C/G/T. (*Typically with low codon usage; suppressor mutation may be used to block it. **Typically with low codon usage, especially in yeast; suppressor mutation may be used to block it).

The most common method (QuikChange and derivatives thereof) uses mutagenic oligonucleotides complementary to both strands of a target sequence, which are used as primers for a PCR amplification of the plasmid encoding the gene. Following DpnI digestion of the template, the PCR product is transformed into E. coli and the nicked plasmid is repaired in vivo.550,551 Despite its popularity, QuikChange is somewhat limited by aspects like primer design and efficiency, and a variety of derivatives have been published that improve upon the original method.552,553

Given that site-directed mutagenesis provides a way of mutating a small number of residues with high levels of accuracy, several approaches have been developed to identify possible positions to target to increase the hit rate and success. Combinatorial alanine scanning554,555 is well known, while other flavours include the Mutagenesis Assistant Program,531,556 and the semi-rational CASTing and B-FIT approaches323,557 that employ a Mutagenic Plasmid Amplification method.558

In addition to these more conventional methods, new approaches are continually being developed to improve efficiency and to reduce the number of steps in the workflow, for example Mutagenic Oligonucleotide-Directed PCR Amplification (MOD-PCR),559 Overlap Extension PCR (OE-PCR),560–564 Sequence Saturation Mutagenesis,565–571 Iterative Saturation Mutagenesis,557,572–579 and a variety of transposon-based methods.580–583 However, a common issue with site-directed mutagenesis methods is the large number of steps involved and the limited number of positions that can be efficiently targeted at a time. The ability to mutate residues in multiple positions in a sequence is of particular interest as this can be used to address the question of combinatorial mutations simultaneously. Hence, methods like those by Liu et al.,584 Seyfang et al.,585 Fushan et al.586 and Kegler-Ebo et al.587 are important developments in mutagenesis strategies. Rational approaches have been reviewed,588 including from the perspective of the necessary library size.589 As a result, there is significant interest in the development of novel methodologies that can address these issues to produce accurate variant libraries, with larger numbers of simultaneous mutations in an economical workflow.

Optimising nucleotide substitutions

Following the selection of residues to target for mutation an important choice is the type of mutation to create. This choice is not obvious but determines the type of mutations that are made and the level of screening required. The experimenter needs to consider the nature of the mutations that they want to introduce for each position and this relates to the objective of the study. Using the common mixed base IUPAC terminology590 (Fig. 11) there are a large number of codons that can be chosen, ranging from those encoding all 20 amino acids (the NNK or NNS codons), to a particular characteristic (e.g. NTN encodes just nonpolar residues64) and a limited number of defined residues (GAN encoding just aspartate or glutamate). Importantly, choosing to use these specified mixed base codons in mutagenesis can reduce the possibility of premature stop codons and increase the chance of creating functional variants. For example, if a wild-type sequence encodes a nonpolar residue at a particular position then the number of functional variants is likely to be higher if the nonpolar codon NTN is used, encoding what are conserved substitutions, compared to encoding all possible residues with the NNK codon.591,592

Indeed, it is known to be better to search a large library sparsely than a small library thoroughly.593 Thus, a general strategy that seeks to move the trade-off between numbers of changes and numbers of clones to be assessed recognizes that one can design libraries that cover different general amino acid properties (such as charged, hydrophobic) while not encoding all 20 amino acids, thereby reducing (somewhat) the size of the search space. These are known as reduced library designs (see Fig. 11).

Reduced library designs

One limitation with the use of single degenerate codons is that for some sequences not all amino acids are equally represented and sometimes rare codons or stop codons are encoded. To circumvent this issue “small-intelligent” or “smart” libraries have been developed to provide equal frequency of each amino acid without bias.594 Using a mixture of oligonucleotides, Kille et al.595 created a restricted library with three codons NDT, VHG and TGG that encode 12, 9 and 1 codon, respectively. Together these encode 22 codons for all 20 amino acids in equal frequency, which provides good coverage of possible mutations but reduces the screening effort required to cover the sequence space. Alternative methods with the same objective include the MAX randomisation strategy596 and using ratios of different degenerate codons designed by software (DC-Analyzer597). Alternatively, the use of a reduced amino acid alphabet can also search a relevant sequence space whilst reducing the screening effort further. For example, the NDT codon encodes 12 amino acids of different physicochemical properties without encoding stop codons and has been shown to increase the number of positive hits (versus full randomization) in directed evolution studies.324 Overall, a considerable number of such strategies have been used (e.g.ref. 64, 67, 68, 81, 82, 324, 513, 556, 592 and 596–603).

The opposite strategy to reduced library designs is to increase them by modifying the genetic code. While one may think that there is enough potential in the very large search spaces using just 20 amino acids, such approaches have led to some exceptionally elegant work that bears description.

Non-canonical amino acid incorporation

If the existing protein synthetic machinery of the host cell is able to recognise a novel amino acid, it is possible to take an auxotroph and add the non-canonical amino acid (NCAA)604 that is thereby incorporated non-selectively. If one wishes to have site specificity, there are two main ways to increase the number of amino acids that can be incorporated into proteins.605 First, the specificity of a tRNA molecule (e.g. one encoding a stop codon) can be modified to accommodate non-canonical amino acids; in this way, the use of the relevant codon can introduce an NCAA at the specified position.606,607 Using this method, eight NCAAs were incorporated into the active site of nitroreductase (NTR, at Phe124) and screened for activity. One Phe analogue, p-nitrophenylalanine (pNF), exhibited more than a two-fold increase in activity over the best mutant containing a natural amino acid substitution (P124K), showing that NCAAs can produce higher enzyme activity than is possible with natural amino acids.608

The other, considerably more radical and potentially ground-breaking, is effectively to evolve the genetic code and other apparatus such that instead of recognising triplets a subset of mRNAs and the relevant translational machinery can recognise and decode quadruplets.609–619 To date, some 100 such NCAAs have been incorporated. However, the incorporation of NCAAs can often impact negatively on protein folding and thermostability, an issue that can be addressed through further rounds of directed evolution.620

Recombination

In contrast to the mutagenesis methods of library creation outlined above, but entirely consistent with our knowledge from strategies used in evolutionary computing (e.g.ref. 106), recombination is an alternative (or complementary) and effective strategy for DE (Fig. 12). Recombination techniques offer several advantages that reflect aspects of natural evolution that differ from random mutagenesis methods, not least because such changes can be combinatorial and hence able to search more areas of the sequence space in a given experiment. Recombination for the purposes of DE was popularized by Stemmer and his colleagues under the term ‘DNA shuffling’.621–625 This used a pool of parental genes with point mutations that were randomly fragmented by DNAseI and then reassembled using OE-PCR. Since then, a variety of further methods have been developed using different fragmentation and assembly protocols.626–629 Parental genes for DNA shuffling can be generated by random mutagenesis (epPCR) or from homologous gene families; such chimaeras may be particularly effective.630–633
image file: c4cs00351a-f12.tif
Fig. 12 The traditional recombination method for diversity creation. Recombination requires a sample of different variants of a gene (parents), which can be derived from a family of homologous genes or generated by random mutagenesis methods. The random fragmentation of these genes (using DNase I or other method) cleaves them into small constituent parts. Importantly, as the parental genes are all homologous, the fragments overlap in sequence thus allowing them to be reassembled by overlap extension PCR (OE-PCR) producing products that encode a random mixture of the parental genes. A key advantage of recombination methods is the improved ability to create combinatorial mutations. This is illustrated using two mutations (present in two different parental sequences) that when recombined separately produce no fitness improvement, but when combined together produce a variant with improved fitness.

Despite its advantages for searching wider sequence space, however, such recombination does not yield chimaeric proteins with balanced mutation distribution. Bias occurs in crossover regions of high sequence identity because the assembly of these sequences is more favourable during OE-PCR.634,635 As a result, this reduces the diversity of sequences in the variant library. Alternative methods like SCRATCHY636,637 generate chimaeras from genes of low sequence homology and so may help to reduce the extent of bias at the crossover points.

In addition to these more traditional methods of DNA shuffling, a number of variations have been developed (often with a penchant for a quirky acronym), such as Iterative Truncation for the Creation of HYbrid enzymes (ITCHY638,639), RAndom CHImeragenesis on Transient Templates (RACHITT),640 Recombined Extension on Truncated Templates (RETT),641 One-pot Simple methodology for CAssette Randomization and Recombination) OSCARR,642,643 DNA shuffling Frame shuffling,644 Synthetic shuffling,645 Degenerate Oligonucleotide Gene Shuffling (DOGS),646 USERec,647,648 SCOPE649–651 and Incorporation of Synthetic Oligos duRing gene shuffling (ISOR).652,653 Other methods of recombination that have been used for the improvement of proteins include the Protamax approach,654 DNA assembler,655,656 homologous recombination in vitro657 and Recombineering (e.g.ref. 658 and 659). Circular permutation, in which the beginning and end of a protein are effectively recombined in different places, provides a (perhaps surprisingly) effective strategy.17,660–667

There has long been a recognition that the better kind of chimaeragenesis strategies are those that maintain major structural elements,668,669 by ensuring that crossover occurs mainly or solely in what are seen as structurally the most ‘suitable’ locations. This is the basis of the OPTCOMB,670,671 RASPP,672 SCHEMA (e.g.ref. 673–683) and other types of approach.109,684–691

Thus, in the directed evolution of a cytochrome P450, Otey et al.674 utilized the SCHEMA algorithm to approximate the effect of recombination with different parent P450s on the protein structure. SCHEMA provided a prediction of preferred positions for crossovers, which enabled the creation of a mutant with a 40-fold higher peroxidase activity.673,678 Similarly, the recombination of stabilizing fragments was also able to increase the thermostability of P450s using the same approach.692

Cell-free synthesis

Although the majority of the mutations and recombinations described above have been performed in vitro, the actual expression of the proteins themselves, and the analysis of their functionalities, is usually done in vivo. However, we should mention a series of purely in vitro strategies that have also been used to identify good sequences when coupled to suitable in vitro translation systems with functional assays.693–700

Synthetic biology for directed evolution

With the recent improvements in DNA synthesis technology and reducing costs it is becoming increasingly feasible to synthesise sequences on a large scale. The most widely used methods for DNA synthesis continue to be short single-stranded oligodeoxyribonucleotides (typically 10–100 nt in length, often abbreviated to oligonucleotides or oligos) using phosphoramidite chemistry,701,702 although syntheses from microarrays have particular promise.546,703–708 Following synthesis, these oligonucleotides are assembled into larger constructs using enzymatic methods.

Hence, the foundation of synthetic biology is based on the ability to design and assemble novel biological systems ‘from the ground up’, i.e. synthetically at the DNA level.709–713 As a result, gene synthesis and genome assembly methods have been developed to create novel sequences of several kilobases in length.714 In particular, Gibson et al. recently assembled sections of the Mycoplasma genitalium genome (each 136 to 166 kb) using overlapping synthesised oligonucleotides.5,6

These developments in DNA synthesis technology (and lowered cost) can greatly benefit directed evolution studies. In particular, gene synthesis using overlapping oligonucleotides presents a particularly promising method for introducing controlled mutations into a gene sequence. As these methods assemble the gene de novo, multiple mutations at different positions in the gene can be introduced simultaneously in a single workflow, decreasing the need for iterative rounds of mutagenesis.

In this process, oligonucleotide sequences are designed to be overlapping and span the length of the gene of interest, following synthesis they are assembled by either PCR-based715,716 or ligation-based717–720 methods. Variant libraries can be created using this process by encoding mixed base codons on the oligonucleotides and at multiple positions if required.721 However, a limitation of the conventional gene synthesis procedure is the inherent error rate (primarily single base inserts or deletions),722,723 which arises from errors in the phosphoramidite synthesis of the oligonucleotides. As a result, clones encoding the desired sequence must be verified by DNA sequencing and an error-correction procedure is often required. Several error-correction methods are used, including site-directed mutagenesis,724 mismatch binding proteins725 and mismatch cleaving endonucleases.726,727 Of these, mismatch endonucleases are the most commonly used, and they are amenable to high throughput and automation.

SpeedyGenes and GeneGenie: tools for synthetic biology applied to the directed evolution of biocatalysts

Mismatch endonucleases recognise and cleave heteroduplexes in a DNA sequence. Consequently, they can be used as an effective method for the removal of errors during gene synthesis. However, when using mixed-base codons in directed evolution this is problematic, as these mixed sequences will form heteroduplexes and so will be heavily cleaved, thus preventing assembly of the required full-length sequence. Hence, we have developed an improved gene synthesis method, SpeedyGenes, which both improves the accurate synthesis of larger genes and can also accommodate mixed-base codon sequences.728 SpeedyGenes integrates a mismatch endonuclease step to cleave mismatched bases and, anticipating complete digestion of the mixed-base sequences, then restores these mixed base sequences by reintroducing the oligonucleotides encoding the mutation back into the PCR (“spiking in”) to allow the full length, error corrected gene to be synthesised. Importantly, multiple variant codons can be encoded at different positions of the gene simultaneously, enabling greater search of the sequence space through combinatorial mutations. This was illustrated728 by the synthesis of a monoamine oxidase (MAO-N) with three contiguous mixed-base codons mutated at two different positions in the gene. The known structure of MAO-N showed that the side chains of these residues were known to interact, hence these libraries could be screened for combinatorial coevolutionary mutations.

As with most synthetic biology methods, the use of sequence design in silico is crucial to the successful synthesis in vitro. In the case of SpeedyGenes, a parallel, online software design tool, GeneGenie, was developed to automate the design of DNA sequences and the desired variant library.729 By calculating the melting temperature (Tm) of the overlapping sequences, and minimising the potential mis-annealing of oligomers, GeneGenie greatly improves the success rate of assembly by PCR in vitro. In addition, codons are selected according to the codon usage of the expression host organism, and cloning sequences can be encoded ab initio to facilitate downstream cloning. Importantly, any mixed base codon can be added to incorporate into the designed sequence, hence automating the design of the variant library. As an example, a limited library of enhanced green fluorescent protein (EGFP) were designed to encode two variant codons (YAT at Y66 and TWT at Y145), the product of which would encode a limited variant library of green and blue variants of EGFP728 (Fig. 13).


image file: c4cs00351a-f13.tif
Fig. 13 GeneGenie and SpeedyGenes: synthetic biology tools for the purposes of directed evolution. The integration of computational design and accurate gene synthesis methodology provide a strong platform that can be utilised for directed evolution. As an example, the design, synthesis and screening of a small library of EGFP variants is shown. Mixed base codons are used to encode the green and blue variants of EGFP in a single library. (A) GeneGenie (http://www.gene-genie.org/) designs overlapping oligonucleotides for a given protein together with any specific mixed base codon (here YAT denoting C/T,A,T). (B) SpeedyGenes assembles the gene sequence using these oligonucleotides, accurately (using error correction) producing variant libraries with the desired mutations. (C) Direct expression (no pre-selection) of the library in E. coli yielded colonies with the desired mutations (green or blue fluorescence).

Genetic selection and screening

An important aspect of any experiment exploiting directed evolution for the development of improved biocatalysts is how one determines which of the many millions (or more) of the different clones that are created is worth testing further and/or retaining for subsequent generations. If it is possible to include a (genetic) selection step prior to any screening, this is always likely to prove valuable.303,730–732

Genetic selection

Most strategies for selection are unique to the protein of interest, and hence need to be designed empirically. Generally, this entails selection of a clone containing a desirable protein because it leads the cell to have a higher fitness.599,733 Examples including those based on enantioselectivity,734,735 substrate utilisation,736 chemical complementation,737,738 riboswitches,739–743 and counter-selection744 can be given. An ideal is when the selection rescues cells from a toxic insult that would otherwise kill them745 (see Fig. 14) or repairs a growth defect746–748). Two such examples749,750 of genetic selection are based on transporter engineering. However, most of the time it is quite difficult to develop such a genetic selection assay, so one must resort to screening.
image file: c4cs00351a-f14.tif
Fig. 14 The principle of genetic selection, here illustrated with a transporter gene knockout mutant in competition with others749 that does not take up toxic levels of an otherwise cytotoxic drug D.

Screening

Microtitre plates are the standard in biomolecular screening, and this is no different in DE.751 Herein, clones are seeded such that one clone per well is cultured, the substrates added, and the activity or products screened, primarily using chromogenic or fluorogenic substrates. This said, flow cytometry and fluorescence-activated cell sorting (FACS) have the benefit of much higher throughputs and have been widely applied (e.g.ref. 415 and 752–790) (and see below for microchannels and picodroplets). 2D arrays using immobilized proteins may also be used.791,792 However, not all products of interest are fluorescent, and these therefore need alternative methods of detection.

Thus, other techniques have included Raman spectroscopy for the chemical imaging of productive clones,793,794 while IR spectroscopy has been used to assess secondary structure (i.e. folding).795 Various constructs have been used to make non-fluorescent substrates or product produce a fluorescence signal.796 These include substrate-induced gene expression screening797–799 and product-induced gene expression,800 fluorescent RNAs,801 reporter bacteria,773,802 the detection of metabolites by fluorogenic RNA aptamers,803–811 colourimetric aptamers and Au particles,812 or appropriate transcription factors.787 Riboswitches that respond to product formation,742,743 chemical tags,813,814 and chemical proteomics815 have also been used as reporters for the production of small molecules.

Solid-phase screening with digital imaging is another alternative used for the engineering of biocatalysts. These methods generally use microbial colonies expressing the protein of interest to screen for activity directly in situ.816–818 Advantages to this include the ability to use enzyme-coupled assays (like HRP)819,820 or substrates of poor solubility or viscosity.821

Microfluidics, microdroplets and microcompartments

Sometimes the ‘host’ and the screen are virtually synonymous, as this kind of miniaturisation can also offer considerable speeds.822–825 Thus, there are trends towards the analysis of directed evolution experiments in microcompartments,766,826–831 using suitable microfluidics777,832–838 or picodroplets.831,839–843 Agresti et al.844 have shown that microfluidics using picolitre-volume droplets can screen a library of 108 HRP mutants in 10 hours. Although further refinement of microfluidics-based screening is required before its use becomes commonplace, it is clear that it has the capability to process the larger and more diverse libraries that one wishes to investigate.

Assessment of diversity and its maintenance

By now we have acquired a population of clones that are ‘better’ in some sense(s) than those of their parents. If we measure only fitnesses, however, as we have implicitly done thus far, we have only half the story, and we now return to the question of using knowledge of where we are or have been in a search space to optimize how we navigate it. There is of course a considerable literature on the role of ‘genetic’ and related searches in all kinds of single and multi-objective optimisation (see e.g.ref. 106, 107, 110, 113, 116, 210–213 and 845–858), all of which recognises that there is a trade-off between ‘exploration’ (looking for productive parts of the landscape) and ‘exploitation’ (performing more local searches in those parts). Methods or algorithms such as ‘efficient global optimisation’859 calculate these explicitly. Of course ‘where’ we are in the search space is simply encoded by the protein's sequence.

There is thus an increasing recognition that for the assessment860–863 and maintenance864 of diversity under selection one needs to study sequence-activity relationships. When DNA sequencing was much more expensive, methods were focused on assessing functionally important residues (e.g.ref. 865–868). As sequences became more readily available, methods such as PROSAR219,232,233,869 were used to fix favourable amino acids, a strategy that proved rather effective (albeit that it does not consider epistasis). Now (although sequence-free methods are also possible340,870–872), as large-scale DNA (including ‘next-generation’) sequencing becomes commonplace in DE,873–876 we may hope to see large and rich datasets becoming openly available to those who care to analyse them.

Sequence-activity relationships and machine learning

A historically important development in what is nowadays usually known as machine learning (ML)877–879 was the recognition that it is possible to learn relationships (in the form of mathematical models) between paired inputs and outputs – in the classical case between mass spectra and the structures of the molecules that had generated them880–884 – and more importantly that one could apply such models successfully in a predictive manner to molecules and spectra not used in the generation of the model. Such models are thus said to ‘learn’, or to ‘generalise’ to unseen samples (Fig. 15).
image file: c4cs00351a-f15.tif
Fig. 15 The principles of building and testing a machine learning model, illustrated here with a QSAR model. We start with paired inputs and outputs (here sequences and activities) and learn a nonlinear mapping between the two. Methods for doing this that we have found effective include genetic programming1345 and random forests.23 In a second phase, the learned model is used to make predictions on an unseen validation and/or test set1346 to establish that the model has generalized well.

In a similar vein, the first implementation of the idea that one could learn a mathematical model that captured the (normally rather nonlinear) relationships between a macromolecule's sequence and its activity in an assay of some kind, and thereby use that model to predict (in silico) the activities of untested sequences, seems to be that of Jonsson et al.885 These authors885 used partial least squares regression (a statistical model rather than ML – for the key differences see ref. 886) to establish a ‘quantitative sequence-activity model’ (QSAM) between (a numerical description of) 68-base-pair fragments of 25 E. coli promoters and their corresponding promoter strengths. The QSAM was then used to predict two 68 bp fragments that it was hoped would be more potent promoters than any in the training set. While extrapolation, to ‘fitnesses’ beyond what had been seen thus far, was probably a little optimistic, this work showed that such kinds of mappings were indeed possible (e.g.ref. 887–891). We have used such methods for a variety of protein-related problems, including predicting the nature and visibility of protein mass spectra.892–894

As a separate example, we used another ML method known as ‘random forests’895 to learn the relationship between features of some 40[thin space (1/6-em)]000 macromolecular (DNA aptamer) sequences and their activities,23 and could use this to predict (from a search space some 14 orders of magnitude greater) the activities of previously untested sequences. While considerable work is going on in structural biology, we are always going to have very many more (indeed increasingly more) sequences than we have structures; thus we consider that approaches such as this are going to be very important in speeding up DE in biocatalysis and improving the functional annotation of proteins. In particular, those performing directed evolution can have simultaneous access to all sequences and activities for a given protein.896,897 In contrast, an individual protein undergoing natural evolution cannot in any sense have a detailed ‘memory’ of its evolutionary past or pathway and in any event cannot (so far as is known, but cf.ref. 122 and 898) itself determine where to make mutations (only what to select on the basis of a poorly specified fitness). Machine learning methods seem extremely well suited for searching landscapes of this type.23,56,107,677,899 Overall, this is a very important difference between natural evolution and (Experimenter-) Directed Evolution.

The objective function(s): metrics for the effectiveness of biocatalysts

This is not a review of enzyme kinetics and kinetic mechanisms,549,900–902 and for our purposes we shall mainly assume that we are dealing with enzymes that catalyse thermodynamically favourable reactions, operating via a Michaelis–Menten type of reaction whose kinetic properties can largely be characterized via binding or Michaelis constants plus a (slower) catalytic rate constant kcat that is equivalent to the enzyme's turnover number (with units of reciprocal time). Much literature (e.g.ref. 549, 902 and 903) summarises the view that an appropriate measure of the effectiveness of an enzyme is a high value of kcat/Km, effected via the transduction of the initial energy of substrate/cofactor binding.903–905 Certainly the lowering of Km alone is a very poor target for most purposes in directed evolution where initial substrate concentrations are large. Better (as an objective function) than enantiomeric excess for chiral reactions producing a preferred R form (preferred over the S form) is a P factor or E factor (kcat,R/Km,R)/(kcat,S/Km,S)906 of a product. For industrial purposes, we are normally much more interested in the overall conversion in a reactor, rather than any specific enzyme kinetic parameter. Hence, the space-time yield (STY) or volume–time output (VTO) over a specified period, whose units are expressed in amount × (volume × time)−1 (e.g.ref. 907–911) has also been preferred as an objective function. This is clearly more logical from the engineering point of view, but for understanding how best to drive directed evolution at the molecular level, it is arguably best to concentrate on kcat, i.e. the turnover number, which is what we do here.

The distribution of kcat values among natural proteins

Not least because of the classic and beautiful work on triose phosphate isomerase, an enzyme that is operating almost at the diffusion-controlled limit,383,912 there is a quite pervasive view that natural evolution has taken enzymes ‘as far as they can go’ to make ‘proficient’ enzymes (e.g.ref. 913–915). Were this to be the case, there would be little point in developing directed evolution save for artificial substrates. However, it is not; most enzymes operate in vivo (and in vitro) at rates much lower than diffusion-controlled limits916,917 (online databases of enzyme kinetic parameters include BRENDA918 and SABIO-RK919). One assumes that this is largely because evolution simply had no need (i.e. faster enzymes did not confer sufficient evolutionary advantage)920 to select for them to increase their rates beyond that necessary to lose substantial flux control (a systems property921–925). It is this in particular that makes it both desirable and possible to improve kcat or kcat/Km values over what Nature thus far has commonly achieved.

In biotransformations studies, most papers appear to report processes in terms of g product × (g enzyme × day)−1; while process parameters are important,907 this serves (and is probably designed) to hide the very poor molecular kinetic parameters that actually pertain. Km is largely irrelevant because the concentrations in use are huge; thus our focus is on kcat. While DE has been shown to be capable of improving enzyme turnover numbers significantly, calculations show that even the ‘poster child’ examples (prositagliptin ketone transaminase,926 ∼0.03 s−1; halohydrin dehydrogenase,219 ∼2 s−1; isopropylmalate dehydrogenase,927 ∼5 s−1; lovD,368 ∼2 s−1) have turnover numbers that are very poor compared to those typical of primary metabolism, let alone the diffusion-controlled rates (depending on Km) of nearer 106–107 s−1.916,917

What enzyme properties determine kcat values?

Almost since the beginning of molecular enzymology, scientists have come to wonder what particular features of enzymes are ‘responsible’ for enzyme catalytic power (i.e. can be used to explain it, from a mechanistic point of view).928–930 It is implausible that there will be a unitary answer, as different sources will contribute differently in different cases. Scientifically, one may assume from the many successes of protein engineering that comparing various related sequences (and structures) by homology will be productive for our understanding of enzymology. Directed evolution studies increase the opportunities massively.

In general terms (e.g.ref. 930–933), preferred contributions to mechanisms have their different proponents, with such contributions being ascribed variously to a ‘Circe effect’,904 strain or distortion,934 electrostatic pre-organisation,933,935–939 hydrogen tunneling,940–947 reorganization energy,948 and in particular various kinds of fluctuations and enzyme dynamics.254,379,930,942,944–946,949–978 Less well-known flavours of dynamics include the idea that solitons may be involved.979,980 Overall, we consider that the ‘dynamics’ view of enzyme action is especially attractive for those seeking to increase the turnover number of an enzyme.

This is because what is not in doubt is that following substrate binding at the active site (that is dominantly responsible for substrate affinity and the degree of specificity), the binding energy has been ‘used up’ (Fig. 16) and is not thereby available to drive the catalytic step in a thermodynamic sense. This means that the protein must explore its energy landscape via conformational fluctuations that are essentially isoenergetic,951,966,981 before finding a configuration that places the active site residues into positions appropriate to effect the chemical catalysis itself, that happens as a ‘protein quake’981,982 in picoseconds or less.983 The source of these motions, whether normal mode or otherwise,984,985 can only be the protein and solvent fluctuations in the heat bath, and this means that their origins can lie in any parts of the protein, not just the few amino acids at the active site. Two exceedingly important corollaries of this follow. The first is that one may hope to predict or reflect this through the methods of molecular dynamics even when the active site is essentially unchanged, and this has recently been shown368). The second corollary is that one should expect it to be found that successful directed evolution programs that increase kcat lead to many mutations that are very distant from the active site. This can also serve, at least in part, to account for why surface post-translational modifications such as glycosylation can have significant effects on turnover (e.g.ref. 986).


image file: c4cs00351a-f16.tif
Fig. 16 A standard representation of an energy diagram for enzyme catalysis. Substrate binding is thermodynamically favourable, but to effect the catalytic reaction thermal energy is used to take the reaction to the right, often shown as a barrier represented by one or more ‘transition states’. Changes in the Km and Kd (affecting substrate affinity) can be influenced most directly by mutagenesis of the residues at the active site whilst changes in the kcat occur primarily from mutagenesis of residues away from the active site (which can affect the fluctuations in enzyme structure required either for crossing the transition state ES or by tunnelling under the barrier). At all points there are multiple roughly iso-energetic conformational (sub)states. Figure based on elements of those in ref. 930 and 966.

The importance of non-active-site mutations in increasing kcat values

In general terms, it is known that there is a considerable amount of long-range allostery in proteins,987 such that distant mutations couple to the active site.988 Indeed, most mutations (and amino acid residues) are necessarily distant from the active site, and there is a lack of correlation between a mutation's influence on kcat and its proximity to the active site989 (by contrast, specificity is determined much more by the closeness to the active site990). We still do not have that much data, since we require 3D structural information on many related variants; however, a number of excellent examples (Table 2) are indeed consistent with this recognition that effective DE strategies that raise kcat require that we spread our attention throughout the protein, and do not simply concentrate on the active site (Fig. 17). Indeed mutants with major improvements in kcat may display only minor changes in active site structure.368,938,974,991
Table 2 Some examples of improvements in biocatalytic activities that have been achieved using directed evolution, focusing on examples where most relevant mutations are in amino acids that are distal to the active site
Target Fold improvement over starting point Ref. Other notes
Cytochrome P450 9000 56 and 992 20 from generation 5, more than 15 away from active site
Diels–Alderase k cat 108-fold; catalytic power 9000-fold 993 21 aa, 16 outside active site
Glycerol dehydratase 336 994 2 aa, both very distant from active site
Glyphosate acyltransferase 200 in kcat 995 21 mutations, only 4 at active site
Halohydrin dehalogenase 4000 in volumetric productivity 219 35 mutations, only 8 at active site
3-Isopropylmalate dehydrogenase 65 927 8 mutations, 6 distant from active site
LovD >1000 368 29 mutations, 18 on enzyme surface
Phosphotriesterase 25 996 7, only 1 at active site
Prositagliptin ketone transaminase ∞ (no starting activity) 926 27 mutations, 17 binding substrate. 200 g L−1, >99.5 ee
Triose phosphate isomerase >10[thin space (1/6-em)]000 386 36 mutations, only 1 at active site (NB effects on dimerisation, also implying distant effects)
Valine aminotransferase 21[thin space (1/6-em)]000[thin space (1/6-em)]000 997 17 mutations, only 1 at active site



image file: c4cs00351a-f17.tif
Fig. 17 The residues that influence kcat tend to be distributed throughout an enzyme. The amino acid side chains of each of the 24 mutations obtained993 by the directed evolution of a Dielsalderase (PDB: 4O5T) are highlighted. The active site pocket is shown in grey, while all mutated residues within 5 Å of the ligand (blue) are differentiated from those more than 5 Å away (red). This illustrates that the majority of mutations influencing kcat are not in close proximity to the active (substrate-binding) site. The figure was prepared using PyMol.

What if we lack the structure? Folding up proteins ‘from scratch’

A very particular kind of dynamics is that which leads a protein to fold into its tertiary structure in the first place, and the purpose of this brief section is to draw attention to some recent advances that might allow us to do this computationally. Advances in specialist computer hardware (albeit not yet widely available) can now make a prediction of how a given (smallish) protein sequence will fold up ‘ab initio’.998–1002 However, there are many more protein sequences than there are structures, and this gap is destined to become considerably wider1003 as sequencing methods continue to increase in speed.102 The need for methods that can fold up proteins accurately ‘de novo’ (from their sequences) is thus acute.1004 However, despite a number of advances (e.g.ref. 1000, 1005 and 1006) this is not yet routine. The problem is, of course, that the search space of possible structures is enormous,1 and largely unconstrained. As well as using more powerful hardware, the real key is finding suitable constraints. An important recognition is that the covariance of residues in a series of homologous functional proteins provides a massive constraint on the inter-residue contacts and thus what structures they might adopt, and substantial advances have recently been made by a number of groups197,199,1007–1011 in this regard. Directed evolution supplies an obvious means of creating and assessing suitable sequences.

Metals

As mentioned above, many proteins use metals and cofactors to aid the chemistry that they can catalyse, and while we shall not discuss cofactors, a short section on metalloenzymes is warranted, not least since nearly half of natural enzymes contain metals,1012 albeit that free metals can be quite toxic.1013–1015

To this end, if one wishes to keep open the possibility of incorporating metals into proteins undergoing DE (sometimes referred to as hybrid enzymes1016–1019), it is necessary to understand the common mechanisms, residues and structures involved.460,461,1020–1042

Some specific and unusual examples include high-valent metal catalysis,1043 multi-metal designs as in a di-iron hydrylation reaction,1044 a protein whose fluorescence is metal-dependent1045 and various chelators, quantum dots and so on1046–1050 and metallo-enzymes based on (strept)avidin–biotin technology.1051–1053

A particular attraction of DE is that it becomes possible to incorporate metal ions that are rarely (or never) used in living organisms, to provide novel functions. Examples include iridium,1054 rhodium1055 and uranium (uranyl).1056,1057

Enzyme stability, including thermostability

In general, the rates of chemical reactions increase with temperature, and if we evolve kcat to high levels we may create processes in which temperature may rise naturally anyway (and some processes may simply require it1058). In a similar vein, protein stability tends to decrease with increasing temperature, and there is commonly1059–1061 (though not always1062) a trade-off between kcat and thermostability, including at the cellular level.1063 This relationship depends effectively on the evolutionary pathway followed.1062 As discussed above, thermostability may also sometimes (but not always171–173) correlate with evolvability,175 and is the result of multiple mutations each contributing a small amount.1064–1069

Of course the ‘first law’ of directed evolution is that you get what you select for (even if you did not mean to). Thus if thermostability is important one must incorporate it into one's selection regime, typically by screening for it.1070,1071 Of course if one uses a thermophile such as T. thermophilus then in vivo selection is possible, too.1072

As rehearsed above, protein flexibility (a somewhat ill-defined concept87,1073) is related to kcat, and most residues involved in improving kcat are away from the active site, at the protein surface (where they are bombarded by solvent thermal fluctuations). The connection between flexibility and thermostability is not well understood, and it does not always follow that less flexibility provides greater stability.1074,1075 However, one might suppose that some residues that contribute flexibility are most important for (i.e. contribute significantly to) thermostability too. This is indeed the case.989,1076,1077 Indeed, the same blend of design and focused (thus semi-empirical) DE that has proved valuable for improving kcat values seems to be the best strategy for enhancing thermostability too.1078–1080

Some aspects of thermostability1081–1087 can be related to individual amino acids (e.g. an ionic or H-bond formed by an arginine is of greater strength than that formed by a corresponding lysine, or thermophilic enzymes have more charged and hydrophobic but fewer polar residues1088,1089). However, some aspects are best based on analyses of the 3D structure,456,1090,1091e.g. intra-helix ion pairs1092,1093 and packing density.1068,1094 Thus, Greaves and Warwicker1095 conclude that “charge number relates to solubility, whereas protein stability is determined by charge location”. The choice of which residues to focus on can be assisted (if a structure is available) by looking at the local flexibility via methods such as mutability1096,1097 or via B-factors,1062,1098,1099 or via certain kinds of mass spectrometry.1100–1108 Constraint Network Analysis1109 provides a useful strategy for choosing which residues might be most important for thermostability. Unnatural amino acids may be beneficial too; thus fluoro-aminoacids can increase stability.1110–1112

To disentangle the various contributions to kcat and thermostability, what we need are detailed studies of sequences and structures as they relate to both of these, and published ones remain largely lacking. However, the goal of finding sequence changes that improve both kcat and thermostability is exceptionally desirable. It should also be attainable, on the grounds that protein structural constraints that increase the rate of desirable conformational fluctuations while minimizing those that do not help the enzyme to its catalytically active confirmations must exist and will tend to have this precise effect.

Finally, thermal stresses are not the only stresses that may pertain during a biocatalytic process, albeit sometimes the same mutations can be beneficial in both (e.g. in permitting resistance to oxidation1113,1114 or catalysis in organic solvents1115).

Solvency

While our focus is on evolving proteins, those that are catalyzing reactions are always immersed in a solvent, and we cannot ignore this completely. Although ‘bulk’ measurements of solvent properties are typically unsuitable for molecular analyses of transport across membranes,269,335,356,362,1116–1119 it is the case that some of the binding energy used in enzyme catalysis is effectively used in transferring a substrate from a usually hydrophilic aqueous phase to a usually more ‘hydrophobic’ protein phase. In general, the increased mass/hydrophobicity is also accompanied by a changed value for Km.916 This can lead to some interesting effects of organic solvents, and solvent mixtures,1120 on the specificity,1121–1126 equilibria1127 and catalytic rate constants1128,1129 of enzymes, for reasons that are still not entirely understood. However, because the intention of many DE programs is the production of enzymes for use in industrial processes, the ability to function in organic solvents is often another important objective function, and can be solved via the above strategies.577 One recent trend of note is the exploitation of ionic liquids1130,1131 and ‘deep eutectic solvents’1132–1135 in biocatalysis.

Reaction classes

Apart from circumstances involving extremely reactive substrates and products, there is no known reason of principle why one might not be able to evolve a biocatalyst for any more-or-less simple (i.e. one-step, mono- or bi-molecular) chemical reaction. Thus, one's imagination is limited only by the reactions chosen (nowadays, for a more complex pathway, via retrosynthetic and related strategies (ref. 1136–1148)). Given that these are practically limitless (even if one might wish to start with ‘core’ molecules1149,1150), we choose to be illustrative, and thereby provide a table of some of the kinds of reaction, reaction class or products for which the methods of DE have been used, with a slight focus on non-mainstream reactions. (Curiously, a convenient online database for these is presently lacking.) Our main point is that there seems no obvious limitation on reactions, beyond the case of very highly reactive substrates, intermediates or products, for which an enzymatic reaction cannot be evolved. Since the search space of possible enzymes can never be tested exhaustively, it is a safe prediction that we should expect this to hold for many more, and more complex, chemistries than have been tried to date, provided that the thermodynamics are favourable.

While the focus of this review is about how best to navigate the very large search spaces that pertain in directed enzyme evolution, we recognize that a number of processes including enzymes evolved by DE are now operated industrially.327,328 Examples include sitagliptin,926 generic chiral amine APIs,1151 bio-isoprene,1152 and atorvastatin.1153

Concluding remarks and future prospects

In our review above, we have developed the idea that the most appropriate strategy for improving biocatalysts involves a judicious interplay between design and empiricism, the former more focused at the active site that determines binding and specificity, while the latter might usefully be focussed more on other surface and non-active-site residues to improve kcat and (in part) (thermo)stability. As our knowledge improves, design may begin to play a larger role in optimising kcat, but we consider that this will still require a considerable improvement in our understanding of the relationships between enzyme sequence, structure and dynamics. Thus, protein improvement is likely to involve the creation of increasingly ‘smart’ variant libraries over larger parts of the protein.

Another such interplay relates to the combination of experimental (‘wet’) and computational (‘dry’) approaches. We detect a significant trend towards more of the latter,519 for instance in the use368,1235,1296,1297 of molecular dynamics to calculate properties that suggest which residues might be creating internal friction1298,1299 and hence lowering kcat. These examples help to illustrate that predictions and simulations in silico are likely to play an increasingly important role in predicting strategies for mutagenesis in vitro.

The increasing availability of genomic and metagenomic data, coupled to improvements in the design and prediction of protein structures (and maybe activities) will certainly contribute to improving the initialisation steps of DE. The availability of large sets of protein homologues and analogues will lead to a greater understanding of the relationships (Fig. 1) between protein sequence, structure, dynamics and catalytic activities, all of which can contribute to the design of DE experiments. Together with the development of improved synthetic biology methodology for DNA synthesis and variation, the tools for designing and initialising DE experiments are increasing greatly.

Specifically, the availability of large numbers of sequence-activity pairs may be used to learn to predict where mutations might best be tested. This decreases the empiricism of entirely random mutations in favour of synthetic biology strategies in which one has (at least statistically) more or less complete control over which sequences to make and test. Thus we see a considerable role for modern versions of sequence-activity mapping based on appropriate machine learning methods as a means of predicting where searches might optimally be conducted; this can be done in silico before creating the sequences themselves.23 No doubt many useful datasets of this type exist in the databases of commercial organisations, but they need to become public as the likelihood is that crowdsourcing analyses would add value for their originators1300 as well as for the common good.1301

In terms of optimisation algorithms, we have already pointed out that very few of the modern algorithms of evolutionary optimisation have been applied to the DE problem,107 and the advent of synthetic biology now makes their development and comparison (given that no one size will fit all1302–1306) a worthwhile and timely endeavour. Complex DE algorithms that have no real counterpart in natural evolution can also now be carried out using the methods of synthetic biology.

Searching our empirical knowledge of reactions is becoming increasingly straightforward as it becomes digitised. As implied above, we expect to see an increasing cross-fertilisation between the fields of bioinformatics and cheminformatics1307,1308 and text mining;1309–1311 a very interesting development in this direction is that of Cadeddu et al.1136

Conspicuous by their absence in Table 3 are the members of one important set of reactions that are widely ignored (because they do not always involve actual chemical transformations). These are the transmembrane transporters, and they make up fully one third of the reactions in the reconstructed yeast1312 and human25,1313 metabolic networks. Despite a widespread and longstanding assumption (e.g.ref. 1314) that xenobiotics simply tend to ‘float’ across biological membranes according to their lipophilicity, it is here worth highlighting the considerable literature (that we have reviewed elsewhere, e.g.ref. 269, 335, 356, 357, 362, 1116–1119 and 1315), including a couple of experimental examples (ref. 749 and 750), that implies that the diffusion of xenobiotics through phospholipid bilayers in intact cells is normally negligible. It is now clear that transporters enhance (and are probably required for) the transmembrane transport even of hydrophobic molecules such as alkanes,1316–1321 terpenoids,1319,1322,1323 long-chain,1324–1328 and short-chain1329–1332 fatty acids, and even CO2.1333,1334 This may imply a significantly enhanced role for transporter engineering in whole cell biocatalysis.

Table 3 Some reactions, reaction classes or product types for which DE has proved successful. We largely exclude the very well-established programmes such as ketone and other stereoselective reductases, which along with various other reactions aimed at pharmaceutical intermediates have recently been reviewed in e.g.ref. 326, 328 and 1154–1162. Chiralities are implicit
Reaction (class) or substrate/product Illustrative ref.
Aldolases e.g. R1CHO + R2C([double bond, length as m-dash]O)R3 ⇌ R1C([double bond, length as m-dash]O)CH2C(O)R3 462, 1163 and 1164
Alkenyl and arylmalonate decarboxylases e.g. HOOCC(R1R2)COOH → HC(R1R2)COOH 1165
Amines 1166–1169
Amine dehydrogenase RC([double bond, length as m-dash]O)Me + NH3 + NADH + H+ → RCHNH2Me + H2O + NAD+ 1167
Antifreeze proteins 1170
Azidation RH → RN3 1171 and 1172
Baeyer–Villiger monooxygenases image file: c4cs00351a-u1.tif 1173 and 1174
Beta-keto adipate HOOCCH2CH2C([double bond, length as m-dash]O)CH2COOH 1175
Carotenoid biosynthesis 1176
Chlorinase Ar–H → Ar–Cl 1177–1180
Chloroperoxidase RH + Cl + H2O2 → RCl + H2O + OH 1181–1187
CO groups 1157
Cytochromes P450 e.g. R–H → R–OH 56, 366, 632, 633, 674, 692, 992 and 1188–1208
Diels–Alderases e.g. CH2[double bond, length as m-dash]CHCH[double bond, length as m-dash]CH2 + CH2[double bond, length as m-dash]CH2 → cyclohexene 378, 380, 993, 1209 and 1210
DNA polymerase 1211 and 1212
Endopeptidases 769
Esterase enantioselectivity 1213
Epoxide hydrolase image file: c4cs00351a-u2.tif 947
Flavanones image file: c4cs00351a-u3.tif 1214–1221
Fluorinase 1178 and 1222 (and see ref. 1223)
Fatty acids 1224–1226
Glyphosate acyltransferase HOOCCH2NHCH2PO32− + AcCoA → HOOCCH2N(CH3C[double bond, length as m-dash]O)CH2PO32− + CoA 995 and 1227–1229
Glycine (glyphosate) oxidase HOOCCH2NHCH2PO32− + O2 → OHC–COOH + H2NCH2 PO32− + H2O2 1229–1231
Haloalkane dehalogenase R1C(HBr)R2 + H2O → R1C(HOH)R2 + H+ + Br 1232–1235
Halogenase Ar–H → Ar–Hal 1236–1239
Hydroxytyrosol 2–NO2–Ph–CH2CH2OH + O2 → 2–OH, 3–OH–Ph–CH2CH2OH + NO2 1240
Kemp eliminase 377 and 1241–1247
Ketone reductions R1–C([double bond, length as m-dash]O)–R2 → R1–CH(OH)–R2 1248 and 1249
Laccase 1250–1252
Michael addition R–CH2CHO + Ph–CH[double bond, length as m-dash]CHNO2 → OHC–CH(R)–CH(Ph)CH2NO2 1253
Monoamine oxidase R1R2CHCH2NR3R4 + 1/2O2 →R1R2C[double bond, length as m-dash]NR3 (R4 = H) or R1R2CH[double bond, length as m-dash]N+R3R4 (R4 = alkyl) + H2O 728, 1151 and 1254–1256
Nitrogenase N2 + 3H2 → 2NH3 1257
Nucleases 1258
Old yellow enzyme (activated alkene reductions) 664, 1259 and 1260
Paraoxonase R1(R2O)(R3O)–P[double bond, length as m-dash]O → R1(R2O)(HO)P[double bond, length as m-dash]O + R3OH 1261–1263
Peroxidase 1264 and 1265
Phospho(mono/di/tri)esterases 255, 831 and 1266–1271
Polyketides 1272–1275
Polylactate 1276
Redox enzymes 1277–1279
Reductive cyclisation 1280
Restriction protease 1281
Retro-aldolase e.g. R1C([double bond, length as m-dash]O)CH2C(O)R3 ⇌ R1CHO + R2C([double bond, length as m-dash]O)R3 463 and 1282
Sesquiterpene synthases 1283
Tautomerases Ar–CH[double bond, length as m-dash]C(OH)COOH ⇌ Ar–CH2COCOOH 1284
Terpene synthase/cyclase 1285–1287
Transaldolase erythrose-4-phosphate + fructose-6-phosphate → glyceraldehyde-3-phosphate + sedoheptulose-7-phosphate 1288 and 1289
Transketolase RCHO + HOCH2COCOOH → RCH(OH)COCH2OH 1290–1293
Zinc finger proteins 1294 and 1295


The recent introduction of the community standard Synthetic Biology Open Language (SBOL) will certainly facilitate the sharing and re-use of synthetic biology designed sequences and modules. Beginning in 2008, the development of SBOL has been driven by an international community of computational synthetic biologists, and has led to the introduction of an initial standard for the sharing of synthetic DNA sequences1335 and also for their visualisation. A recent proposal has introduced a more complete extension to the language, covering interactions between synthetic sequences, the design of modules and specification of their overall function.1336 Just as with the Systems Biology Markup Language,1337 the Systems Biology Graphical Notation,1338 and related controlled vocabularies, metadata and ontologies for knowledge exchange in systems biology1339,1340 and metabolomics,1341 the availability of these kinds of standards will help move the field forward considerably.

Overall, we conclude that existing and emerging knowledge-based methods exploiting the strategies and capabilities of synthetic biology and the power of e-science will be a huge driver for the improvement of biocatalysts by directed evolution. We have only just begun.

Acknowledgements

We thank Chris Knight, Rainer Breitling, Nigel Scrutton and Nick Turner for very useful comments on the manuscript, Colin Levy for some material for the figures, and the Biotechnology and Biological Sciences Research Council for financial support (grant BB/M017702/1). This is a contribution from the Manchester Centre for Synthetic Biology of Fine and Speciality Chemicals (SYNBIOCHEM).

References

  1. D. B. Kell, Scientific discovery as a combinatorial optimisation problem: how best to navigate the landscape of possible experiments?, BioEssays, 2012, 34, 236–244 CrossRef PubMed.
  2. Phylogenetic analysis of DNA sequences, ed. M. M. Miyamoto and J. Cracraft, Oxford University Press, Oxford, 1991 Search PubMed.
  3. R. D. M. Page and E. C. Holmes, Molecular evolution: a phylogenetic approach, Blackwell Science, Oxford, 1998 Search PubMed.
  4. M. J. Harms and J. W. Thornton, Evolutionary biochemistry: revealing the historical and physical causes of protein properties, Nat. Rev. Genet., 2013, 14, 559–571 CrossRef CAS PubMed.
  5. D. G. Gibson, G. A. Benders, K. C. Axelrod, J. Zaveri, M. A. Algire, M. Moodie, M. G. Montague and J. C. Venter, H. O. Smith and C. A. Hutchison, 3rd, One-step assembly in yeast of 25 overlapping DNA fragments to form a complete synthetic Mycoplasma genitalium genome, Proc. Natl. Acad. Sci. U. S. A., 2008, 105, 20404–20409 CrossRef CAS PubMed.
  6. D. G. Gibson, G. A. Benders, C. Andrews-Pfannkoch, E. A. Denisova, H. Baden-Tillson, J. Zaveri, T. B. Stockwell, A. Brownley, D. W. Thomas, M. A. Algire, C. Merryman, L. Young, V. N. Noskov, J. I. Glass, J. C. Venter, C. A. Hutchison 3rd and H. O. Smith, Complete chemical synthesis, assembly, and cloning of a Mycoplasma genitalium genome, Science, 2008, 319, 1215–1220 CrossRef CAS PubMed.
  7. H. J. Frasch, M. H. Medema, E. Takano and R. Breitling, Design-based re-engineering of biosynthetic gene clusters: plug-and-play in practice, Curr. Opin. Biotechnol., 2013, 24, 1144–1150 CrossRef CAS PubMed.
  8. S. E. Ongley, X. Bian, B. A. Neilan and R. Muller, Recent advances in the heterologous expression of microbial natural product biosynthetic pathways, Nat. Prod. Rep., 2013, 30, 1121–1138 RSC.
  9. A. Pourmir and T. W. Johannes, Directed evolution: selection of the host organism, Comput. Struct. Biotechnol. J., 2012, 2, e201209012 Search PubMed.
  10. S. V. Taylor, K. U. Walter, P. Kast and D. Hilvert, Searching sequence space for protein catalysts, Proc. Natl. Acad. Sci. U. S. A., 2001, 98, 10596–10601 CrossRef CAS PubMed.
  11. K. J. Waldron, J. C. Rutherford, D. Ford and N. J. Robinson, Metalloproteins and metal sensing, Nature, 2009, 460, 823–830 CrossRef CAS PubMed.
  12. C. J. Jeffery, Moonlighting proteins—an update, Mol. BioSyst., 2009, 5, 345–350 RSC.
  13. J. C. Moore, H. M. Jin, O. Kuchner and F. H. Arnold, Strategies for the in vitro evolution of protein function: Enzyme evolution by random recombination of improved sequences, J. Mol. Biol., 1997, 272, 336–347 CrossRef CAS PubMed.
  14. R. Bellman, Adaptive control processes: a guided tour, Princeton University Press, Princeton, NJ, 1961 Search PubMed.
  15. C. D. Manning, P. Raghavan and H. Schütze, Introduction to information retrieval, CUP, Cambridge, 2009 Search PubMed.
  16. T. Hastie, R. Tibshirani and J. Friedman, The elements of statistical learning: data mining, inference and prediction, Springer-Verlag, Berlin, 2001 Search PubMed.
  17. F. H. Arnold, Fancy footwork in the sequence space shuffle, Nat. Biotechnol., 2006, 24, 328–330 CrossRef CAS PubMed.
  18. W. S. Cleveland, Visualizing data, Hobart Press, Summit, NJ, 1993 Search PubMed.
  19. Beautiful visualization: looking at data through the eyes of experts, ed. J. Steele and N. Iliinsky, O'Reilly, Sebastopol, CA, 2010 Search PubMed.
  20. C. Ware, Information visualization, Morgan Kaufmann, San Francisco, 2000 Search PubMed.
  21. D. M. McCandlish, Visualizing fitness landscapes, Evolution, 2011, 65, 1544–1558 CrossRef PubMed.
  22. N. Yau, Visualize this: the FlowingData guide to design, visualization and statistics, Wiley, Indianapolis, IN, 2011 Search PubMed.
  23. C. G. Knight, M. Platt, W. Rowe, D. C. Wedge, F. Khan, P. Day, A. McShea, J. Knowles and D. B. Kell, Array-based evolution of DNA aptamers allows modelling of an explicit sequence-fitness landscape, Nucleic Acids Res., 2009, 37, e6 CrossRef PubMed.
  24. S. Sahoo, L. Franzson, J. J. Jonsson and I. Thiele, A compendium of inborn errors of metabolism mapped onto the human metabolic network, Mol. BioSyst., 2012, 8, 2545–2558 RSC.
  25. I. Thiele, N. Swainston, R. M. T. Fleming, A. Hoppe, S. Sahoo, M. K. Aurich, H. Haraldsdottír, M. L. Mo, O. Rolfsson, M. D. Stobbe, S. G. Thorleifsson, R. Agren, C. Bölling, S. Bordel, A. K. Chavali, P. Dobson, W. B. Dunn, L. Endler, I. Goryanin, D. Hala, M. Hucka, D. Hull, D. Jameson, N. Jamshidi, J. Jones, J. J. Jonsson, N. Juty, S. Keating, I. Nookaew, N. Le Novère, N. Malys, A. Mazein, J. A. Papin, Y. Patel, N. D. Price, E. Selkov Sr, M. I. Sigurdsson, E. Simeonidis, N. Sonnenschein, K. Smallbone, A. Sorokin, H. V. Beek, D. Weichart, J. B. Nielsen, H. V. Westerhoff, D. B. Kell, P. Mendes and B. Ø. Palsson, A community-driven global reconstruction of human metabolism, Nat. Biotechnol., 2013, 31, 419–425 CrossRef CAS PubMed.
  26. D. Dimova, M. Wawer, A. M. Wassermann and J. Bajorath, Design of multitarget activity landscapes that capture hierarchical activity cliff distributions, J. Chem. Inf. Model., 2011, 51, 258–266 CrossRef CAS PubMed.
  27. G. M. Maggiora, On outliers and activity cliffs—why QSAR often disappoints, J. Chem. Inf. Model., 2006, 46, 1535 CrossRef CAS PubMed.
  28. R. Guha and J. H. Van Drie, Structure–activity landscape index: identifying and quantifying activity cliffs, J. Chem. Inf. Model., 2008, 48, 646–658 CrossRef CAS PubMed.
  29. J. L. Medina-Franco, Activity cliffs: facts or artifacts?, Chem. Biol. Drug Des., 2013, 81, 553–556 CAS.
  30. D. Stumpfe and J. Bajorath, Frequency of occurrence and potency range distribution of activity cliffs in bioactive compounds, J. Chem. Inf. Model., 2012, 52, 2348–2353 CrossRef CAS PubMed.
  31. D. Stumpfe and J. Bajorath, Exploring activity cliffs in medicinal chemistry, J. Med. Chem., 2012, 55, 2932–2942 CrossRef CAS PubMed.
  32. D. Stumpfe, Y. Hu, D. Dimova and J. Bajorath, Recent progress in understanding activity cliffs and their utility in medicinal chemistry, J. Med. Chem., 2014, 57, 18–28 CrossRef CAS PubMed.
  33. R. Guha and J. L. Medina-Franco, On the validity versus utility of activity landscapes: are all activity cliffs statistically significant?, J. Cheminf., 2014, 6, 11 Search PubMed.
  34. R. Guha, What makes a good structure activity landscape? Network metrics and structure representations as a way of exploring activity landscapes, Abstracts of Papers of the American Chemical Society, 2010, 240 Search PubMed.
  35. R. Guha, The ups and downs of structure-activity landscapes, Methods Mol. Biol., 2011, 672, 101–117 CAS.
  36. R. Guha, Exploring uncharted territories: predicting activity cliffs in structure-activity landscapes, J. Chem. Inf. Model., 2012, 52, 2181–2191 CrossRef CAS PubMed.
  37. R. Guha, Exploring Structure-Activity Data Using the Landscape Paradigm, Wiley Interdiscip. Rev.: Comput. Mol. Sci., 2012, 2, 829–841,  DOI:10.1002/wcms.1087.
  38. S. T. Rao and M. G. Rossmann, Comparison of super-secondary structures in proteins, J. Mol. Biol., 1973, 76, 241–256 CrossRef CAS.
  39. M. Saraste, P. R. Sibbald and A. Wittinghofer, The P-loop: a common motif in ATP- and GTP-binding proteins, Trends Biochem. Sci., 1990, 15, 430–434 CrossRef.
  40. P. D. Dobson and A. J. Doig, Distinguishing enzyme structures from non-enzymes without alignments, J. Mol. Biol., 2003, 330, 771–783 CrossRef CAS.
  41. P. D. Dobson and A. J. Doig, Predicting enzyme class from protein structure without alignments, J. Mol. Biol., 2005, 345, 187–199 CrossRef CAS PubMed.
  42. J. Minshull, J. E. Ness, C. Gustafsson and S. Govindarajan, Predicting enzyme function from protein sequence, Curr. Opin. Chem. Biol., 2005, 9, 202–209 CrossRef CAS PubMed.
  43. L. Han, J. Cui, H. Lin, Z. Ji, Z. Cao, Y. Li and Y. Chen, Recent progresses in the application of machine learning approach for predicting protein functional class independent of sequence similarity, Proteomics, 2006, 6, 4023–4037 CrossRef CAS PubMed.
  44. Z. Q. Tang, H. H. Lin, H. L. Zhang, L. Y. Han, X. Chen and Y. Z. Chen, Prediction of functional class of proteins and peptides irrespective of sequence homology by support vector machines, Bioinf. Biol. Insights, 2007, 1, 19–47 Search PubMed.
  45. J. L. Faulon, M. Misra, S. Martin, K. Sale and R. Sapra, Genome scale enzyme-metabolite and drug-target interaction predictions using the signature molecular descriptor, Bioinformatics, 2008, 24, 225–233 CrossRef CAS PubMed.
  46. H. Strömbergsson, P. Daniluk, A. Kryshtafovych, K. Fidelis, J. E. Wikberg, G. J. Kleywegt and T. R. Hvidsten, Interaction model based on local protein substructures generalizes to the entire structural enzyme-ligand space, J. Chem. Inf. Model., 2008, 48, 2278–2288 CrossRef PubMed.
  47. T. Bray, P. Chan, S. Bougouffa, R. Greaves, A. J. Doig and J. Warwicker, Sites Identify: a protein functional site prediction tool, BMC Bioinf., 2009, 10, 379 CrossRef PubMed.
  48. T. R. Hvidsten, A. Lægreid, A. Kryshtafovych, G. Andersson, K. Fidelis and J. Komorowski, A comprehensive analysis of the structure-function relationship in proteins based on local structure similarity, PLoS One, 2009, 4, e6266 Search PubMed.
  49. D. M. Fowler, C. L. Araya, S. J. Fleishman, E. H. Kellogg, J. J. Stephany, D. Baker and S. Fields, High-resolution mapping of protein sequence-function relationships, Nat. Methods, 2010, 7, 741–746 CrossRef CAS PubMed.
  50. T. Lee, H. Min, S. J. Kim and S. Yoon, Application of maximin correlation analysis to classifying protein environments for function prediction, Biochem. Biophys. Res. Commun., 2010, 400, 219–224 CrossRef CAS PubMed.
  51. C. R. Shyu, B. Pang, P. H. Chi, N. Zhao, D. Korkin and D. Xu, ProteinDBS v2.0: a web server for global and local protein structure search, Nucleic Acids Res., 2010, 38, W53–W58 CrossRef CAS PubMed.
  52. S. D. Brown and P. C. Babbitt, Inference of functional properties from large-scale analysis of enzyme superfamilies, J. Biol. Chem., 2012, 287, 35–42 CrossRef CAS PubMed.
  53. L. De Ferrari, S. Aitken, J. van Hemert and I. Goryanin, EnzML: multi-label prediction of enzyme classes using InterPro signatures, BMC Bioinf., 2012, 13, 61 CrossRef CAS PubMed.
  54. Y. Qi, M. Oja, J. Weston and W. S. Noble, A unified multitask architecture for predicting local protein properties, PLoS One, 2012, 7, e32235 CAS.
  55. S. Wright, The roles of mutation, inbreeding, crossbreeding and selection in evolution, in Proc. Sixth Int. Conf. Genetics, ed. D. F. Jones, Genetics Society of America, Austin TX, Ithaca, NY, 1932, pp. 356–366 Search PubMed.
  56. P. A. Romero and F. H. Arnold, Exploring protein fitness landscapes by directed evolution, Nat. Rev. Mol. Cell Biol., 2009, 10, 866–876 CrossRef CAS PubMed.
  57. J. A. G. M. de Visser and J. Krug, Empirical fitness landscapes and the predictability of evolution, Nat. Rev. Genet., 2014, 15, 480–490 CrossRef CAS PubMed.
  58. J. Maynard Smith, Natural selection and the concept of a protein space, Nature, 1970, 225, 563–564 CrossRef.
  59. A. R. Davidson, K. J. Lumb and R. T. Sauer, Cooperatively folded proteins in random sequence libraries, Nat. Struct. Biol., 1995, 2, 856–864 CrossRef CAS PubMed.
  60. J. R. Beasley and M. H. Hecht, Protein design: the choice of de novo sequences, J. Biol. Chem., 1997, 272, 2031–2034 CrossRef CAS PubMed.
  61. T. Matsuura, A. Ernst and A. Plückthun, Construction and characterization of protein libraries composed of secondary structure modules, Protein Sci., 2002, 11, 2631–2643 CrossRef CAS PubMed.
  62. T. Matsuura and A. Plückthun, Strategies for selection from protein libraries composed of de novo designed secondary structure modules, Origins Life Evol. Biospheres, 2004, 34, 151–157 CrossRef CAS.
  63. D. D. Axe, Estimating the prevalence of protein sequences adopting functional enzyme folds, J. Mol. Biol., 2004, 341, 1295–1315 CrossRef CAS PubMed.
  64. L. H. Bradley, P. P. Thumfort and M. H. Hecht, De novo proteins from binary-patterned combinatorial libraries, Methods Mol. Biol., 2006, 340, 53–69 CAS.
  65. J. J. Graziano, W. Liu, R. Perera, B. H. Geierstanger, S. A. Lesley and P. G. Schultz, Selecting folded proteins from a library of secondary structural elements, J. Am. Chem. Soc., 2008, 130, 176–185 CrossRef CAS PubMed.
  66. T. Schmidt-Goenner, A. Guerler, B. Kolbeck and E. W. Knapp, Circular permuted proteins in the universe of protein folds, Proteins, 2010, 78, 1618–1630 CrossRef CAS PubMed.
  67. J. Tanaka, N. Doi, H. Takashima and H. Yanagawa, Comparative characterization of random-sequence proteins consisting of 5, 12, and 20 kinds of amino acids, Protein Sci., 2010, 19, 786–795 CrossRef CAS PubMed.
  68. M. H. Hecht, M. W. West, J. Patterson, J. D. Mancias, J. R. Beasley, B. M. Broome and W. Wang, Designed combinatorial libraries of novel amyloid-like proteins, Self-Assem. Pept. Syst. Biol., Med. Eng., [Workshop], 2001, 127–138 CAS.
  69. Y. Ito, T. Kawama, I. Urabe and T. Yomo, Evolution of an arbitrary sequence in solubility, J. Mol. Evol., 2004, 58, 196–202 CrossRef CAS PubMed.
  70. A. D. Keefe and J. W. Szostak, Functional proteins from a random-sequence library, Nature, 2001, 410, 715–718 CrossRef CAS PubMed.
  71. B. Kuhlman and D. Baker, Native protein sequences are close to optimal for their structures, Proc. Natl. Acad. Sci. U. S. A., 2000, 97, 10383–10388 CrossRef CAS.
  72. H. P. Yockey, Information Theory and Molecular Biology, Cambridge University Press, Cambridge, 1992 Search PubMed.
  73. Y. Wei and M. H. Hecht, Enzyme-like proteins from an unselected library of designed amino acid sequences, Protein Eng., Des. Sel., 2004, 17, 67–75 CrossRef CAS PubMed.
  74. G. Minervini, G. Evangelista, L. Villanova, D. Slanzi, D. De Lucrezia, I. Poli, P. L. Luisi and F. Polticelli, Massive non-natural proteins structure prediction using grid technologies, BMC Bioinf., 2009, 10(suppl 6), S22 CrossRef PubMed.
  75. K. Prymula, M. Piwowar, M. Kochanczyk, L. Flis, M. Malawski, T. Szepieniec, G. Evangelista, G. Minervini, F. Polticelli, Z. Wisniowski, K. Salapa, E. Matczynska and I. Roterman, In silico structural study of random amino acid sequence proteins not present in nature, Chem. Biodiversity, 2009, 6, 2311–2336 CAS.
  76. M. Scalley-Kim and D. Baker, Characterization of the folding energy landscapes of computer generated proteins suggests high folding free energy barriers and cooperativity may be consequences of natural selection, J. Mol. Biol., 2004, 338, 573–583 CrossRef CAS PubMed.
  77. S. Kamtekar, J. M. Schiffer, H. Xiong, J. M. Babik and M. H. Hecht, Protein design by binary patterning of polar and nonpolar amino acids, Science, 1993, 262, 1680–1685 CAS.
  78. M. W. West and M. H. Hecht, Binary patterning of polar and nonpolar amino acids in the sequences and structures of native proteins, Protein Sci., 1995, 4, 2032–2039 CrossRef CAS PubMed.
  79. H. Xiong, B. L. Buckwalter, H. M. Shieh and M. H. Hecht, Periodicity of polar and nonpolar amino acids is the major determinant of secondary structure in self-assembling oligomeric peptides, Proc. Natl. Acad. Sci. U. S. A., 1995, 92, 6349–6353 CrossRef CAS.
  80. D. A. Moffet, J. Foley and M. H. Hecht, Midpoint reduction potentials and heme binding stoichiometries of de novo proteins from designed combinatorial libraries, Biophys. Chem., 2003, 105, 231–239 CrossRef CAS.
  81. M. H. Hecht, A. Das, A. Go, L. H. Bradley and Y. Wei, De novo proteins from designed combinatorial libraries, Protein Sci., 2004, 13, 1711–1723 CrossRef CAS PubMed.
  82. L. H. Bradley, Y. Wei, P. Thumfort, C. Wurth and M. H. Hecht, Protein design by binary patterning of polar and nonpolar amino acids, Methods Mol. Biol., 2007, 352, 155–166 CAS.
  83. S. C. Patel, L. H. Bradley, S. P. Jinadasa and M. H. Hecht, Cofactor binding and enzymatic activity in an unevolved superfamily of de novo designed 4-helix bundle proteins, Protein Sci., 2009, 18, 1388–1400 CrossRef CAS PubMed.
  84. M. A. Fisher, K. L. McKinley, L. H. Bradley, S. R. Viola and M. H. Hecht, De novo designed proteins from a library of artificial sequences function in Escherichia coli and enable cell growth, PLoS One, 2011, 6, e15364 CAS.
  85. C. H. Shih, C. M. Chang, Y. S. Lin, W. C. Lo and J. K. Hwang, Evolutionary information hidden in a single protein structure, Proteins, 2012, 80, 1647–1657 CrossRef CAS PubMed.
  86. C. M. Chang, Y. W. Huang, C. H. Shih and J. K. Hwang, On the relationship between the sequence conservation and the packing density profiles of the protein complexes, Proteins, 2013, 81, 1192–1199 CrossRef CAS PubMed.
  87. T. T. Huang, M. L. D. Marcos, J. K. Hwang and J. Echave, A mechanistic stress model of protein evolution accounts for site-specific evolutionary rates and their relationship with packing density and flexibility, BMC Evol. Biol., 2014, 14, 78 CrossRef PubMed.
  88. S. W. Yeh, T. T. Huang, J. W. Liu, S. H. Yu, C. H. Shih, J. K. Hwang and J. Echave, Local Packing Density Is the Main Structural Determinant of the Rate of Protein Sequence Evolution at Site Level, BioMed Res. Int., 2014, 572409 Search PubMed.
  89. S. W. Yeh, J. W. Liu, S. H. Yu, C. H. Shih, J. K. Hwang and J. Echave, Site-Specific Structural Constraints on Protein Sequence Evolutionary Divergence: Local Packing Density versus Solvent Exposure, Mol. Biol. Evol., 2014, 31, 135–139 CrossRef CAS PubMed.
  90. A. Yamauchi, T. Nakashima, N. Tokuriki, M. Hosokawa, H. Nogami, S. Arioka, I. Urabe and T. Yomo, Evolvability of random polypeptides through functional selection within a small library, Protein Eng., 2002, 15, 619–626 CrossRef CAS PubMed.
  91. A. R. Davidson and R. T. Sauer, Folded proteins occur frequently in libraries of random amino acid sequences, Proc. Natl. Acad. Sci. U. S. A., 1994, 91, 2146–2150 CrossRef CAS.
  92. H. H. Guo, J. Choe and L. A. Loeb, Protein tolerance to random amino acid change, Proc. Natl. Acad. Sci. U. S. A., 2004, 101, 9205–9210 CrossRef CAS PubMed.
  93. N. Silver, The signal and the noise: the art and science of prediction, Penguin, New York, 2012 Search PubMed.
  94. F. J. Poelwijk, D. J. Kiviet, D. M. Weinreich and S. J. Tans, Empirical fitness landscapes reveal accessible evolutionary paths, Nature, 2007, 445, 383–386 CrossRef CAS PubMed.
  95. M. J. Keiser, B. L. Roth, B. N. Armbruster, P. Ernsberger, J. J. Irwin and B. K. Shoichet, Relating protein pharmacology by ligand chemistry, Nat. Biotechnol., 2007, 25, 197–206 CrossRef CAS PubMed.
  96. M. Bashton, I. Nobeli and J. M. Thornton, PROCOGNATE: a cognate ligand domain mapping for enzymes, Nucleic Acids Res., 2008, 36, D618–D622 CrossRef CAS PubMed.
  97. R. Adams, C. L. Worth, S. Guenther, M. Dunkel, R. Lehmann and R. Preissner, Binding sites in membrane proteins - diversity, druggability and prospects, Eur. J. Cell Biol., 2012, 91, 326–339 CrossRef CAS PubMed.
  98. A. M. Wassermann and J. Bajorath, BindingDB and ChEMBL: online compound databases for drug discovery, Expert Opin. Drug Discovery, 2011, 6, 683–687 CrossRef CAS PubMed.
  99. D. E. Featherstone and K. Broadie, Wrestling with pleiotropy: genomic and topological analysis of the yeast gene expression network, BioEssays, 2002, 24, 267–274 CrossRef CAS PubMed.
  100. M. Soskine and D. S. Tawfik, Mutational effects and the evolution of new protein functions, Nat. Rev. Genet., 2010, 11, 572–582 CrossRef CAS PubMed.
  101. J. Shendure and H. Ji, Next-generation DNA sequencing, Nat. Biotechnol., 2008, 26, 1135–1145 CrossRef CAS PubMed.
  102. J. Shendure and E. L. Aiden, The expanding scope of DNA sequencing, Nat. Biotechnol., 2012, 30, 1084–1094 CrossRef CAS PubMed.
  103. J. I. Jiménez, R. Xulvi-Brunet, G. W. Campbell, R. Turk-MacLeod and I. A. Chen, Comprehensive experimental fitness landscape and evolutionary network for small RNA, Proc. Natl. Acad. Sci. U. S. A., 2013, 110, 14984–14989 CrossRef PubMed.
  104. W. Rowe, M. Platt, D. Wedge, P. J. Day, D. B. Kell and J. Knowles, Analysis of a complete DNA-protein affinity landscape, J. R. Soc., Interface, 2010, 7, 397–408 CrossRef CAS PubMed.
  105. Á. E. Eiben, R. Hinterding and Z. Michalewicz, Parameter control in evolutionary algorithms, IEEE Trans. Evol. Comput., 1999, 3, 124–141 CrossRef.
  106. Handbook of evolutionary computation., ed. T. Bäck, D. B. Fogel and Z. Michalewicz, IOP Publishing/Oxford University Press, Oxford, 1997 Search PubMed.
  107. S. O'Hagan, J. Knowles and D. B. Kell, Exploiting genomic knowledge in optimising molecular breeding programmes: algorithms from evolutionary computing, PLoS One, 2012, 7, e48862 Search PubMed.
  108. G. Syswerda, Uniform crossover in genetic algorithms, in Proc 3rd Int Conf on Genetic Algorithms, ed. J. Schaffer, Morgan Kaufmann, 1989, pp. 2–9 Search PubMed.
  109. L. He, A. M. Friedman and C. Bailey-Kellogg, A divide-and-conquer approach to determine the Pareto frontier for optimization of protein engineering experiments, Proteins, 2012, 80, 790–806 CrossRef CAS PubMed.
  110. J. Handl, D. B. Kell and J. Knowles, Multiobjective optimization in bioinformatics and computational biology, IEEE/ACM Trans. Comput. Biol. Bioinf., 2007, 4, 279–292 CrossRef CAS PubMed.
  111. J. D. Knowles and D. W. Corne, Approximating the non-dominated front using the Pareto Archived Evolution Strategy, Evol. Comput., 2000, 8, 149–172 CrossRef CAS PubMed.
  112. J. D. Knowles and D. W. Corne, M-PAES: a memetic algorithm for multiobjective optimization, Proc. 2000 Congr. Evol. Computation, 2000, vol. 1 and 2, pp. 325–332 Search PubMed.
  113. K. Deb, Multi-objective optimization using evolutionary algorithms, Wiley, New York, 2001 Search PubMed.
  114. E. Zitzler, L. Thiele, M. Laumanns, C. M. Fonseca and V. G. da Fonseca, Performance assessment of multiobjective optimizers: An analysis and review, IEEE Trans. Evol. Comput., 2003, 7, 117–132 CrossRef.
  115. J. Knowles, ParEGO: A hybrid algorithm with on-line landscape approximation for expensive multiobjective optimization problems, IEEE Trans. Evol. Comput., 2006, 10, 50–66 CrossRef.
  116. Multiobjective Problem Solving from Nature, ed. J. Knowles, D. Corne and K. Deb, Springer, Berlin, 2008 Search PubMed.
  117. B. Maher, The case of the missing heritability, Nature, 2008, 456, 18–21 CrossRef CAS PubMed.
  118. D. M. Weinreich, N. F. Delaney, M. A. Depristo and D. L. Hartl, Darwinian evolution can follow only very few mutational paths to fitter proteins, Science, 2006, 312, 111–114 CrossRef CAS PubMed.
  119. W. Sung, M. S. Ackerman, S. F. Miller, T. G. Doak and M. Lynch, Drift-barrier hypothesis and mutation-rate evolution, Proc. Natl. Acad. Sci. U. S. A., 2012, 109, 18488–18492 CrossRef CAS PubMed.
  120. M. W. Nachman and S. L. Crowell, Estimate of the mutation rate per nucleotide in humans, Genetics, 2000, 156, 297–304 CAS.
  121. P. D. Keightley, Rates and fitness consequences of new mutations in humans, Genetics, 2012, 190, 295–304 CrossRef PubMed.
  122. P. D. Sniegowski, P. J. Gerrish and R. E. Lenski, Evolution of high mutation rates in experimental populations of E. coli, Nature, 1997, 387, 703–705 CrossRef CAS PubMed.
  123. M. V. Rockman and L. Kruglyak, Recombinational landscape and population genomics of Caenorhabditis elegans, PLoS Genet., 2009, 5, e1000419 Search PubMed.
  124. P. D. Sniegowski and R. E. Lenski, Mutation and Adaptation - the Directed Mutation Controversy in Evolutionary Perspective, Annu. Rev. Ecol. Evol. Syst., 1995, 26, 553–578 Search PubMed.
  125. J. R. Peck and D. Waxman, Is life impossible? Information, sex, and the origin of complex organisms, Evolution, 2010, 64, 3300–3309 CrossRef PubMed.
  126. J. Franke, A. Klözer, J. A. G. M. de Visser and J. Krug, Evolutionary accessibility of mutational pathways, PLoS Comput. Biol., 2011, 7, e1002134 CAS.
  127. B. Papp, R. A. Notebaart and C. Pál, Systems-biology approaches for predicting genomic evolution, Nat. Rev. Genet., 2011, 12, 591–602 CrossRef CAS PubMed.
  128. C. A. Orengo and J. M. Thornton, Protein families and their evolution-a structural perspective, Annu. Rev. Biochem., 2005, 74, 867–900 CrossRef CAS PubMed.
  129. A. D. Moore, S. Grath, A. Schuler, A. K. Huylmans and E. Bornberg-Bauer, Quantification and functional analysis of modular protein evolution in a dense phylogenetic tree, Biochim. Biophys. Acta, 2013, 1834, 898–907 CrossRef CAS PubMed.
  130. T. E. Lewis, I. Sillitoe, A. Andreeva, T. L. Blundell, D. W. Buchan, C. Chothia, A. Cuff, J. M. Dana, I. Filippis, J. Gough, S. Hunter, D. T. Jones, L. A. Kelley, G. J. Kleywegt, F. Minneci, A. Mitchell, A. G. Murzin, B. Ochoa-Montano, O. J. Rackham, J. Smith, M. J. Sternberg, S. Velankar, C. Yeats and C. Orengo, Genome3D: a UK collaborative project to annotate genomic sequences with predicted 3D structures based on SCOP and CATH domains, Nucleic Acids Res., 2013, 41, D499–D507 CrossRef CAS PubMed.
  131. D. Xu, Protein databases on the internet, Curr. Protoc. Mol. Biol., 2012, ch. 19, Unit 19.14 Search PubMed.
  132. L. H. Greene, T. E. Lewis, S. Addou, A. Cuff, T. Dallman, M. Dibley, O. Redfern, F. Pearl, R. Nambudiry, A. Reid, I. Sillitoe, C. Yeats, J. M. Thornton and C. A. Orengo, The CATH domain structure database: new protocols and classification levels give a more comprehensive resource for exploring evolution, Nucleic Acids Res., 2007, 35, D291–D297 CrossRef CAS PubMed.
  133. A. L. Cuff, I. Sillitoe, T. Lewis, O. C. Redfern, R. Garratt, J. Thornton and C. A. Orengo, The CATH classification revisited—architectures reviewed and new ways to characterize structural divergence in superfamilies, Nucleic Acids Res., 2009, 37, D310–D314 CrossRef CAS PubMed.
  134. A. L. Cuff, I. Sillitoe, T. Lewis, A. B. Clegg, R. Rentzsch, N. Furnham, M. Pellegrini-Calace, D. Jones, J. Thornton and C. A. Orengo, Extending CATH: increasing coverage of the protein structure universe and linking structure with function, Nucleic Acids Res., 2011, 39, D420–D426 CrossRef CAS PubMed.
  135. A. G. Murzin, S. E. Brenner, T. Hubbard and C. Chothia, SCOP: a structural classification of proteins database for the investigation of sequences and structures, J. Mol. Biol., 1995, 247, 536–540 CAS.
  136. A. Andreeva, D. Howorth, J. M. Chandonia, S. E. Brenner, T. J. Hubbard, C. Chothia and A. G. Murzin, Data growth and its impact on the SCOP database: new developments, Nucleic Acids Res., 2008, 36, D419–D425 CrossRef CAS PubMed.
  137. N. K. Fox, S. E. Brenner and J. M. Chandonia, SCOPe: Structural Classification of Proteins—extended, integrating SCOP and ASTRAL data and classification of new structures, Nucleic Acids Res., 2014, 42, D304–D309 CrossRef CAS PubMed.
  138. S. Hunter, P. Jones, A. Mitchell, R. Apweiler, T. K. Attwood, A. Bateman, T. Bernard, D. Binns, P. Bork, S. Burge, E. de Castro, P. Coggill, M. Corbett, U. Das, L. Daugherty, L. Duquenne, R. D. Finn, M. Fraser, J. Gough, D. Haft, N. Hulo, D. Kahn, E. Kelly, I. Letunic, D. Lonsdale, R. Lopez, M. Madera, J. Maslen, C. McAnulla, J. McDowall, C. McMenamin, H. Mi, P. Mutowo-Muellenet, N. Mulder, D. Natale, C. Orengo, S. Pesseat, M. Punta, A. F. Quinn, C. Rivoire, A. Sangrador-Vegas, J. D. Selengut, C. J. Sigrist, M. Scheremetjew, J. Tate, M. Thimmajanarthanan, P. D. Thomas, C. H. Wu, C. Yeats and S. Y. Yong, InterPro in 2011: new developments in the family and domain prediction database, Nucleic Acids Res., 2011, 40, D306–D312 CrossRef PubMed.
  139. S. Burge, E. Kelly, D. Lonsdale, P. Mutowo-Muellenet, C. McAnulla, A. Mitchell, A. Sangrador-Vegas, S. Y. Yong, N. Mulder and S. Hunter, Manual GO annotation of predictive protein signatures: the InterPro approach to GO curation, Database, 2012, 2012, bar068 Search PubMed.
  140. A. Cuff, O. C. Redfern, L. Greene, I. Sillitoe, T. Lewis, M. Dibley, A. Reid, F. Pearl, T. Dallman, A. Todd, R. Garratt, J. Thornton and C. Orengo, The CATH hierarchy revisited-structural divergence in domain superfamilies and the continuity of fold space, Structure, 2009, 17, 1051–1062 CrossRef CAS PubMed.
  141. L. Xie and P. E. Bourne, Detecting evolutionary relationships across existing fold space, using sequence order-independent profile-profile alignments, Proc. Natl. Acad. Sci. U. S. A., 2008, 105, 5441–5446 CrossRef CAS PubMed.
  142. N. Furnham, I. Sillitoe, G. L. Holliday, A. L. Cuff, R. A. Laskowski, C. A. Orengo and J. M. Thornton, Exploring the evolution of novel enzyme functions within structurally defined protein superfamilies, PLoS Comput. Biol., 2012, 8, e1002403 CAS.
  143. G. J. Bartlett, N. Borkakoti and J. M. Thornton, Catalysing new reactions during evolution: economy of residues and mechanism, J. Mol. Biol., 2003, 331, 829–860 CrossRef CAS.
  144. P. F. Gherardini, M. N. Wass, M. Helmer-Citterich and M. J. Sternberg, Convergent evolution of enzyme active sites is not a rare phenomenon, J. Mol. Biol., 2007, 372, 817–845 CrossRef CAS PubMed.
  145. G. L. Holliday, C. Andreini, J. D. Fischer, S. A. Rahman, D. E. Almonacid, S. T. Williams and W. R. Pearson, MACiE: exploring the diversity of biochemical reactions, Nucleic Acids Res., 2011, 40, D783–D789 CrossRef PubMed.
  146. E. Ferrada and A. Wagner, Evolutionary innovations and the organization of protein functions in genotype space, PLoS One, 2010, 5, e14172 CAS.
  147. U. Bastolla, M. Porto, H. E. Roman and M. Vendruscolo, A protein evolution model with independent sites that reproduces site-specific amino acid distributions from the Protein Data Bank, BMC Evol. Biol., 2006, 6, 43 CrossRef PubMed.
  148. C. L. Worth, S. Gong and T. L. Blundell, Structural and functional constraints in the evolution of protein families, Nat. Rev. Mol. Cell Biol., 2009, 10, 709–720 CAS.
  149. C. L. Worth, R. Preissner and T. L. Blundell, SDM—a server for predicting effects of mutations on protein stability and malfunction, Nucleic Acids Res., 2011, 39, W215–W222 CrossRef CAS PubMed.
  150. J. Overington, D. Donnelly, M. S. Johnson, A. Šali and T. L. Blundell, Environment-specific amino acid substitution tables: tertiary templates and prediction of protein folds, Protein Sci., 1992, 1, 216–226 CrossRef CAS PubMed.
  151. A. Šali and T. L. Blundell, Comparative protein modelling by satisfaction of spatial restraints, J. Mol. Biol., 1993, 234, 779–815 CrossRef PubMed.
  152. M. E. Peterson, F. Chen, J. G. Saven, D. S. Roos, P. C. Babbitt and A. Sali, Evolutionary constraints on structural similarity in orthologs and paralogs, Protein Sci., 2009, 18, 1306–1315 CrossRef CAS PubMed.
  153. R. Mendez, M. Fritsche, M. Porto and U. Bastolla, Mutation bias favors protein folding stability in the evolution of small populations, PLoS Comput. Biol., 2010, 6, e1000767 Search PubMed.
  154. D. M. Taverna and R. A. Goldstein, The distribution of structures in evolving protein populations, Biopolymers, 2000, 53, 1–8 CrossRef CAS.
  155. D. M. Taverna and R. A. Goldstein, Why are proteins so robust to site mutations?, J. Mol. Biol., 2002, 315, 479–484 CrossRef CAS PubMed.
  156. D. M. Taverna and R. A. Goldstein, Why are proteins marginally stable?, Proteins, 2002, 46, 105–109 CrossRef CAS PubMed.
  157. M. A. DePristo, D. M. Weinreich and D. L. Hartl, Missense meanderings in sequence space: a biophysical view of protein evolution, Nat. Rev. Genet., 2005, 6, 678–687 CrossRef CAS PubMed.
  158. P. D. Williams, D. D. Pollock and R. A. Goldstein, Functionality and the evolution of marginal stability in proteins: inferences from lattice simulations, Evol. Bioinf. Online, 2006, 2, 91–101 Search PubMed.
  159. R. A. Goldstein, The evolution and evolutionary consequences of marginal thermostability in proteins, Proteins, 2011, 79, 1396–1407 CrossRef CAS PubMed.
  160. C. G. Langton, Life at the Edge of Chaos, SFI S Sci. C, 1992, 10, 41–91 Search PubMed.
  161. S. A. Kauffman, The origins of order, Oxford University Press, Oxford, 1993 Search PubMed.
  162. P. Csermely, K. S. Sandhu, E. Hazai, Z. Hoksza, H. J. Kiss, F. Miozzo, D. V. Veres, F. Piazza and R. Nussinov, Disordered proteins and network disorder in network descriptions of protein structure, dynamics and function: hypotheses and a comprehensive review, Curr. Protein Pept. Sci., 2012, 13, 19–33 CrossRef CAS.
  163. P. Csermely, T. Korcsmáros, H. J. M. Kiss, G. London and R. Nussinov, Structure and dynamics of molecular networks: A novel paradigm of drug discovery. A comprehensive review, Pharmacol. Therapeut., 2013, 138, 333–408 CrossRef CAS PubMed.
  164. J. M. Carlson and J. Doyle, Highly optimized tolerance: a mechanism for power laws in designed systems, Phys. Rev. E: Stat. Phys., Plasmas, Fluids, Relat. Interdiscip. Top., 1999, 60, 1412–1427 CrossRef CAS.
  165. J. M. Carlson and J. Doyle, Complexity and robustness, Proc. Natl. Acad. Sci. U. S. A., 2002, 99(suppl 1), 2538–2545 CrossRef PubMed.
  166. M. E. Csete and J. C. Doyle, Reverse engineering of biological complexity, Science, 2002, 295, 1664–1669 CrossRef CAS PubMed.
  167. M. Csete and J. Doyle, Bow ties, metabolism and disease, Trends Biotechnol., 2004, 22, 446–450 CrossRef CAS PubMed.
  168. T. Zhou, J. M. Carlson and J. Doyle, Evolutionary dynamics and highly optimized tolerance, J. Theor. Biol., 2005, 236, 438–447 CrossRef PubMed.
  169. F. J. Doyle 3rd and J. Stelling, Systems interface biology, J. R. Soc., Interface, 2006, 3, 603–616 CrossRef PubMed.
  170. P. Ao, Global view of bionetwork dynamics: adaptive landscape, J. Genet. Genomics, 2009, 36, 63–73 CrossRef.
  171. L. Giver, A. Gershenson, P. O. Freskgard and F. H. Arnold, Directed evolution of a thermostable esterase, Proc. Natl. Acad. Sci. U. S. A., 1998, 95, 12809–12813 CrossRef CAS.
  172. F. H. Arnold, P. L. Wintrode, K. Miyazaki and A. Gershenson, How enzymes adapt: lessons from directed evolution, Trends Biochem. Sci., 2001, 26, 100–106 CrossRef CAS.
  173. N. Tokuriki, F. Stricher, L. Serrano and D. S. Tawfik, How protein stability and new functions trade off, PLoS Comput. Biol., 2008, 4, e1000002 Search PubMed.
  174. J. D. Bloom, D. A. Drummond, F. H. Arnold and C. O. Wilke, Structural determinants of the rate of protein evolution in yeast, Mol. Biol. Evol., 2006, 23, 1751–1761 CrossRef CAS PubMed.
  175. J. D. Bloom, S. T. Labthavikul, C. R. Otey and F. H. Arnold, Protein stability promotes evolvability, Proc. Natl. Acad. Sci. U. S. A., 2006, 103, 5869–5874 CrossRef CAS PubMed.
  176. C. O. Wilke, J. D. Bloom, D. A. Drummond and A. Raval, Predicting the tolerance of proteins to random amino acid substitution, Biophys. J., 2005, 89, 3714–3720 CrossRef CAS PubMed.
  177. C. O. Wilke and D. A. Drummond, Signatures of protein biophysics in coding sequence evolution, Curr. Opin. Struct. Biol., 2010, 20, 385–389 CrossRef CAS PubMed.
  178. U. Bastolla, M. Porto, H. E. Roman and M. Vendruscolo, Looking at structure, stability, and evolution of proteins through the principal eigenvector of contact matrices and hydrophobicity profiles, Gene, 2005, 347, 219–230 CrossRef CAS PubMed.
  179. C. L. Worth and T. L. Blundell, Satisfaction of hydrogen-bonding potential influences the conservation of polar sidechains, Proteins, 2009, 75, 413–429 CrossRef CAS PubMed.
  180. J. D. Bloom, A. Raval and C. O. Wilke, Thermodynamics of neutral protein evolution, Genetics, 2007, 175, 255–266 CrossRef CAS PubMed.
  181. J. D. Bloom, P. A. Romero, Z. Lu and F. H. Arnold, Neutral genetic drift can alter promiscuous protein functions, potentially aiding functional evolution, Biol. Direct, 2007, 2, 17 CrossRef PubMed.
  182. R. D. Gupta and D. S. Tawfik, Directed enzyme evolution via small and effective neutral drift libraries, Nat. Methods, 2008, 5, 939–942 CrossRef CAS PubMed.
  183. D. L. Hartl, D. E. Dykhuizen and A. M. Dean, Limits of adaptation: the evolution of selective neutrality, Genetics, 1985, 111, 655–674 CAS.
  184. M. A. Huynen, P. F. Stadler and W. Fontana, Smoothness within ruggedness: The role of neutrality in adaptation, Proc. Natl. Acad. Sci. U. S. A., 1996, 93, 397–401 CrossRef CAS.
  185. C. M. Reidys and P. F. Stadler, Neutrality in fitness landscapes, Appl. Math. Comput., 2001, 117, 321–350 CrossRef.
  186. G. Amitai, R. D. Gupta and D. S. Tawfik, Latent evolutionary potentials under the neutral mutational drift of an enzyme, HFSP J., 2007, 1, 67–78 CrossRef CAS PubMed.
  187. J. Noirel and T. Simonson, Neutral evolution of proteins: The superfunnel in sequence space and its relation to mutational robustness, J. Chem. Phys., 2008, 129, 185104 CrossRef PubMed.
  188. A. Wagner, Neutralism and selectionism: a network-based reconciliation, Nat. Rev. Genet., 2008, 9, 965–974 CrossRef CAS PubMed.
  189. S. G. Peisajovich, L. Rockah and D. S. Tawfik, Evolution of new protein topologies through multistep gene rearrangements, Nat. Genet., 2006, 38, 168–174 CrossRef CAS PubMed.
  190. L. Pritchard, P. Bladon, J. M. O. Mitchell and M. J. Dufton, Evaluation of a novel method for the identification of coevolving protein residues, Protein Eng., 2001, 14, 549–555 CrossRef CAS PubMed.
  191. S. Govindarajan, J. E. Ness, S. Kim, E. C. Mundorff, J. Minshull and C. Gustafsson, Systematic variation of amino acid substitutions for stringent assessment of pairwise covariation, J. Mol. Biol., 2003, 328, 1061–1069 CrossRef CAS.
  192. A. F. Neuwald, Surveying the Manifold Divergence of an Entire Protein Class for Statistical Clues to Underlying Biochemical Mechanisms, Stat. Appl. Genet. Mol. Biol., 2011, 10, 36 Search PubMed.
  193. L. Pritchard and M. J. Dufton, Do proteins learn to evolve? The Hopfield network as a basis for the understanding of protein evolution, J. Theor. Biol., 2000, 202, 77–86 CrossRef CAS PubMed.
  194. V. Chelliah, L. Chen, T. L. Blundell and S. C. Lovell, Distinguishing structural and functional restraints in evolution in order to identify interaction sites, J. Mol. Biol., 2004, 342, 1487–1504 CrossRef CAS PubMed.
  195. V. Chelliah and T. L. Blundell, Quantifying structural and functional restraints on amino acid substitutions in evolution of proteins, Biochemistry, 2005, 70, 835–840 CAS.
  196. V. Chelliah, T. Blundell and K. Mizuguchi, Functional restraints on the patterns of amino acid substitutions: application to sequence-structure homology recognition, Proteins, 2005, 61, 722–731 CrossRef CAS PubMed.
  197. D. S. Marks, L. J. Colwell, R. Sheridan, T. A. Hopf, A. Pagnani, R. Zecchina and C. Sander, Protein 3D structure computed from evolutionary sequence variation, PLoS One, 2011, 6, e28766 CAS.
  198. T. A. Hopf, L. J. Colwell, R. Sheridan, B. Rost, C. Sander and D. S. Marks, Three-dimensional structures of membrane proteins from genomic sequencing, Cell, 2012, 149, 1607–1621 CrossRef CAS PubMed.
  199. F. Morcos, A. Pagnani, B. Lunt, A. Bertolino, D. S. Marks, C. Sander, R. Zecchina, J. N. Onuchic, T. Hwa and M. Weigt, Direct-coupling analysis of residue coevolution captures native contacts across many protein families, Proc. Natl. Acad. Sci. U. S. A., 2011, 108, E1293–E1301 CrossRef CAS PubMed.
  200. R. N. McLaughlin Jr, F. J. Poelwijk, A. Raman, W. S. Gosal and R. Ranganathan, The spatial architecture of protein function and adaptation, Nature, 2012, 491, 138–142 CrossRef PubMed.
  201. P. E. Tomatis, S. M. Fabiane, F. Simona, P. Carloni, B. J. Sutton and A. J. Vila, Adaptive protein evolution grants organismal fitness by improving catalysis and flexibility, Proc. Natl. Acad. Sci. U. S. A., 2008, 105, 20605–20610 CrossRef CAS PubMed.
  202. M. I. Sadowski and D. T. Jones, An automatic method for assessing structural importance of amino acid positions, BMC Struct. Biol., 2009, 9, 10 CrossRef PubMed.
  203. L. A. Abriata, M. S. ML and P. E. Tomatis, Sequence-function-stability relationships in proteins from datasets of functionally annotated variants: The case of TEM beta-lactamases, FEBS Lett., 2012, 586, 3330–3335 CrossRef CAS PubMed.
  204. L. Burger and E. van Nimwegen, Accurate prediction of protein-protein interactions from sequence alignments using a Bayesian method, Mol. Syst. Biol., 2008, 4, 165 CrossRef PubMed.
  205. M. Weigt, R. A. White, H. Szurmant, J. A. Hoch and T. Hwa, Identification of direct residue contacts in protein-protein interaction by message passing, Proc. Natl. Acad. Sci. U. S. A., 2009, 106, 67–72 CrossRef CAS PubMed.
  206. L. Burger and E. van Nimwegen, Disentangling Direct from Indirect Co-Evolution of Residues in Protein Alignments, PLoS Comput. Biol., 2010, 6, e1000633 Search PubMed.
  207. A. Rausell, D. Juan, F. Pazos and A. Valencia, Protein interactions and ligand binding: from protein subfamilies to functional specificity, Proc. Natl. Acad. Sci. U. S. A., 2010, 107, 1995–2000 CrossRef CAS PubMed.
  208. J. Strafford, P. Payongsri, E. G. Hibbert, P. Morris, S. S. Batth, D. Steadman, M. E. B. Smith, J. M. Ward, H. C. Hailes and P. A. Dalby, Directed evolution to re-adapt a co-evolved network within an enzyme, J. Biotechnol., 2012, 157, 237–245 CrossRef CAS PubMed.
  209. D. de Juan, F. Pazos and A. Valencia, Emerging methods in protein co-evolution, Nat. Rev. Genet., 2013, 14, 249–261 CrossRef CAS PubMed.
  210. J. H. Holland, Adaption in natural and artificial systems: an introductory analysis with applications to biology, control, and artificial intelligence, MIT Press, 1992 Search PubMed.
  211. D. E. Goldberg, Genetic algorithms in search, optimization and machine learning, Addison-Wesley, 1989 Search PubMed.
  212. D. E. Goldberg, The design of innovation: lessons from and for competent genetic algorithms, Kluwer, Boston, 2002 Search PubMed.
  213. J. Knowles, Closed-Loop Evolutionary Multiobjective Optimization, IEEE Computational Intelligence Magazine, 2009, 4, 77–91 CrossRef.
  214. D. B. Fogel, Evolutionary computation: toward a new philosophy of machine intelligence, IEEE Press, Piscataway, 1995 Search PubMed.
  215. Evolutionary computation in bioinformatics, ed. G. B. Fogel and D. W. Corne, Morgan Kaufmann, Amsterdam, 2003 Search PubMed.
  216. D. Ashlock, Evolutionary computation for modeling and optimization, Springer, New York, 2006 Search PubMed.
  217. N. Hamamatsu, T. Aita, Y. Nomiya, H. Uchiyama, M. Nakajima, Y. Husimi and Y. Shibanaka, Biased mutation-assembling: an efficient method for rapid directed evolution through simultaneous mutation accumulation, Protein Eng., Des. Sel., 2005, 18, 265–271 CrossRef CAS PubMed.
  218. N. Hamamatsu, Y. Nomiya, T. Aita, M. Nakajima, Y. Husimi and Y. Shibanaka, Directed evolution by accumulating tailored mutations: thermostabilization of lactate oxidase with less trade-off with catalytic activity, Protein Eng., Des. Sel., 2006, 19, 483–489 CrossRef CAS PubMed.
  219. R. J. Fox, S. C. Davis, E. C. Mundorff, L. M. Newman, V. Gavrilovic, S. K. Ma, L. M. Chung, C. Ching, S. Tam, S. Muley, J. Grate, J. Gruber, J. C. Whitman, R. A. Sheldon and G. W. Huisman, Improving catalytic function by ProSAR-driven enzyme evolution, Nat. Biotechnol., 2007, 25, 338–344 CrossRef CAS PubMed.
  220. D. Wedge, W. Rowe, D. B. Kell and J. Knowles, In silico modelling of directed evolution: implications for experimental design and stepwise evolution, J. Theor. Biol., 2009, 257, 131–141 CrossRef PubMed.
  221. M. Carneiro and D. L. Hartl, Adaptive landscapes and protein evolution, Proc. Natl. Acad. Sci. U. S. A., 2011, 107(suppl 1), 1747–1751 Search PubMed.
  222. J. H. Gillespie, A simple stochastic gene substitution model, Theor. Popul. Biol., 1983, 23, 202–215 CrossRef CAS.
  223. J. H. Gillespie, Molecular Evolution over the Mutational Landscape, Evolution, 1984, 38, 1116–1129 CrossRef CAS.
  224. H. A. Orr, The genetic theory of adaptation: a brief history, Nat. Rev. Genet., 2005, 6, 119–127 CrossRef CAS PubMed.
  225. H. A. Orr, The population genetics of adaptation on correlated fitness landscapes: the block model, Evolution, 2006, 60, 1113–1124 CrossRef PubMed.
  226. H. A. Orr, The distribution of fitness effects among beneficial mutations in Fisher's geometric model of adaptation, J. Theor. Biol., 2006, 238, 279–285 CrossRef PubMed.
  227. H. A. Orr, Fitness and its role in evolutionary genetics, Nat. Rev. Genet., 2009, 10, 531–539 CrossRef CAS PubMed.
  228. R. L. Unckless and H. A. Orr, The population genetics of adaptation: multiple substitutions on a smooth fitness landscape, Genetics, 2009, 183, 1079–1086 CrossRef PubMed.
  229. I. G. Szendro, J. Franke, J. A. G. M. de Visser and J. Krug, Predictability of evolution depends nonmonotonically on population size, Proc. Natl. Acad. Sci. U. S. A., 2013, 110, 571–576 CrossRef CAS PubMed.
  230. J. A. Wells, Additivity of mutational effects in proteins, Biochemistry, 1990, 29, 8509–8517 CrossRef CAS.
  231. M. Lunzer, S. P. Miller, R. Felsheim and A. M. Dean, The biochemical architecture of an ancient adaptive landscape, Science, 2005, 310, 499–501 CrossRef CAS PubMed.
  232. R. Fox, A. Roy, S. Govindarajan, J. Minshull, C. Gustafsson, J. T. Jones and R. Emig, Optimizing the search algorithm for protein engineering by directed evolution, Protein Eng., 2003, 16, 589–597 CrossRef CAS PubMed.
  233. R. Fox, Directed molecular evolution by machine learning and the influence of nonlinear interactions, J. Theor. Biol., 2005, 234, 187–199 CrossRef CAS PubMed.
  234. M. Iwakura, K. Maki, H. Takahashi, T. Takenawa, A. Yokota, K. Katayanagi, T. Kamiyama and K. Gekko, Evolutional design of a hyperactive cysteine- and methionine-free mutant of Escherichia coli dihydrofolate reductase, J. Biol. Chem., 2006, 281, 13234–13246 CrossRef CAS PubMed.
  235. C. A. Tracewell and F. H. Arnold, Directed enzyme evolution: climbing fitness peaks one amino acid at a time, Curr. Opin. Chem. Biol., 2009, 13, 3–9 CrossRef CAS PubMed.
  236. M. T. Reetz, The importance of additive and non-additive mutational effects in protein engineering, Angew. Chem., Int. Ed., 2013, 52, 2658–2666 CrossRef CAS PubMed.
  237. Y. Hayashi, T. Aita, H. Toyota, Y. Husimi, I. Urabe and T. Yomo, Experimental rugged fitness landscape in protein sequence space, PLoS One, 2006, 1, e96 Search PubMed.
  238. J. D. Bloom, F. H. Arnold and C. O. Wilke, Breaking proteins with mutations: threads and thresholds in evolution, Mol. Syst. Biol., 2007, 3, 76 CrossRef PubMed.
  239. K. Jain and S. Seetharaman, Multiple adaptive substitutions during evolution in novel environments, Genetics, 2011, 189, 1029–1043 CrossRef PubMed.
  240. J. D. Bloom, L. I. Gong and D. Baltimore, Permissive secondary mutations enable the evolution of influenza oseltamivir resistance, Science, 2010, 328, 1272–1275 CrossRef CAS PubMed.
  241. T. Aita, N. Hamamatsu, Y. Nomiya, H. Uchiyama, Y. Shibanaka and Y. Husimi, Surveying a local fitness landscape of a protein with epistatic sites for the study of directed evolution, Biopolymers, 2002, 64, 95–105 CrossRef CAS PubMed.
  242. B. Østman, A. Hintze and C. Adami, Impact of epistasis and pleiotropy on evolutionary adaptation, Proc. R. Soc. B, 2011, 279, 247–256 CrossRef PubMed.
  243. M. S. Breen, C. Kemena, P. K. Vlasov, C. Notredame and F. A. Kondrashov, Epistasis as the primary factor in molecular evolution, Nature, 2012, 490, 535–538 CrossRef CAS PubMed.
  244. C. Natarajan, N. Inoguchi, R. E. Weber, A. Fago, H. Moriyama and J. F. Storz, Epistasis among adaptive mutations in deer mouse hemoglobin, Science, 2013, 340, 1324–1327 CrossRef CAS PubMed.
  245. J. L. Rummer, D. J. McKenzie, A. Innocenti, C. T. Supuran and C. J. Brauner, Root effect hemoglobin may have evolved to enhance general tissue oxygen delivery, Science, 2013, 340, 1327–1329 CrossRef CAS PubMed.
  246. S. Mirceta, A. V. Signore, J. M. Burns, A. R. Cossins, K. L. Campbell and M. Berenbrink, Evolution of mammalian diving capacity traced by myoglobin net surface charge, Science, 2013, 340, 1234192 CrossRef PubMed.
  247. P. A. Alexander, Y. He, Y. Chen, J. Orban and P. N. Bryan, A minimal sequence code for switching protein structure and function, Proc. Natl. Acad. Sci. U. S. A., 2009, 106, 21149–21154 CrossRef CAS PubMed.
  248. L. I. Gong and J. D. Bloom, Epistatically Interacting Substitutions Are Enriched during Adaptive Protein Evolution, PLoS Genet., 2014, 10, e1004328 Search PubMed.
  249. M. Lunzer, G. B. Golding and A. M. Dean, Pervasive cryptic epistasis in molecular evolution, PLoS Genet., 2010, 6, e1001162 Search PubMed.
  250. S. G. Williams and S. C. Lovell, The effect of sequence evolution on protein structural divergence, Mol. Biol. Evol., 2009, 26, 1055–1065 CrossRef CAS PubMed.
  251. Y. Yoshikuni, T. E. Ferrin and J. D. Keasling, Designed divergent evolution of enzyme function, Nature, 2006, 440, 1078–1082 CrossRef CAS PubMed.
  252. K. Hult and P. Berglund, Enzyme promiscuity: mechanism and applications, Trends Biotechnol., 2007, 25, 231–238 CrossRef CAS PubMed.
  253. I. Nobeli, A. D. Favia and J. M. Thornton, Protein promiscuity and its implications for biotechnology, Nat. Biotechnol., 2009, 27, 157–167 CrossRef CAS PubMed.
  254. N. Tokuriki and D. S. Tawfik, Protein dynamism and evolvability, Science, 2009, 324, 203–207 CrossRef CAS PubMed.
  255. A. Babtie, N. Tokuriki and F. Hollfelder, What makes an enzyme promiscuous?, Curr. Opin. Chem. Biol., 2010, 14, 200–207 CrossRef CAS PubMed.
  256. O. Khersonsky and D. S. Tawfik, Enzyme promiscuity: a mechanistic and evolutionary perspective, Annu. Rev. Biochem., 2010, 79, 471–505 CrossRef CAS PubMed.
  257. O. Khersonsky and D. S. Tawfik, Enzyme promiscuity: evolutionary and mechanistic aspects, in Comprehensive Natural Products II Chemistry and Biology, ed. L. Mander and H. -W. Lui, Elsevier, Oxford, 2010, pp. 48–90 Search PubMed.
  258. A. Aharoni, L. Gaidukov, O. Khersonsky, Q. G. S. Mc, C. Roodveldt and D. S. Tawfik, The 'evolvability' of promiscuous protein functions, Nat. Genet., 2005, 37, 73–76 CAS.
  259. M. S. Humble and P. Berglund, Biocatalytic Promiscuity, Eur. J. Org. Chem., 2011, 3391–3401 CrossRef CAS.
  260. R. Huang, F. Hippauf, D. Rohrbeck, M. Haustein, K. Wenke, J. Feike, N. Sorrelle, B. Piechulla and T. J. Barkman, Enzyme functional evolution through improved catalysis of ancestrally nonpreferred substrates, Proc. Natl. Acad. Sci. U. S. A., 2012, 109, 2966–2971 CrossRef CAS PubMed.
  261. D. R. Burton, P. Poignard, R. L. Stanfield and I. A. Wilson, Broadly neutralizing antibodies present new prospects to counter highly antigenically diverse viruses, Science, 2012, 337, 183–186 CrossRef CAS PubMed.
  262. H. Garcia-Seisdedos, B. Ibarra-Molero and J. M. Sanchez-Ruiz, Probing the mutational interplay between primary and promiscuous protein functions: a computational-experimental approach, PLoS Comput. Biol., 2012, 8, e1002558 CAS.
  263. H. Nam, N. E. Lewis, J. A. Lerman, D. H. Lee, R. L. Chang, D. Kim and B. Ø. Palsson, Network context and selection in the evolution to enzyme specificity, Science, 2012, 337, 1101–1104 CrossRef CAS PubMed.
  264. J. H. Luo, B. van Loo and S. C. L. Kamerlin, Catalytic promiscuity in Pseudomonas aeruginosa arylsulfatase as an example of chemistry-driven protein evolution, FEBS Lett., 2012, 586, 1622–1630 CrossRef CAS PubMed.
  265. S. Chakraborty, An automated flow for directed evolution based on detection of promiscuous scaffolds using spatial and electrostatic properties of catalytic residues, PLoS One, 2012, 7, e40408 CAS.
  266. P. Gatti-Lafranconi and F. Hollfelder, Flexibility and reactivity in promiscuous enzymes, ChemBioChem, 2013, 14, 285–292 CrossRef CAS PubMed.
  267. P. Carbonell, G. Lecointre and J. L. Faulon, Origins of specificity and promiscuity in metabolic networks, J. Biol. Chem., 2011, 286, 43994–44004 CrossRef CAS PubMed.
  268. P. Carbonell and J. L. Faulon, Molecular signatures-based prediction of enzyme promiscuity, Bioinformatics, 2010, 26, 2012–2019 CrossRef CAS PubMed.
  269. D. B. Kell, P. D. Dobson, E. Bilsland and S. G. Oliver, The promiscuous binding of pharmaceutical drugs and their transporter-mediated uptake into cells: what we (need to) know and how we can do so, Drug Discovery Today, 2013, 18, 218–239 CrossRef CAS PubMed.
  270. S. Chakraborty, R. Minda, L. Salaye, A. M. Dandekar, S. K. Bhattacharjee and B. J. Rao, Promiscuity-based enzyme selection for rational directed evolution experiments, Methods Mol. Biol., 2013, 978, 205–216 CAS.
  271. S. A. Kauffman and E. D. Weinberger, The NK model of rugged fitness landscapes and its application to maturation of the immune response, J. Theor. Biol., 1989, 141, 211–245 CrossRef CAS.
  272. L. Barnett, Ruggedness and neutrality: the NKp family of fitness landscapes, Proc. 6th Int'l Conf. on Artificial Life, MIT Press, 1998, pp. 17–27 Search PubMed.
  273. T. Aita, Hierarchical distribution of ascending slopes, nearly neutral networks, highlands, and local optima at the dth order in an NK fitness landscape, J. Theor. Biol., 2008, 254, 252–263 CrossRef PubMed.
  274. T. Aita and Y. Husimi, Fitting protein-folding free energy landscape for a certain conformation to an NK fitness landscape, J. Theor. Biol., 2008, 253, 151–161 CrossRef CAS PubMed.
  275. B. Østman, A. Hintze and C. Adami, Critical Properties of Complex Fitness Landscapes, Proc XII Conf Alife, 2010, 126–132 Search PubMed.
  276. W. Rowe, D. C. Wedge, M. Platt, D. B. Kell and J. Knowles, Predictive models for population performance on real biological fitness landscapes, Bioinformatics, 2010, 26, 2125–2142 CrossRef PubMed.
  277. N. A. Rosenberg, A sharp minimum on the mean number of steps taken in adaptive walks, J. Theor. Biol., 2005, 237, 17–22 CrossRef PubMed.
  278. D. B. Kell and E. Lurie-Luke, The Virtue of Innovation: Innovation through the Lenses of Biological Evolution, J. R. Soc., Interface, 2015 Search PubMed , in the press.
  279. C. R. Reeves and J. E. Rowe, Genetic algorithms – principles and perspectives: a guide to GA theory, Kluwer Academic Publishers, Dordrecht, 2002 Search PubMed.
  280. T. Aita and Y. Husimi, Adaptive walks by the fittest among finite random mutants on a Mt. Fuji-type fitness landscape - II. Effect of small non- additivity, J. Math. Biol., 2000, 41, 207–231 CrossRef CAS.
  281. T. Aita, H. Uchiyama, T. Inaoka, M. Nakajima, T. Kokubo and Y. Husimi, Analysis of a local fitness landscape with a model of the rough Mt. Fuji-type landscape: Application to prolyl endopeptidase and thermolysin, Biopolymers, 2000, 54, 64–79 CrossRef CAS.
  282. E. D. Weinberger, NP completeness of Kauffman's NK model: a tuneably rugged fitness landscape, Santa Fe Institute Technical Report, 1996, 96-02-003 Search PubMed.
  283. N. Tokuriki, C. J. Jackson, L. Afriat-Jurnou, K. T. Wyganowski, R. Tang and D. S. Tawfik, Diminishing returns and tradeoffs constrain the laboratory optimization of an enzyme, Nat. Commun., 2012, 3, 1257 CrossRef PubMed.
  284. D. T. Campbell, Blind variation and selective retention in creative thought as in other knowledge processes, Psychol. Rev., 1960, 67, 380–400 CrossRef CAS PubMed.
  285. J. W. Rivkin and N. Siggelkow, Organisational sticking points on NK landscapes, Complexity, 2002, 7, 31–43 CrossRef.
  286. K. Frenken, Technological innovation and complexity theory, Econ. Innov. New Technol., 2006, 15, 137–155 CrossRef.
  287. S. Geisendorf, Searching NK Fitness Landscapes: On the Trade Off Between Speed and Quality in Complex Problem Solving, Comput. Econ., 2010, 35, 395–406 CrossRef PubMed.
  288. S. Johnson, Where good ideas come from: the seven patterns of innovation, Penguin, London, 2011 Search PubMed.
  289. P. Auerswald, S. Kauffman, J. Lobo and K. Shell, The production recipes approach to modeling technological innovation: An application to learning by doing, J. Econ. Dyn. Control, 2000, 24, 389–450 CrossRef.
  290. S. Kauffman, J. Lobo and W. G. Macready, Optimal search on a technology landscape, J. Econ. Behav. Organ., 2000, 43, 141–166 CrossRef.
  291. M. Ganco and G. Hoetker, NK Modeling Methodology in the Strategy Literature: Bounded Search on a Rugged Landscape, Res Methodol Strat Mangement, 2009, 5, 237–268 Search PubMed.
  292. G. M. Hodgson and T. Knudsen, Balancing inertia, innovation, and imitation in complex environments, J. Econ. Issues, 2006, 40, 287–295 Search PubMed.
  293. L. Gabora, An evolutionary framework for cultural change: Selectionism versus communal exchange, Phys. Life Rev., 2013, 10, 117–145 CrossRef PubMed.
  294. R. V. Solé, S. Valverde, M. R. Casals, S. A. Kauffman, D. Farmer and N. Eldredge, The Evolutionary Ecology of Technological Innovations, Complexity, 2013, 18, 15–27 CrossRef.
  295. A. Wagner and W. Rosen, Spaces of the possible: universal Darwinism and the wall between technological and biological innovation, J. R. Soc., Interface, 2014, 11, 20131190 CrossRef PubMed.
  296. Directed evolution library creation: methods and protocols, ed. F. H. Arnold and G. Georgiou, Springer, Berlin, 1996 Search PubMed.
  297. Directed molecular evolution of proteins, ed. S. Brakmann and K. Johnsoon, Wiley-VCH, Weinheim, 2002 Search PubMed.
  298. C. A. Voigt, S. Kauffman and Z. G. Wang, Rational evolutionary design: The theory of in vitro protein evolution, in Adv. Protein Chem., ed. F. M. Arnold, 2001, vol. 55, pp. 79–160 Search PubMed.
  299. F. H. Arnold and A. A. Volkov, Directed evolution of biocatalysts, Curr. Opin. Biotechnol., 1999, 3, 54–59 CrossRef CAS.
  300. Evolutionary protein design, ed. F. H. Arnold, Academic Press, San Diego, 2001 Search PubMed.
  301. K. A. Powell, S. W. Ramer, S. B. Del Cardayré, W. P. C. Stemmer, M. B. Tobin, P. F. Longchamp and G. W. Huisman, Directed Evolution and Biocatalysis, Angew. Chem., Int. Ed., 2001, 40, 3948–3959 CrossRef CAS.
  302. C. Schmidt-Dannert, Directed evolution of single proteins, metabolic pathways, and viruses, Biochemistry, 2001, 40, 13125–13136 CrossRef CAS.
  303. S. V. Taylor, P. Kast and D. Hilvert, Investigating and engineering enzymes by genetic selection, Angew. Chem., Int. Ed., 2001, 40, 3311–3335 CrossRef.
  304. H. Tao and V. W. Cornish, Milestones in directed enzyme evolution, Curr. Opin. Chem. Biol., 2002, 6, 858–864 CrossRef CAS.
  305. P. A. Dalby, Optimising enzyme function by directed evolution, Curr. Opin. Struct. Biol., 2003, 13, 500–505 CrossRef CAS.
  306. N. J. Turner, Directed evolution of enzymes for applied biocatalysis, Trends Biotechnol., 2003, 21, 474–478 CrossRef CAS PubMed.
  307. C. Neylon, Chemical and biochemical strategies for the randomization of protein encoding DNA sequences: library construction methods for directed evolution, Nucleic Acids Res., 2004, 32, 1448–1459 CrossRef CAS PubMed.
  308. H. Leemhuis, V. Stein, A. D. Griffiths and F. Hollfelder, New genotype-phenotype linkages for directed evolution of functional proteins, Curr. Opin. Struct. Biol., 2005, 15, 472–478 CrossRef CAS PubMed.
  309. L. Yuan, I. Kurek, J. English and R. Keenan, Laboratory-directed protein evolution, Microbiol. Mol. Biol. Rev., 2005, 69, 373–392 CrossRef CAS PubMed.
  310. T. Matsuura and T. Yomo, In vitro evolution of proteins, J. Biosci. Bioeng., 2006, 101, 449–456 CrossRef CAS PubMed.
  311. P. A. Dalby, Engineering enzymes for biocatalysis, Recent Pat. Biotechnol., 2007, 1, 1–9 CrossRef CAS.
  312. S. Sen, V. Venkata Dasu and B. Mandal, Developments in directed evolution for improving enzyme functions, Appl. Biochem. Biotechnol., 2007, 143, 212–223 CrossRef CAS.
  313. C. C. Akoh, S. W. Chang, G. C. Lee and J. F. Shaw, Biocatalysis for the production of industrial products and functional foods from rice and other agricultural produce, J. Agric. Food Chem., 2008, 56, 10445–10451 CrossRef CAS PubMed.
  314. J. C. Chaput, N. W. Woodbury, L. A. Stearns and B. A. Williams, Creating protein biocatalysts as tools for future industrial applications, Expert Opin. Biol. Ther., 2008, 8, 1087–1098 CrossRef CAS PubMed.
  315. R. N. Patel, Synthesis of chiral pharmaceutical intermediates by biocatalysis, Coord. Chem. Rev., 2008, 252, 659–701 CrossRef CAS PubMed.
  316. C. Jäckel, P. Kast and D. Hilvert, Protein design by directed evolution, Annu. Rev. Biophys., 2008, 37, 153–173 CrossRef PubMed.
  317. F. H. Arnold, How proteins adapt: lessons from directed evolution, Cold Spring Harbor Symp. Quant. Biol., 2009, 74, 41–46 CrossRef CAS PubMed.
  318. S. C. Stebel, A. Gaida, K. M. Arndt and K. M. Müller, Directed Protein Evolution, in Molecular Biomethods Handbook, ed. J. M. Walker and R. Rapley, Humana Press, Totowa, NJ, 2nd edn, 2008, pp. 631–656 Search PubMed.
  319. N. J. Turner, Directed evolution drives the next generation of biocatalysts, Nat. Chem. Biol., 2009, 5, 567–573 CrossRef CAS PubMed.
  320. C. Jäckel and D. Hilvert, Biocatalysts by evolution, Curr. Opin. Biotechnol., 2010, 21, 753–759 CrossRef PubMed.
  321. P. A. Dalby, Strategy and success for the directed evolution of enzymes, Curr. Opin. Struct. Biol., 2011, 21, 473–480 CrossRef CAS PubMed.
  322. N. J. Turner and M. D. Truppo, Biocatalysis enters a new era, Curr. Opin. Chem. Biol., 2013, 17, 212–214 CrossRef CAS PubMed.
  323. M. T. Reetz, J. D. Carballeira, J. Peyralans, H. Hobenreich, A. Maichele and A. Vogel, Expanding the substrate scope of enzymes: combining mutations obtained by CASTing, Chemistry, 2006, 12, 6031–6038 CrossRef CAS PubMed.
  324. M. T. Reetz, D. Kahakeaw and R. Lohmer, Addressing the numbers problem in directed evolution, ChemBioChem, 2008, 9, 1797–1804 CrossRef CAS PubMed.
  325. G. A. Behrens, A. Hummel, S. K. Padhi, S. Schätzle and U. T. Bornscheuer, Discovery and Protein Engineering of Biocatalysts for Organic Synthesis, Adv. Synth. Catal., 2011, 353, 2191–2215 CrossRef CAS.
  326. M. T. Reetz, Laboratory evolution of stereoselective enzymes: a prolific source of catalysts for asymmetric reactions, Angew. Chem., Int. Ed., 2011, 50, 138–174 CrossRef CAS PubMed.
  327. G. A. Strohmeier, H. Pichler, O. May and M. Gruber-Khadjawi, Application of designed enzymes in organic synthesis, Chem. Rev., 2011, 111, 4141–4164 CrossRef CAS PubMed.
  328. U. T. Bornscheuer, G. W. Huisman, R. J. Kazlauskas, S. Lutz, J. C. Moore and K. Robins, Engineering the third wave of biocatalysis, Nature, 2012, 485, 185–194 CrossRef CAS PubMed.
  329. M. Goldsmith and D. S. Tawfik, Directed enzyme evolution: beyond the low-hanging fruit, Curr. Opin. Struct. Biol., 2012, 22, 406–412 CrossRef CAS PubMed.
  330. R. Verma, U. Schwaneberg and D. Roccatano, Computer-Aided Protein Directed Evolution: a Review of Web Servers, Databases and other Computational Tools for Protein Engineering, Comput. Struct. Biotechnol. J., 2012, 2, e201209008 Search PubMed.
  331. T. Davids, M. Schmidt, D. Bottcher and U. T. Bornscheuer, Strategies for the discovery and engineering of enzymes for biocatalysis, Curr. Opin. Chem. Biol., 2013, 17, 215–220 CrossRef CAS PubMed.
  332. A. Kumar and S. Singh, Directed evolution: tailoring biocatalysts for industrial applications, Crit. Rev. Biotechnol., 2013, 33, 365–378 CrossRef CAS PubMed.
  333. M. T. Reetz, Biocatalysis in organic chemistry and biotechnology: past, present, and future, J. Am. Chem. Soc., 2013, 135, 12480–12496 CrossRef CAS PubMed.
  334. J. Damborsky and J. Brezovsky, Computational tools for designing and engineering enzymes, Curr. Opin. Chem. Biol., 2014, 19, 8–16 CrossRef CAS PubMed.
  335. P. D. Dobson, Y. Patel and D. B. Kell, “Metabolite-likeness” as a criterion in the design and selection of pharmaceutical drug libraries, Drug Discovery Today, 2009, 14, 31–40 CrossRef CAS PubMed.
  336. S. O'Hagan, N. Swainston, J. Handl and D. B. Kell, A 'rule of 0.5' for the metabolite-likeness of approved pharmaceutical drugs, Metabolomics, 2015 DOI:10.1007/s11306-11014-10733-z.
  337. E. Akiva, S. Brown, D. E. Almonacid, A. E. Barber, 2nd, A. F. Custer, M. A. Hicks, C. C. Huang, F. Lauck, S. T. Mashiyama, E. C. Meng, D. Mischel, J. H. Morris, S. Ojha, A. M. Schnoes, D. Stryke, J. M. Yunes, T. E. Ferrin, G. L. Holliday and P. C. Babbitt, The structure-function linkage database, Nucleic Acids Res., 2014, 42, D521–D530 CrossRef CAS PubMed.
  338. K. A. Armstrong and B. Tidor, Computationally mapping sequence space to understand evolutionary protein engineering, Biotechnol. Prog., 2008, 24, 62–73 CrossRef CAS PubMed.
  339. Y. Nov, Probabilistic methods in directed evolution: library size, mutation rate, and diversity, Methods Mol. Biol., 2014, 1179, 261–278 Search PubMed.
  340. J. Zaugg, Y. Gumulya, E. M. Gillam and M. Boden, Computational tools for directed evolution: a comparison of prospective and retrospective strategies, Methods Mol. Biol., 2014, 1179, 315–333 Search PubMed.
  341. A. Pavelka, E. Chovancova and J. Damborsky, Hot Spot Wizard: a web server for identification of hot spots in protein engineering, Nucleic Acids Res., 2009, 37, W376–W383 CrossRef CAS PubMed.
  342. M. Höhne, S. Schätzle, H. Jochens, K. Robins and U. T. Bornscheuer, Rational assignment of key motifs for function guides in silico enzyme identification, Nat. Chem. Biol., 2010, 6, 807–813 CrossRef PubMed.
  343. H. Jochens and U. T. Bornscheuer, Natural Diversity to Guide Focused Directed Evolution, ChemBioChem, 2010, 11, 1861–1866 CrossRef CAS PubMed.
  344. F. Sievers, A. Wilm, D. Dineen, T. J. Gibson, K. Karplus, W. Li, R. Lopez, H. McWilliam, M. Remmert, J. Soding, J. D. Thompson and D. G. Higgins, Fast, scalable generation of high-quality protein multiple sequence alignments using Clustal Omega, Mol. Syst. Biol., 2011, 7, 539 CrossRef PubMed.
  345. F. Sievers and D. G. Higgins, Clustal Omega, accurate alignment of very large numbers of sequences, Methods Mol. Biol., 2014, 1079, 105–116 Search PubMed.
  346. R. C. Edgar, MUSCLE: multiple sequence alignment with high accuracy and high throughput, Nucleic Acids Res., 2004, 32, 1792–1797 CrossRef CAS PubMed.
  347. J. Pei, B. H. Kim, M. Tang and N. V. Grishin, PROMALS web server for accurate multiple protein sequence alignments, Nucleic Acids Res., 2007, 35, W649–W652 CrossRef PubMed.
  348. J. Pei and N. V. Grishin, PROMALS3D: multiple protein sequence alignment enhanced with evolutionary and three-dimensional structural information, Methods Mol. Biol., 2014, 1079, 263–271 Search PubMed.
  349. L. Xie, L. Xie and P. E. Bourne, A unified statistical model to support local sequence order independent similarity searching for ligand-binding sites and its application to genome-based drug discovery, Bioinformatics, 2009, 25, i305–312 CrossRef CAS PubMed.
  350. L. Xie, L. Xie and P. E. Bourne, Structure-based systems biology for analyzing off-target binding, Curr. Opin. Struct. Biol., 2011, 21, 189–199 CrossRef CAS PubMed.
  351. T. Uchiyama and K. Miyazaki, Functional metagenomics for enzyme discovery: challenges to efficient screening, Curr. Opin. Biotechnol., 2009, 20, 616–622 CrossRef CAS PubMed.
  352. M. M. Schofield and D. H. Sherman, Meta-omic characterization of prokaryotic gene clusters for natural product biosynthesis, Curr. Opin. Biotechnol., 2013, 24, 1151–1158 CrossRef CAS PubMed.
  353. P. Medawar, Pluto's republic, Oxford University Press, Oxford, 1982 Search PubMed.
  354. K. R. Popper, Conjectures and refutations: the growth of scientific knowledge, Routledge & Kegan Paul, London, 5th edn, 1992 Search PubMed.
  355. A. F. Chalmers, What is this thing called Science? An assessment of the nature and status of science and its methods, Open University Press, Maidenhead, 1999 Search PubMed.
  356. D. B. Kell and S. G. Oliver, How drugs get into cells: tested and testable predictions to help discriminate between transporter-mediated uptake and lipoidal bilayer diffusion, Front. Pharmacol., 2014, 5, 231 Search PubMed.
  357. D. B. Kell, What would be the observable consequences if phospholipid bilayer diffusion of drugs into cells is negligible?, Trends Pharmacol. Sci., 2015 DOI:10.1016/j.tips.2014.10.005.
  358. D. B. Kell and S. G. Oliver, Here is the evidence, now what is the hypothesis? The complementary roles of inductive and hypothesis-driven science in the post-genomic era., BioEssays, 2004, 26, 99–105 CrossRef PubMed.
  359. D. B. Kell, Metabolomics and systems biology: making sense of the soup, Curr. Opin. Microbiol., 2004, 7, 296–307 CrossRef CAS PubMed.
  360. D. B. Kell and J. D. Knowles, The role of modeling in systems biology, in System modeling in cellular biology: from concepts to nuts and bolts, ed. Z. Szallasi, J. Stelling and V. Periwal, MIT Press, Cambridge, 2006, pp. 3–18 Search PubMed.
  361. D. B. Kell, Metabolomics, modelling and machine learning in systems biology: towards an understanding of the languages of cells. The 2005 Theodor Bücher lecture, FEBS J., 2006, 273, 873–894 CrossRef CAS PubMed.
  362. D. B. Kell, Finding novel pharmaceuticals in the systems biology era using multiple effective drug targets, phenotypic screening, and knowledge of transporters: where drug discovery went wrong and how to fix it, FEBS J., 2013, 280, 5957–5980 CrossRef CAS PubMed.
  363. L. R. Franklin, Exploratory experiments, Philos. Sci., 2005, 72, 888–899 CrossRef.
  364. K. C. Elliott, Epistemic and methodological iteration in scientific research, Stud. Hist. Philos Sci., 2012, 43, 376–382 CrossRef PubMed.
  365. S. A. Doyle, S. Y. Fung and D. E. Koshland, Jr., Redesigning the substrate specificity of an enzyme: isocitrate dehydrogenase, Biochemistry, 2000, 39, 14348–14355 CrossRef CAS PubMed.
  366. Q. S. Li, U. Schwaneberg, M. Fischer, J. Schmitt, J. Pleiss, S. Lutz-Wahl and R. D. Schmid, Rational evolution of a medium chain-specific cytochrome P-450 BM-3 variant, Biochim. Biophys. Acta, 2001, 1545, 114–121 CrossRef CAS.
  367. C. Gustafsson, S. Govindarajan and J. Minshull, Putting engineering back into protein engineering: bioinformatic approaches to catalyst design, Curr. Opin. Biotechnol., 2003, 14, 366–370 CrossRef CAS.
  368. G. Jiménez-Osés, S. Osuna, X. Gao, M. R. Sawaya, L. Gilson, S. J. Collier, G. W. Huisman, T. O. Yeates, Y. Tang and K. N. Houk, The role of distant mutations and allosteric regulation on LovD active site dynamics, Nat. Chem. Biol., 2014, 10, 431–436 CrossRef PubMed.
  369. P. Tian, Computational protein design, from single domain soluble proteins to membrane proteins, Chem. Soc. Rev., 2010, 39, 2071–2082 RSC.
  370. B. R. Lichtenstein, T. A. Farid, G. Kodali, L. A. Solomon, J. L. Anderson, M. M. Sheehan, N. M. Ennist, B. A. Fry, S. E. Chobot, C. Bialas, J. A. Mancini, C. T. Armstrong, Z. Zhao, T. V. Esipova, D. Snell, S. A. Vinogradov, B. M. Discher, C. C. Moser and P. L. Dutton, Engineering oxidoreductases: maquette proteins designed from scratch, Biochem. Soc. Trans., 2012, 40, 561–566 CrossRef CAS PubMed.
  371. T. A. Farid, G. Kodali, L. A. Solomon, B. R. Lichtenstein, M. M. Sheehan, B. A. Fry, C. Bialas, N. M. Ennist, J. A. Siedlecki, Z. Zhao, M. A. Stetz, K. G. Valentine, J. L. Anderson, A. J. Wand, B. M. Discher, C. C. Moser and P. L. Dutton, Elementary tetrahelical protein design for diverse oxidoreductase functions, Nat. Chem. Biol., 2013, 9, 826–833 CrossRef CAS PubMed.
  372. C. W. Wood, M. Bruning, A. A. Ibarra, G. J. Bartlett, A. R. Thomson, R. B. Sessions, R. L. Brady and D. N. Woolfson, CCBuilder: an interactive web-based tool for building, designing and assessing coiled-coil protein assemblies, Bioinformatics, 2014, 30, 3029–3035 CrossRef PubMed.
  373. A. R. Thomson, C. W. Wood, A. J. Burton, G. J. Bartlett, R. B. Sessions, R. L. Brady and D. N. Woolfson, Computational design of water-soluble alpha-helical barrels, Science, 2014, 346, 485–488 CrossRef CAS PubMed.
  374. F. Richter, A. Leaver-Fay, S. D. Khare, S. Bjelic and D. Baker, De novo enzyme design using Rosetta3, PLoS One, 2011, 6, e19230 CAS.
  375. L. Wang, E. A. Althoff, J. Bolduc, L. Jiang, J. Moody, J. K. Lassila, L. Giger, D. Hilvert, B. Stoddard and D. Baker, Structural analyses of covalent enzyme-substrate analog complexes reveal strengths and limitations of de novo enzyme design, J. Mol. Biol., 2012, 415, 615–625 CrossRef CAS PubMed.
  376. L. Jiang, E. A. Althoff, F. R. Clemente, L. Doyle, D. Rothlisberger, A. Zanghellini, J. L. Gallaher, J. L. Betker, F. Tanaka, C. F. Barbas, 3rd, D. Hilvert, K. N. Houk, B. L. Stoddard and D. Baker, De novo computational design of retro-aldol enzymes, Science, 2008, 319, 1387–1391 CrossRef CAS PubMed.
  377. D. Röthlisberger, O. Khersonsky, A. M. Wollacott, L. Jiang, J. DeChancie, J. Betker, J. L. Gallaher, E. A. Althoff, A. Zanghellini, O. Dym, S. Albeck, K. N. Houk, D. S. Tawfik and D. Baker, Kemp elimination catalysts by computational enzyme design, Nature, 2008, 453, 190–195 CrossRef PubMed.
  378. J. B. Siegel, A. Zanghellini, H. M. Lovick, G. Kiss, A. R. Lambert, J. L. St Clair, J. L. Gallaher, D. Hilvert, M. H. Gelb, B. L. Stoddard, K. N. Houk, F. E. Michael and D. Baker, Computational design of an enzyme catalyst for a stereoselective bimolecular Diels-Alder reaction, Science, 2010, 329, 309–313 CrossRef CAS PubMed.
  379. J. Sykora, J. Brezovsky, T. Koudelakova, M. Lahoda, A. Fortova, T. Chernovets, R. Chaloupkova, V. Stepankova, Z. Prokop, I. K. Smatanova, M. Hof and J. Damborsky, Dynamics and hydration explain failed functional transformation in dehalogenase design, Nat. Chem. Biol., 2014, 10, 428–430 CrossRef CAS PubMed.
  380. C. B. Eiben, J. B. Siegel, J. B. Bale, S. Cooper, F. Khatib, B. W. Shen, F. Players, B. L. Stoddard, Z. Popovic and D. Baker, Increased Diels-Alderase activity through backbone remodeling guided by Foldit players, Nat. Biotechnol., 2012, 30, 190–192 CrossRef CAS PubMed.
  381. Y. Kipnis and D. Baker, Comparison of designed and randomly generated catalysts for simple chemical reactions, Protein Sci., 2012, 21, 1388–1395 CrossRef CAS PubMed.
  382. Y. L. Boersma and A. Pluckthun, DARPins and other repeat protein scaffolds: advances in engineering and applications, Curr. Opin. Biotechnol., 2011, 22, 849–857 CrossRef CAS PubMed.
  383. W. J. Albery and J. R. Knowles, Evolution of enzyme function and the development of catalytic efficiency, Biochemistry, 1976, 15, 5631–5640 CrossRef CAS.
  384. E. B. Nickbarg and J. R. Knowles, Triosephosphate isomerase: energetics of the reaction catalyzed by the yeast enzyme expressed in Escherichia coli, Biochemistry, 1988, 27, 5939–5947 CrossRef CAS.
  385. F. Zárate-Pérez, M. E. Chánez-Cárdenas and E. Vázquez-Contreras, The folding pathway of triosephosphate isomerase, Prog. Mol. Biol. Transl. Sci., 2008, 84, 251–267 Search PubMed.
  386. B. J. Sullivan, V. Durani and T. J. Magliery, Triosephosphate isomerase by consensus design: dramatic differences in physical properties and activity of related variants, J. Mol. Biol., 2011, 413, 195–208 CrossRef CAS PubMed.
  387. M. Alahuhta, M. Salin, M. G. Casteleijn, C. Kemmer, I. El-Sayed, K. Augustyns, P. Neubauer and R. K. Wierenga, Structure-based protein engineering efforts with a monomeric TIM variant: the importance of a single point mutation for generating an active site with suitable binding properties, Protein Eng., Des. Sel., 2008, 21, 257–266 CrossRef CAS PubMed.
  388. T. V. Borchert, R. Abagyan, R. Jaenicke and R. K. Wierenga, Design, creation, and characterization of a stable, monomeric triosephosphate isomerase, Proc. Natl. Acad. Sci. U. S. A., 1994, 91, 1515–1518 CrossRef CAS.
  389. M. Holmquist, Alpha/Beta-hydrolase fold enzymes: structures, functions and mechanisms, Curr. Protein Pept. Sci., 2000, 1, 209–235 CrossRef CAS.
  390. M. Henn-Sax, B. Hocker, M. Wilmanns and R. Sterner, Divergent evolution of (betaalpha)8-barrel enzymes, Biol. Chem., 2001, 382, 1315–1320 CrossRef CAS PubMed.
  391. R. K. Wierenga, The TIM-barrel fold: a versatile framework for efficient enzymes, FEBS Lett., 2001, 492, 193–198 CrossRef CAS.
  392. N. Nagano, C. A. Orengo and J. M. Thornton, One fold with many functions: the evolutionary relationships between TIM barrel families based on their sequences, structures and functions, J. Mol. Biol., 2002, 321, 741–765 CrossRef CAS.
  393. D. M. Z. Schmidt, E. C. Mundorff, M. Dojka, E. Bermudez, J. E. Ness, S. Govindarajan, P. C. Babbitt, J. Minshull and J. A. Gerlt, Evolutionary potential of (beta/alpha)8-barrels: functional promiscuity produced by single substitutions in the enolase superfamily, Biochemistry, 2003, 42, 8387–8393 CrossRef CAS PubMed.
  394. J. E. Vick, D. M. Schmidt and J. A. Gerlt, Evolutionary potential of (beta/alpha)8-barrels: in vitro enhancement of a “new” reaction in the enolase superfamily, Biochemistry, 2005, 44, 11722–11729 CrossRef CAS PubMed.
  395. H. S. Park, S. H. Nam, J. K. Lee, C. N. Yoon, B. Mannervik, S. J. Benkovic and H. S. Kim, Design and evolution of new catalytic activity with an existing protein scaffold, Science, 2006, 311, 535–538 CrossRef CAS PubMed.
  396. J. E. Vick and J. A. Gerlt, Evolutionary potential of (beta/alpha)8-barrels: stepwise evolution of a “new” reaction in the enolase superfamily, Biochemistry, 2007, 46, 14589–14597 CrossRef CAS PubMed.
  397. S. Leopoldseder, J. Claren, C. Jurgens and R. Sterner, Interconverting the catalytic activities of (betaalpha)(8)-barrel enzymes from different metabolic pathways: sequence requirements and molecular analysis, J. Mol. Biol., 2004, 337, 871–879 CrossRef CAS PubMed.
  398. R. Sterner and B. Hocker, Catalytic versatility, stability, and evolution of the (betaalpha)8-barrel enzyme fold, Chem. Rev., 2005, 105, 4038–4055 CrossRef CAS PubMed.
  399. T. Seitz, M. Bocola, J. Claren and R. Sterner, Stabilisation of a (betaalpha)8-barrel protein designed from identical half barrels, J. Mol. Biol., 2007, 372, 114–129 CrossRef CAS PubMed.
  400. J. Claren, C. Malisi, B. Höcker and R. Sterner, Establishing wild-type levels of catalytic activity on natural and artificial (beta alpha)8-barrel protein scaffolds, Proc. Natl. Acad. Sci. U. S. A., 2009, 106, 3704–3709 CrossRef CAS PubMed.
  401. B. Höcker, A. Lochner, T. Seitz, J. Claren and R. Sterner, High-resolution crystal structure of an artificial (betaalpha)(8)-barrel protein designed from identical half-barrels, Biochemistry, 2009, 48, 1145–1147 CrossRef PubMed.
  402. X. Yang, S. V. Kathuria, R. Vadrevu and C. R. Matthews, Betaalpha-hairpin clamps brace betaalphabeta modules and can make substantive contributions to the stability of TIM barrel proteins, PLoS One, 2009, 4, e7179 Search PubMed.
  403. J. A. Gerlt, New wine from old barrels, Nat. Struct. Biol., 2000, 7, 171–173 CrossRef CAS PubMed.
  404. T. A. Bharat, S. Eisenbeis, K. Zeth and B. Höcker, A beta alpha-barrel built by the combination of fragments from different folds, Proc. Natl. Acad. Sci. U. S. A., 2008, 105, 9942–9947 CrossRef CAS PubMed.
  405. B. Höcker, Directed evolution of (betaalpha)(8)-barrel enzymes, Biomol. Eng., 2005, 22, 31–38 CrossRef PubMed.
  406. A. Fischer, T. Seitz, A. Lochner, R. Sterner, R. Merkl and M. Bocola, A fast and precise approach for computational saturation mutagenesis and its experimental validation by using an artificial (betaalpha)8-barrel protein, ChemBioChem, 2011, 12, 1544–1550 CrossRef CAS PubMed.
  407. S. Eisenbeis, W. Proffitt, M. Coles, V. Truffault, S. Shanmugaratnam, J. Meiler and B. Höcker, Potential of Fragment Recombination for Rational Design of Proteins, J. Am. Chem. Soc., 2012, 134, 4019–4022 CrossRef CAS PubMed.
  408. X. Deng, J. Lee, A. J. Michael, D. R. Tomchick, E. J. Goldsmith and M. A. Phillips, Evolution of substrate specificity within a diverse family of beta/alpha-barrel-fold basic amino acid decarboxylases: X-ray structure determination of enzymes with specificity for L-arginine and carboxynorspermidine, J. Biol. Chem., 2010, 285, 25708–25719 CrossRef CAS PubMed.
  409. S. Deechongkit, H. Nguyen, M. Jager, E. T. Powers, M. Gruebele and J. W. Kelly, Beta-sheet folding mechanisms from perturbation energetics, Curr. Opin. Struct. Biol., 2006, 16, 94–101 CrossRef CAS PubMed.
  410. W. R. Forsyth, O. Bilsel, Z. Gu and C. R. Matthews, Topology and sequence in the folding of a TIM barrel protein: global analysis highlights partitioning between transient off-pathway and stable on-pathway folding intermediates in the complex folding mechanism of a (betaalpha)8 barrel of unknown function from B. subtilis, J. Mol. Biol., 2007, 372, 236–253 CrossRef CAS PubMed.
  411. Z. Gu, M. K. Rao, W. R. Forsyth, J. M. Finke and C. R. Matthews, Structural analysis of kinetic folding intermediates for a TIM barrel protein, indole-3-glycerol phosphate synthase, by hydrogen exchange mass spectrometry and Go model simulation, J. Mol. Biol., 2007, 374, 528–546 CrossRef CAS PubMed.
  412. C. Kalyanaraman, K. Bernacki and M. P. Jacobson, Virtual screening against highly charged active sites: identifying substrates of alpha-beta barrel enzymes, Biochemistry, 2005, 44, 2059–2071 CrossRef CAS PubMed.
  413. A. Sakai, A. A. Fedorov, E. V. Fedorov, A. M. Schnoes, M. E. Glasner, S. Brown, M. E. Rutter, K. Bain, S. Chang, T. Gheyi, J. M. Sauder, S. K. Burley, P. C. Babbitt, S. C. Almo and J. A. Gerlt, Evolution of enzymatic activities in the enolase superfamily: stereochemically distinct mechanisms in two families of cis,cis-muconate lactonizing enzymes, Biochemistry, 2009, 48, 1445–1453 CrossRef CAS PubMed.
  414. R. Kourist, H. Jochens, S. Bartsch, R. Kuipers, S. K. Padhi, M. Gall, D. Böttcher, H. J. Joosten and U. T. Bornscheuer, The alpha/beta-hydrolase fold 3DM database (ABHDB) as a tool for protein engineering, ChemBioChem, 2010, 11, 1635–1643 CrossRef CAS PubMed.
  415. H. Jochens, M. Hesseler, K. Stiba, S. K. Padhi, R. J. Kazlauskas and U. T. Bornscheuer, Protein engineering of alpha/beta-hydrolase fold enzymes, ChemBioChem, 2011, 12, 1508–1517 CrossRef CAS PubMed.
  416. J. A. Gerlt, P. C. Babbitt, M. P. Jacobson and S. C. Almo, Divergent evolution in enolase superfamily: strategies for assigning functions, J. Biol. Chem., 2012, 287, 29–34 CrossRef CAS PubMed.
  417. T. Lukk, A. Sakai, C. Kalyanaraman, S. D. Brown, H. J. Imker, L. Song, A. A. Fedorov, E. V. Fedorov, R. Toro, B. Hillerich, R. Seidel, Y. Patskovsky, M. W. Vetting, S. K. Nair, P. C. Babbitt, S. C. Almo, J. A. Gerlt and M. P. Jacobson, Homology models guide discovery of diverse enzyme specificities among dipeptide epimerases in the enolase superfamily, Proc. Natl. Acad. Sci. U. S. A., 2012, 109, 4122–4127 CrossRef CAS PubMed.
  418. A. Zanghellini, L. Jiang, A. M. Wollacott, G. Cheng, J. Meiler, E. A. Althoff, D. Rothlisberger and D. Baker, New algorithms and an in silico benchmark for computational enzyme design, Protein Sci., 2006, 15, 2785–2794 CrossRef CAS PubMed.
  419. C. Malisi, O. Kohlbacher and B. Höcker, Automated scaffold selection for enzyme design, Proteins, 2009, 77, 74–83 CrossRef CAS PubMed.
  420. E. Dellus-Gur, A. Toth-Petroczy, M. Elias and D. S. Tawfik, What makes a protein fold amenable to functional innovation? Fold polarity and stability trade-offs, J. Mol. Biol., 2013, 425, 2609–2621 CrossRef CAS PubMed.
  421. M. Gebauer and A. Skerra, Anticalins small engineered binding proteins based on the lipocalin scaffold, Methods Enzymol., 2012, 503, 157–188 CAS.
  422. M. Gebauer, A. Schiefner, G. Matschiner and A. Skerra, Combinatorial design of an Anticalin directed against the extra-domain B for the specific targeting of oncofetal fibronectin, J. Mol. Biol., 2012, 425, 780–802 CrossRef PubMed.
  423. A. M. Hohlbaum and A. Skerra, Anticalins: the lipocalin family as a novel protein scaffold for the development of next-generation immunotherapies, Expert Rev. Clin. Immunol., 2007, 3, 491–501 CrossRef CAS PubMed.
  424. S. Schlehuber and A. Skerra, Anticalins as an alternative to antibody technology, Expert Opin. Biol. Ther., 2005, 5, 1453–1462 CrossRef CAS PubMed.
  425. A. Skerra, Anticalins as alternative binding proteins for therapeutic use, Curr. Opin. Mol. Ther., 2007, 9, 336–344 CAS.
  426. A. Skerra, Alternative binding proteins: anticalins - harnessing the structural plasticity of the lipocalin ligand pocket to engineer novel binding activities, FEBS J., 2008, 275, 2677–2683 CrossRef CAS PubMed.
  427. E. Gunneriusson, K. Nord, M. Uhlen and P. A. Nygren, Affinity maturation of a Taq DNA polymerase specific affibody by helix shuffling, Environ. Prot. Eng., 1999, 12, 873–878 CrossRef CAS PubMed.
  428. E. Gunneriusson, P. Samuelson, J. Ringdahl, H. Gronlund, P. A. Nygren and S. Stahl, Staphylococcal surface display of immunoglobulin A (IgA)- and IgE-specific in vitro-selected binding proteins (affibodies) based on Staphylococcus aureus protein A, Appl. Environ. Microbiol., 1999, 65, 4134–4140 CAS.
  429. G. Kronvall and K. Jonsson, Receptins: a novel term for an expanding spectrum of natural and engineered microbial proteins with binding properties for mammalian proteins, J. Mol. Recognit., 1999, 12, 38–44 CrossRef CAS.
  430. K. Nord, E. Gunneriusson, J. Ringdahl, S. Stahl, M. Uhlen and P. A. Nygren, Binding proteins selected from combinatorial libraries of an alpha-helical bacterial receptor domain, Nat. Biotechnol., 1997, 15, 772–777 CrossRef CAS PubMed.
  431. K. Nord, O. Nord, M. Uhlen, B. Kelley, C. Ljungqvist and P. A. Nygren, Recombinant human factor VIII-specific affinity ligands selected from phage-displayed combinatorial libraries of protein A, Eur. J. Biochem., 2001, 268, 4269–4277 CrossRef CAS.
  432. J. Feldwisch and V. Tolmachev, Engineering of affibody molecules for therapy and diagnostics, Methods Mol. Biol., 2012, 899, 103–126 CAS.
  433. J. Löfblom, J. Feldwisch, V. Tolmachev, J. Carlsson, S. Ståhl and F. Y. Frejd, Affibody molecules: engineered proteins for therapeutic, diagnostic and biotechnological applications, FEBS Lett., 2010, 584, 2670–2680 CrossRef PubMed.
  434. P.-Å. Nygren, Alternative binding proteins: affibody binding proteins developed from a small three-helix bundle scaffold, FEBS J., 2008, 275, 2668–2676 CrossRef CAS PubMed.
  435. A. Orlova, J. Feldwisch, L. Abrahmsén and V. Tolmachev, Update: affibody molecules for molecular imaging and therapy for cancer, Cancer Biother. Radiopharm., 2007, 22, 573–584 CrossRef CAS PubMed.
  436. V. Tolmachev, A. Orlova, F. Y. Nilsson, J. Feldwisch, A. Wennborg and L. Abrahmsén, Affibody molecules: potential for in vivo imaging of molecular targets for cancer therapy, Expert Opin. Biol. Ther., 2007, 7, 555–568 CrossRef CAS PubMed.
  437. L. A. Fernández, Prokaryotic expression of antibodies and affibodies, Curr. Opin. Biotechnol., 2004, 15, 364–373 CrossRef PubMed.
  438. R. Sterner and F. X. Schmid, De novo design of an enzyme, Science, 2004, 304, 1916–1917 CrossRef CAS PubMed.
  439. G. Cheng, B. Qian, R. Samudrala and D. Baker, Improvement in protein functional site prediction by distinguishing structural and functional constraints on protein family evolution using computational design, Nucleic Acids Res., 2005, 33, 5861–5867 CrossRef CAS PubMed.
  440. R. A. Chica, N. Doucet and J. N. Pelletier, Semi-rational approaches to engineering enzyme activity: combining the benefits of directed evolution and rational design, Curr. Opin. Biotechnol., 2005, 16, 378–384 CrossRef CAS PubMed.
  441. C. T. Saunders and D. Baker, Recapitulation of protein family divergence using flexible backbone protein design, J. Mol. Biol., 2005, 346, 631–644 CrossRef CAS PubMed.
  442. C. A. Floudas, H. K. Fung, S. R. McAllister, M. Monnigmann and R. Rajgaria, Advances in protein structure prediction and de novo protein design: A review, Chem. Eng. Sci., 2006, 61, 966–988 CrossRef CAS PubMed.
  443. O. Alvizo, B. D. Allen and S. L. Mayo, Computational protein design promises to revolutionize protein engineering, BioTechniques, 2007, 42, 31 CrossRef CAS PubMed 33, 35 passim.
  444. J. L. Anderson, R. L. Koder, C. C. Moser and P. L. Dutton, Controlling complexity and water penetration in functional de novo protein design, Biochem. Soc. Trans., 2008, 36, 1106–1111 CrossRef CAS PubMed.
  445. T. R. Ward, Artificial enzymes made to order: combination of computational design and directed evolution, Angew. Chem., Int. Ed., 2008, 47, 7802–7803 CrossRef CAS PubMed.
  446. M. L. Bellows and C. A. Floudas, Computational methods for de novo protein design and its applications to the human immunodeficiency virus 1, purine nucleoside phosphorylase, ubiquitin specific protease 7, and, histone demethylases, Curr. Drug Targets, 2010, 11, 264–278 CrossRef CAS.
  447. H. C. Fry, A. Lehmann, J. G. Saven, W. F. DeGrado and M. J. Therien, Computational design and elaboration of a de novo heterotetrameric alpha-helical protein that selectively binds an emissive abiological (porphinato)zinc chromophore, J. Am. Chem. Soc., 2010, 132, 3997–4005 CrossRef CAS PubMed.
  448. S. Lutz, Beyond directed evolution–semi-rational protein engineering and design, Curr. Opin. Biotechnol., 2010, 21, 734–743 CrossRef CAS PubMed.
  449. R. J. Pantazes, M. J. Grisewood and C. D. Maranas, Recent advances in computational protein design, Curr. Opin. Struct. Biol., 2011, 21, 467–472 CrossRef CAS PubMed.
  450. J. Pleiss, Protein design in metabolic engineering and synthetic biology, Curr. Opin. Biotechnol., 2011, 22, 611–617 CrossRef CAS PubMed.
  451. I. Samish, C. M. MacDermaid, J. M. Perez-Aguilar and J. G. Saven, Theoretical and computational protein design, Annu. Rev. Phys. Chem., 2011, 62, 129–149 CrossRef CAS PubMed.
  452. D. Hilvert, Design of protein catalysts, Annu. Rev. Biochem., 2013, 82, 447–470 CrossRef CAS PubMed.
  453. H. Kries, R. Blomberg and D. Hilvert, De novo enzymes by computational design, Curr. Opin. Chem. Biol., 2013, 17, 221–228 CrossRef CAS PubMed.
  454. H. K. Privett, G. Kiss, T. M. Lee, R. Blomberg, R. A. Chica, L. M. Thomas, D. Hilvert, K. N. Houk and S. L. Mayo, Iterative approach to computational enzyme design, Proc. Natl. Acad. Sci. U. S. A., 2012, 109, 3790–3795 CrossRef CAS PubMed.
  455. J. G. Saven, Computational protein design: engineering molecular diversity, nonnatural enzymes, nonbiological cofactor complexes, and membrane proteins, Curr. Opin. Chem. Biol., 2011, 15, 452–457 CrossRef CAS PubMed.
  456. P. S. Huang, G. Oberdorfer, C. Xu, X. Y. Pei, B. L. Nannenga, J. M. Rogers, F. DiMaio, T. Gonen, B. Luisi and D. Baker, High thermodynamic stability of parametrically designed helical bundles, Science, 2014, 346, 481–485 CrossRef CAS PubMed.
  457. B. A. Smith and M. H. Hecht, Novel proteins: from fold to function, Curr. Opin. Chem. Biol., 2011, 15, 421–426 CrossRef CAS PubMed.
  458. A. Bhattacherjee and P. Biswas, Combinatorial design of protein sequences with applications to lattice and real proteins, J. Chem. Phys., 2009, 131, 125101 CrossRef PubMed.
  459. C. L. Kleinman, N. Rodrigue, C. Bonnard, H. Philippe and N. Lartillot, A maximum likelihood framework for protein design, BMC Bioinf., 2006, 7, 326 CrossRef PubMed.
  460. S. D. Khare, Y. Kipnis, P. Greisen Jr, R. Takeuchi, Y. Ashani, M. Goldsmith, Y. Song, J. L. Gallaher, I. Silman, H. Leader, J. L. Sussman, B. L. Stoddard, D. S. Tawfik and D. Baker, Computational redesign of a mononuclear zinc metalloenzyme for organophosphate hydrolysis, Nat. Chem. Biol., 2012, 8, 294–300 CrossRef CAS PubMed.
  461. B. Höcker, A metalloenzyme reloaded, Nat. Chem. Biol., 2012, 8, 224–225 CrossRef PubMed.
  462. E. A. Althoff, L. Wang, L. Jiang, L. Giger, J. K. Lassila, Z. Wang, M. Smith, S. Hari, P. Kast, D. Herschlag, D. Hilvert and D. Baker, Robust design and optimization of retroaldol enzymes, Protein Sci., 2012, 21, 717–726 CrossRef CAS PubMed.
  463. L. Giger, S. Caner, R. Obexer, P. Kast, D. Baker, N. Ban and D. Hilvert, Evolution of a designed retro-aldolase leads to complete active site remodeling, Nat. Chem. Biol., 2013, 9, 494–498 CrossRef CAS PubMed.
  464. K. Feldmeier and B. Höcker, Computational protein design of ligand binding and catalysis, Curr. Opin. Chem. Biol., 2013, 17, 929–933 CrossRef CAS PubMed.
  465. S. Božič, T. Doles, H. Gradišar and R. Jerala, New designed protein assemblies, Curr. Opin. Chem. Biol., 2013, 17, 940–945 CrossRef PubMed.
  466. J. Simms and P. J. Booth, Membrane proteins by accident or design, Curr. Opin. Chem. Biol., 2013, 17, 976–981 CrossRef CAS PubMed.
  467. M. A. Hallen, D. A. Keedy and B. R. Donald, Dead-end elimination with perturbations (DEEPer): a provable protein design algorithm with continuous sidechain and backbone flexibility, Proteins, 2013, 81, 18–39 CrossRef CAS PubMed.
  468. G. Kiss, N. Çelebi-ölçüm, R. Moretti, D. Baker and K. N. Houk, Computational enzyme design, Angew. Chem., Int. Ed., 2013, 52, 5700–5725 CrossRef CAS PubMed.
  469. D. J. Tantillo, How an enzyme might accelerate an intramolecular Diels-Alder reaction: theozymes for the formation of salvileucalin B, Org. Lett., 2010, 12, 1164–1167 CrossRef CAS PubMed.
  470. X. Zhang, J. DeChancie, H. Gunaydin, A. B. Chowdry, F. R. Clemente, A. J. Smith, T. M. Handel and K. N. Houk, Quantum mechanical design of enzyme active sites, J. Org. Chem., 2008, 73, 889–899 CrossRef CAS PubMed.
  471. J. Dechancie, F. R. Clemente, A. J. Smith, H. Gunaydin, Y. L. Zhao, X. Zhang and K. N. Houk, How similar are enzyme active site geometries derived from quantum mechanical theozymes to crystal structures of enzyme–inhibitor complexes? Implications for enzyme design, Protein Sci., 2007, 16, 1851–1866 CrossRef CAS PubMed.
  472. D. J. Tantillo, J. Chen and K. N. Houk, Theozymes and compuzymes: theoretical models for biological catalysis, Curr. Opin. Chem. Biol., 1998, 2, 743–750 CrossRef CAS.
  473. G. Bouvignies, P. Vallurupalli, D. F. Hansen, B. E. Correia, O. Lange, A. Bah, R. M. Vernon, F. W. Dahlquist, D. Baker and L. E. Kay, Solution structure of a minor and transiently formed state of a T4 lysozyme mutant, Nature, 2011, 477, 111–114 CrossRef CAS PubMed.
  474. S. Cooper, F. Khatib, A. Treuille, J. Barbero, J. Lee, M. Beenen, A. Leaver-Fay, D. Baker, Z. Popović and F. Players, Predicting protein structures with a multiplayer online game, Nature, 2010, 466, 756–760 CrossRef CAS PubMed.
  475. F. DiMaio, A. Leaver-Fay, P. Bradley, D. Baker and I. Andre, Modeling symmetric macromolecular structures in Rosetta3, PLoS One, 2011, 6, e20450 CAS.
  476. S. J. Fleishman, A. Leaver-Fay, J. E. Corn, E. M. Strauch, S. D. Khare, N. Koga, J. Ashworth, P. Murphy, F. Richter, G. Lemmon, J. Meiler and D. Baker, RosettaScripts: a scripting language interface to the Rosetta macromolecular modeling suite, PLoS One, 2011, 6, e20161 CAS.
  477. J. Handl, J. Knowles, R. Vernon, D. Baker and S. C. Lovell, The dual role of fragments in fragment-assembly methods for de novo protein structure prediction, Proteins, 2012, 80, 490–504 CrossRef CAS PubMed.
  478. A. Leaver-Fay, M. Tyka, S. M. Lewis, O. F. Lange, J. Thompson, R. Jacak, K. Kaufman, P. D. Renfrew, C. A. Smith, W. Sheffler, I. W. Davis, S. Cooper, A. Treuille, D. J. Mandell, F. Richter, Y. E. Ban, S. J. Fleishman, J. E. Corn, D. E. Kim, S. Lyskov, M. Berrondo, S. Mentzer, Z. Popovic, J. J. Havranek, J. Karanicolas, R. Das, J. Meiler, T. Kortemme, J. J. Gray, B. Kuhlman, D. Baker and P. Bradley, ROSETTA3: an object-oriented software suite for the simulation and design of macromolecules, Methods Enzymol., 2011, 487, 545–574 CAS.
  479. O. F. Lange, P. Rossi, N. G. Sgourakis, Y. Song, H. W. Lee, J. M. Aramini, A. Ertekin, R. Xiao, T. B. Acton, G. T. Montelione and D. Baker, Determination of solution structures of proteins up to 40 kDa using CS-Rosetta with sparse NMR data from deuterated samples, Proc. Natl. Acad. Sci. U. S. A., 2012, 109, 10873–10878 CrossRef CAS PubMed.
  480. T. C. Terwilliger, F. Dimaio, R. J. Read, D. Baker, G. Bunkóczi, P. D. Adams, R. W. Grosse-Kunstleve, P. V. Afonine and N. Echols, phenix.mr_rosetta: molecular replacement and model rebuilding with Phenix and Rosetta, J. Struct. Funct. Genomics, 2012, 13, 81–90 CrossRef CAS PubMed.
  481. C. Schmitz, R. Vernon, G. Otting, D. Baker and T. Huber, Protein structure determination from pseudocontact shifts using ROSETTA, J. Mol. Biol., 2012, 416, 668–677 CrossRef CAS PubMed.
  482. D. Gront, D. W. Kulp, R. M. Vernon, C. E. Strauss and D. Baker, Generalized fragment picking in Rosetta: design, protocols and applications, PLoS One, 2011, 6, e23294 CAS.
  483. J. Thompson and D. Baker, Incorporation of evolutionary information into Rosetta comparative modeling, Proteins, 2011, 79, 2380–2388 CrossRef CAS PubMed.
  484. F. Lauck, C. A. Smith, G. F. Friedland, E. L. Humphris and T. Kortemme, RosettaBackrub—a web server for flexible backbone protein structure modeling and design, Nucleic Acids Res., 2010, 38, W569–W575 CrossRef CAS PubMed.
  485. C. A. Smith and T. Kortemme, Predicting the tolerated sequences for proteins and protein interfaces using RosettaBackrub flexible backbone design, PLoS One, 2011, 6, e20451 CAS.
  486. E. H. C. Bromley, K. Channon, E. Moutevelis and D. N. Woolfson, Peptide and protein building blocks for synthetic biology: from programming biomolecules to self-organized biomolecular systems, ACS Chem. Biol., 2008, 3, 38–50 CrossRef CAS PubMed.
  487. C. T. Armstrong, A. L. Boyle, E. H. Bromley, Z. N. Mahmoud, L. Smith, A. R. Thomson and D. N. Woolfson, Rational design of peptide-based building blocks for nanoscience and synthetic biology, Faraday Discuss., 2009, 143, 305–317 RSC , discussion 359–372.
  488. E. H. C. Bromley, R. B. Sessions, A. R. Thomson and D. N. Woolfson, Designed alpha-helical tectons for constructing multicomponent synthetic biological systems, J. Am. Chem. Soc., 2009, 131, 928–930 CrossRef CAS PubMed.
  489. S. R. Gordon, E. J. Stanley, S. Wolf, A. Toland, S. J. Wu, D. Hadidi, J. H. Mills, D. Baker, I. S. Pultz and J. B. Siegel, Computational design of an alpha-gliadin peptidase, J. Am. Chem. Soc., 2012, 134, 20513–20520 CrossRef CAS PubMed.
  490. D. Bhattacharya and J. Cheng, 3Drefine: Consistent protein structure refinement by optimizing hydrogen bonding network and atomic-level energy minimization, Proteins, 2013, 81, 119–131 CrossRef CAS PubMed.
  491. J. A. Davey and R. A. Chica, Multistate approaches in computational protein design, Protein Sci., 2012, 21, 1241–1252 CrossRef CAS PubMed.
  492. K. J. M. Hanf, Protein design for diversity of sequences and conformations using dead-end elimination, Methods Mol. Biol., 2012, 899, 127–144 CAS.
  493. R. Kim and J. Skolnick, Assessment of programs for ligand binding affinity prediction, J. Comput. Chem., Jpn, 2008, 29, 1316–1331 CrossRef CAS PubMed.
  494. P. Cunningham, I. Afzal-Ahmed and R. J. Naftalin, Docking studies show thatD-glucose and quercetin slide through the transporter GLUT1, J. Biol. Chem., 2006, 281, 5797–5803 CrossRef CAS PubMed.
  495. C. Hetényi, U. Maran, A. T. García-Sosa and M. Karelson, Structure-based calculation of drug efficiency indices, Bioinformatics, 2007, 23, 2678–2685 CrossRef PubMed.
  496. S. Cosconati, S. Forli, A. L. Perryman, R. Harris, D. S. Goodsell and A. J. Olson, Virtual Screening with AutoDock: Theory and Practice, Expert Opin. Drug Discovery, 2010, 5, 597–607 CrossRef CAS PubMed.
  497. O. Trott and A. J. Olson, AutoDock Vina: improving the speed and accuracy of docking with a new scoring function, efficient optimization, and multithreading, J. Comput. Chem., 2010, 31, 455–461 CAS.
  498. G. Sandeep, K. P. Nagasree, M. Hanisha and M. M. Kumar, AUDocker LE: A GUI for virtual screening with AUTODOCK Vina, BMC Res. Notes, 2011, 4, 445 CrossRef PubMed.
  499. S. D. Handoko, X. Ouyang, C. T. Su, C. K. Kwoh and Y. S. Ong, QuickVina: accelerating AutoDock Vina using gradient-based heuristics for global optimization, IEEE/ACM Trans. Comput. Biol. Bioinf., 2012, 9, 1266–1272 CrossRef PubMed.
  500. M. Gao and J. Skolnick, APoc: large-scale identification of similar protein pockets, Bioinformatics, 2013, 29, 597–604 CrossRef CAS PubMed.
  501. M. P. Repasky, R. B. Murphy, J. L. Banks, J. R. Greenwood, I. Tubert-Brohman, S. Bhat and R. A. Friesner, Docking performance of the glide program as evaluated on the Astex and DUD datasets: a complete set of glide SP results and selected results for a new scoring function integrating WaterMap and glide, J. Comput.-Aided Mol. Des., 2012, 26, 787–799 CrossRef CAS PubMed.
  502. T. A. Halgren, R. B. Murphy, R. A. Friesner, H. S. Beard, L. L. Frye, W. T. Pollard and J. L. Banks, Glide: a new approach for rapid, accurate docking and scoring. 2. Enrichment factors in database screening, J. Med. Chem., 2004, 47, 1750–1759 CrossRef CAS PubMed.
  503. R. A. Friesner, J. L. Banks, R. B. Murphy, T. A. Halgren, J. J. Klicic, D. T. Mainz, M. P. Repasky, E. H. Knoll, M. Shelley, J. K. Perry, D. E. Shaw, P. Francis and P. S. Shenkin, Glide: a new approach for rapid, accurate docking and scoring. 1. Method and assessment of docking accuracy, J. Med. Chem., 2004, 47, 1739–1749 CrossRef CAS PubMed.
  504. K. T. Schomburg, I. Ardao, K. Gotz, F. Rieckenberg, A. Liese, A. P. Zeng and M. Rarey, Computational biotechnology: Prediction of competitive substrate inhibition of enzymes by buffer compounds with protein-ligand docking, J. Biotechnol., 2012, 161, 391–401 CrossRef CAS PubMed.
  505. M. Vass, A. Tarcsay and G. M. Keserü, Multiple ligand docking by Glide: implications for virtual second-site screening, J. Comput.-Aided Mol. Des., 2012, 26, 821–834 CrossRef CAS PubMed.
  506. P. A. Greenidge, C. Kramer, J. C. Mozziconacci and R. M. Wolf, MM/GBSA Binding Energy Prediction on the PDBbind Data Set: Successes, Failures, and Directions for Further Improvement, J. Chem. Inf. Model., 2012, 53, 201–209 CrossRef PubMed.
  507. D. Plewczynski, M. Laźniewski, R. Augustyniak and K. Ginalski, Can we trust docking results? Evaluation of seven commonly used programs on PDBbind database, J. Comput. Chem., Jpn, 2011, 32, 742–755 CrossRef CAS PubMed.
  508. R. Wang, X. Fang, Y. Lu, C. Y. Yang and S. Wang, The PDBbind database: methodologies and updates, J. Med. Chem., 2005, 48, 4111–4119 CrossRef CAS PubMed.
  509. Y. X. Yuan, J. F. Pei and L. H. Lai, LigBuilder 2: A practical de novo drug design approach, J. Chem. Inf. Model., 2011, 51, 1083–1091 CrossRef CAS PubMed.
  510. S. Sirin, R. Kumar, C. Martinez, M. J. Karmilowicz, P. Ghosh, Y. A. Abramov, V. Martin and W. Sherman, A Computational Approach to Enzyme Design: Predicting omega-Aminotransferase Catalytic Activity Using Docking and MM-GBSA Scoring, J. Chem. Inf. Model., 2014, 54, 2334–2346 CrossRef CAS PubMed.
  511. H. Y. Zhou and J. Skolnick, FINDSITEcomb: A Threading/Structure-Based, Proteomic-Scale Virtual Ligand Screening Approach, J. Chem. Inf. Model., 2013, 53, 230–240 CrossRef CAS PubMed.
  512. J. C. Faver, M. L. Benson, X. A. He, B. P. Roberts, B. Wang, M. S. Marshall, M. R. Kennedy, C. D. Sherrill and K. M. Merz, Formal Estimation of Errors in Computed Absolute Interaction Energies of Protein–Ligand Complexes, J. Chem. Theory Comput., 2011, 7, 790–797 CrossRef CAS PubMed.
  513. M. Goldsmith and D. S. Tawfik, Enzyme engineering by targeted libraries, Methods Enzymol., 2013, 523, 257–283 CAS.
  514. A. J. Ruff, A. Dennig and U. Schwaneberg, To get what we aim for: progress in diversity generation methods, FEBS J., 2013, 280, 2961–2978 CrossRef CAS PubMed.
  515. T. Zhang, Z. F. Wu, H. Chen, Q. Wu, Z. Z. Tang, J. B. Gou, L. H. Wang, W. W. Hao, C. M. Wang and C. M. Li, Progress in strategies for sequence diversity library creation for directed evolution, Afr. J. Biotechnol., 2010, 9, 9277–9285 Search PubMed.
  516. Directed Evolution Library Creation: Methods and Protocols, ed. E. M. J. Gillam, J. N. Copp and D. Ackerley, Springer, Berlin, 2014 Search PubMed.
  517. J. Damborsky and J. Brezovsky, Computational tools for designing and engineering biocatalysts, Curr. Opin. Chem. Biol., 2009, 13, 26–34 CrossRef CAS PubMed.
  518. J. Brezovsky, E. Chovancova, A. Gora, A. Pavelka, L. Biedermannova and J. Damborsky, Software tools for identification, visualization and analysis of protein tunnels and channels, Biotechnol. Adv., 2013, 31, 38–49 CrossRef CAS PubMed.
  519. E. Sebestova, J. Bendl, J. Brezovsky and J. Damborsky, Computational tools for designing smart libraries, Methods Mol. Biol., 2014, 1179, 291–314 Search PubMed.
  520. R. Krašovec, R. V. Belavkin, J. A. D. Aston, A. Channon, E. Aston, B. M. Rash, M. Kadirvel, S. Forbes and C. G. Knight, Mutation rate plasticity in rifampicin resistance depends on Escherichia coli cell-cell interactions, Nat. Commun., 2014, 5, 3742 CrossRef PubMed.
  521. M. Oates, D. Corne and R. Loader, Investigation of a characteristic bimodal convergence-time/mutation-rate feature in evolutionary search, Proc. Congr. Evol. Comput., IEEE, 1999, pp. 2175–2182 Search PubMed.
  522. M. Oates and D. W. Corne, Overcoming fitness barriers in multi-modal search spaces, in Foundation of Genetic Algorithms 6, ed. W. N. Martin and W. M. Spears, Academic Press, London, 2001, pp. 5–26 Search PubMed.
  523. M. J. Oates, D. Corne and R. Loader, Tri-phase performance profile of evolutionary search on uni- and multi-modal search spaces, in Proc. IEEE Congr. Evol. Computation, IEEE Neural Networks Council, San Diego, 2000, pp. 357–364 Search PubMed.
  524. M. J. Oates, D. W. Corne and D. B. Kell, The bimodal feature at large population sizes and high selection pressure: implications for directed evolution, in Recent advances in simulated evolution and learning, ed. K. C. Tan, M. H. Lim, X. Yao and L. Wang, World Scientific, Singapore, 2003, pp. 215–240 Search PubMed.
  525. M. Zaccolo and E. Gherardi, The effect of high-frequency random mutagenesis on in vitro protein evolution: A study on TEM-1 β-lactamase, J. Mol. Biol., 1999, 285, 775–783 CrossRef CAS PubMed.
  526. P. S. Daugherty, G. Chen, B. L. Iverson and G. Georgiou, Quantitative analysis of the effect of the mutation frequency on the affinity maturation of single chain Fv antibodies, Proc. Natl. Acad. Sci. U. S. A., 2000, 97, 2029–2034 CrossRef CAS PubMed.
  527. D. A. Drummond, B. L. Iverson, G. Georgiou and F. H. Arnold, Why high-error-rate random mutagenesis libraries are enriched in functional and improved proteins, J. Mol. Biol., 2005, 350, 806–816 CrossRef CAS PubMed.
  528. A. M. Leconte, B. C. Dickinson, D. D. Yang, I. A. Chen, B. Allen and D. R. Liu, A population-based experimental model for protein evolution: effects of mutation rate and selection stringency on evolutionary outcomes, Biochemistry, 2013, 52, 1490–1499 CrossRef CAS PubMed.
  529. D. W. Leung, E. Chen and D. V. Goeddel, A method for random mutagenesis of a defined DNA segment using a modified polymerase chain reaction., Technique, 1989, 1, 11–15 Search PubMed.
  530. E. O. McCullum, B. A. Williams, J. Zhang and J. C. Chaput, Random mutagenesis by error-prone PCR, Methods Mol. Biol., 2010, 634, 103–109 CAS.
  531. T. S. Wong, D. Roccatano, M. Zacharias and U. Schwaneberg, A statistical analysis of random mutagenesis methods used for directed protein evolution, J. Mol. Biol., 2006, 355, 858–871 CrossRef CAS PubMed.
  532. T. S. Wong, D. Roccatano and U. Schwaneberg, Challenges of the genetic code for exploring sequence space in directed protein evolution, Biocatal. Biotransform., 2007, 25, 229–241 CrossRef CAS.
  533. T. S. Rasila, M. I. Pajunen and H. Savilahti, Critical evaluation of random mutagenesis by error-prone polymerase chain reaction protocols, Escherichia coli mutator strain, and hydroxylamine treatment, Anal. Biochem., 2009, 388, 71–80 CrossRef CAS PubMed.
  534. J. Zhao, T. Kardashliev, A. Joelle Ruff, M. Bocola and U. Schwaneberg, Lessons from diversity of directed evolution experiments by an analysis of 3,000 mutations, Biotechnol. Bioeng., 2014, 111, 2380–2389 CrossRef CAS PubMed.
  535. R. C. Cadwell and G. F. Joyce, Randomization of genes by PCR mutagenesis, PCR Methods Appl., 1992, 2, 28–33 CrossRef CAS.
  536. M. Camps, J. Naukkarinen, B. P. Johnson and L. A. Loeb, Targeted gene evolution in Escherichia coli using a highly error-prone DNA polymerase I, Proc. Natl. Acad. Sci. U. S. A., 2003, 100, 9727–9732 CrossRef CAS PubMed.
  537. J. N. Copp, P. Hanson-Manful, D. F. Ackerley and W. M. Patrick, Error-prone PCR and effective generation of gene variant libraries for directed evolution, Methods Mol. Biol., 2014, 1179, 3–22 Search PubMed.
  538. I. Matsumura and L. A. Rowe, Whole plasmid mutagenic PCR for directed protein evolution, Biomol. Eng., 2005, 22, 73–79 CrossRef CAS PubMed.
  539. D. L. Alexander, J. Lilly, J. Hernandez, J. Romsdahl, C. J. Troll and M. Camps, Random mutagenesis by error-prone pol plasmid replication in Escherichia coli, Methods Mol. Biol., 2014, 1179, 31–44 Search PubMed.
  540. R. Fujii, M. Kitaoka and K. Hayashi, One-step random mutagenesis by error-prone rolling circle amplification, Nucleic Acids Res., 2004, 32, e145 CrossRef PubMed.
  541. R. Fujii, M. Kitaoka and K. Hayashi, RAISE: a simple and novel method of generating random insertion and deletion mutations, Nucleic Acids Res., 2006, 34, e30 CrossRef PubMed.
  542. Y. Kipnis, E. Dellus-Gur and D. S. Tawfik, TRINS: a method for gene modification by randomized tandem repeat insertions, Protein Eng., Des. Sel., 2012, 25, 437–444 CrossRef CAS PubMed.
  543. R. Fujii, M. Kitaoka and K. Hayashi, Random insertional-deletional strand exchange mutagenesis (RAISE): a simple method for generating random insertion and deletion mutations, Methods Mol. Biol., 2014, 1179, 151–158 Search PubMed.
  544. A. Ravikumar, A. Arrieta and C. C. Liu, An orthogonal DNA replication system in yeast, Nat. Chem. Biol., 2014, 10, 175–177 CrossRef CAS PubMed.
  545. L. Pritchard, D. W. Corne, D. B. Kell, J. J. Rowland and M. K. Winson, A general model of error-prone PCR, J. Theor. Biol., 2004, 234, 497–509 CrossRef PubMed.
  546. S. Hoebenreich, F. E. Zilly, C. G. Acevedo-Rocha, M. Zilly and M. T. Reetz, Speeding up Directed Evolution: Combining the Advantages of Solid-Phase Combinatorial Gene Synthesis with Statistically Guided Reduction of Screening Effort, ACS Synth. Biol., 2014 DOI:10.1021/sb5002399.
  547. D. Wei, M. Li, X. Zhang and L. Xing, An improvement of the site-directed mutagenesis method by combination of megaprimer, one-side PCR and DpnI treatment, Anal. Biochem., 2004, 331, 401–403 CrossRef CAS PubMed.
  548. D. L. Steffens and J. G. Williams, Efficient site-directed saturation mutagenesis using degenerate oligonucleotides, J. Biomol. Tech., 2007, 18, 147–149 Search PubMed.
  549. A. Fersht, Enzyme structure and mechanism, W.H. Freeman, San Francisco, 2nd edn, 1977 Search PubMed.
  550. J. Braman, C. Papworth and A. Greener, Site-directed mutagenesis using double-stranded plasmid DNA templates, Methods Mol. Biol., 1996, 57, 31–44 CAS.
  551. C. Papworth, J. C. Bauer, J. Braman and D. A. Wright, Site-directed mutagenesis in one day with >80% efficiency, Strategies, 1996, 9, 3–4 Search PubMed.
  552. L. Zheng, U. Baumann and J. L. Reymond, An efficient one-step site-directed and site-saturation mutagenesis protocol, Nucleic Acids Res., 2004, 32, e115 CrossRef PubMed.
  553. J. Sanchis, L. Fernandez, J. D. Carballeira, J. Drone, Y. Gumulya, H. Hobenreich, D. Kahakeaw, S. Kille, R. Lohmer, J. J. Peyralans, J. Podtetenieff, S. Prasad, P. Soni, A. Taglieber, S. Wu, F. E. Zilly and M. T. Reetz, Improved PCR method for the creation of saturation mutagenesis libraries in directed evolution: application to difficult-to-amplify templates, Appl. Microbiol. Biotechnol., 2008, 81, 387–397 CrossRef CAS PubMed.
  554. K. L. Morrison and G. A. Weiss, Combinatorial alanine-scanning, Curr. Opin. Chem. Biol., 2001, 5, 302–307 CrossRef CAS.
  555. G. A. Weiss, C. K. Watanabe, A. Zhong, A. Goddard and S. S. Sidhu, Rapid mapping of protein functional epitopes by combinatorial alanine scanning, Proc. Natl. Acad. Sci. U. S. A., 2000, 97, 8950–8954 CrossRef CAS PubMed.
  556. T. S. Wong, D. Roccatano and U. Schwaneberg, Steering directed protein evolution: strategies to manage combinatorial complexity of mutant libraries, Environ. Microbiol., 2007, 9, 2645–2659 CrossRef CAS PubMed.
  557. M. T. Reetz, L. W. Wang and M. Bocola, Directed evolution of enantioselective enzymes: iterative cycles of CASTing for probing protein-sequence space, Angew. Chem., Int. Ed., 2006, 45, 1236–1241 CrossRef CAS PubMed.
  558. R. D. Kirsch and E. Joly, An improved PCR-mutagenesis strategy for two-site mutagenesis or sequence swapping between related genes, Nucleic Acids Res., 1998, 26, 1848–1850 CrossRef CAS PubMed.
  559. L. W. Chiang, I. Kovari and M. M. Howe, Mutagenic oligonucleotide-directed PCR amplification (Mod-PCR): an efficient method for generating random base substitution mutations in a DNA sequence element, PCR Methods Appl., 1993, 2, 210–217 CrossRef CAS.
  560. S. N. Ho, H. D. Hunt, R. M. Horton, J. K. Pullen and L. R. Pease, Site-directed mutagenesis by overlap extension using the polymerase chain reaction, Gene, 1989, 77, 51–59 CrossRef CAS.
  561. R. M. Horton, H. D. Hunt, S. N. Ho, J. K. Pullen and L. R. Pease, Engineering hybrid genes without the use of restriction enzymes: gene splicing by overlap extension, Gene, 1989, 77, 61–68 CrossRef CAS.
  562. R. H. Peng, A. S. Xiong and Q. H. Yao, A direct and efficient PAGE-mediated overlap extension PCR method for gene multiple-site mutagenesis, Appl. Microbiol. Biotechnol., 2006, 73, 234–240 CrossRef CAS PubMed.
  563. K. L. Heckman and L. R. Pease, Gene splicing and mutagenesis by PCR-driven overlap extension, Nat. Protoc., 2007, 2, 924–932 CrossRef CAS PubMed.
  564. E. M. Williams, J. N. Copp and D. F. Ackerley, Site-saturation mutagenesis by overlap extension PCR, Methods Mol. Biol., 2014, 1179, 83–101 Search PubMed.
  565. T. S. Wong, K. L. Tee, B. Hauer and U. Schwaneberg, Sequence saturation mutagenesis (SeSaM): a novel method for directed evolution, Nucleic Acids Res., 2004, 32, e26 CrossRef PubMed.
  566. T. S. Wong, D. Roccatano, D. Loakes, K. L. Tee, A. Schenk, B. Hauer and U. Schwaneberg, Transversion-enriched sequence saturation mutagenesis (SeSaM-Tv+): a random mutagenesis method with consecutive nucleotide exchanges that complements the bias of error-prone PCR, Biotechnol. J., 2008, 3, 74–82 CrossRef CAS PubMed.
  567. H. Mundhada, J. Marienhagen, A. Scacioc, A. Schenk, D. Roccatano and U. Schwaneberg, SeSaM-Tv-II generates a protein sequence space that is unobtainable by epPCR, ChemBioChem, 2011, 12, 1595–1601 CrossRef CAS PubMed.
  568. A. J. Ruff, T. Kardashliev, A. Dennig and U. Schwaneberg, The Sequence Saturation Mutagenesis (SeSaM) method, Methods Mol. Biol., 2014, 1179, 45–68 Search PubMed.
  569. A. Dennig, A. V. Shivange, J. Marienhagen and U. Schwaneberg, OmniChange: the sequence independent method for simultaneous site-saturation of five codons, PLoS One, 2011, 6, e26222 CAS.
  570. A. Dennig, J. Marienhagen, A. J. Ruff and U. Schwaneberg, OmniChange: simultaneous site saturation of up to five codons, Methods Mol. Biol., 2014, 1179, 139–149 Search PubMed.
  571. A. V. Shivange, A. Dennig and U. Schwaneberg, Multi-site saturation by OmniChange yields a pH- and thermally improved phytase, J. Biotechnol., 2014, 170, 68–72 CrossRef CAS PubMed.
  572. C. G. Acevedo-Rocha, S. Hoebenreich and M. T. Reetz, Iterative saturation mutagenesis: a powerful approach to engineer proteins by systematically simulating Darwinian evolution, Methods Mol. Biol., 2014, 1179, 103–128 Search PubMed.
  573. L. P. Parra, R. Agudo and M. T. Reetz, Directed evolution by using iterative saturation mutagenesis based on multiresidue sites, ChemBioChem, 2013, 14, 2301–2309 CrossRef CAS PubMed.
  574. M. T. Reetz, J. D. Carballeira and A. Vogel, Iterative saturation mutagenesis on the basis of B factors as a strategy for increasing protein thermostability, Angew. Chem., Int. Ed., 2006, 45, 7745–7751 CrossRef CAS PubMed.
  575. M. T. Reetz and J. D. Carballeira, Iterative saturation mutagenesis (ISM) for rapid directed evolution of functional enzymes, Nat. Protoc., 2007, 2, 891–903 CrossRef CAS PubMed.
  576. M. T. Reetz, D. Kahakeaw and J. Sanchis, Shedding light on the efficacy of laboratory evolution based on iterative saturation mutagenesis, Mol. BioSyst., 2009, 5, 115–122 RSC.
  577. M. T. Reetz, P. Soni, L. Fernandez, Y. Gumulya and J. D. Carballeira, Increasing the stability of an enzyme toward hostile organic solvents by directed evolution based on iterative saturation mutagenesis using the B-FIT method, Chem. Commun., 2010, 46, 8657–8658 RSC.
  578. M. T. Reetz, S. Prasad, J. D. Carballeira, Y. Gumulya and M. Bocola, Iterative saturation mutagenesis accelerates laboratory evolution of enzyme stereoselectivity: rigorous comparison with traditional methods, J. Am. Chem. Soc., 2010, 132, 9144–9152 CrossRef CAS PubMed.
  579. H. Zheng and M. T. Reetz, Manipulating the stereoselectivity of limonene epoxide hydrolase by directed evolution based on iterative saturation mutagenesis, J. Am. Chem. Soc., 2010, 132, 15744–15751 CrossRef CAS PubMed.
  580. A. J. Baldwin, K. Busse, A. M. Simm and D. D. Jones, Expanded molecular diversity generation during directed evolution by trinucleotide exchange (TriNEx), Nucleic Acids Res., 2008, 36, e77 CrossRef PubMed.
  581. A. J. Baldwin, J. A. Arpino, W. R. Edwards, E. M. Tippmann and D. D. Jones, Expanded chemical diversity sampling through whole protein evolution, Mol. BioSyst., 2009, 5, 764–766 RSC.
  582. W. R. Edwards, K. Busse, R. K. Allemann and D. D. Jones, Linking the functions of unrelated proteins using a novel directed evolution domain insertion method, Nucleic Acids Res., 2008, 36, e78 CrossRef PubMed.
  583. D. D. Jones, J. A. J. Arpino, A. J. Baldwin and M. C. Edmundson, Transposon-based approaches for generating novel molecular diversity during directed evolution, Methods Mol. Biol., 2014, 1179, 159–172 Search PubMed.
  584. H. Liu and J. H. Naismith, An efficient one-step site-directed deletion, insertion, single and multiple-site plasmid mutagenesis protocol, BMC Biotechnol., 2008, 8, 91 CrossRef PubMed.
  585. A. Seyfang and J. H. Jin, Multiple site-directed mutagenesis of more than 10 sites simultaneously and in a single round, Anal. Biochem., 2004, 324, 285–291 CrossRef CAS PubMed.
  586. A. A. Fushan and D. T. Drayna, MALS: an efficient strategy for multiple site-directed mutagenesis employing a combination of DNA amplification, ligation and suppression PCR, BMC Biotechnol., 2009, 9, 83 CrossRef PubMed.
  587. D. M. Kegler-Ebo, C. M. Docktor and D. DiMaio, Codon cassette mutagenesis: a general method to insert or replace individual codons by using universal mutagenic cassettes, Nucleic Acids Res., 1994, 22, 1593–1599 CrossRef CAS PubMed.
  588. K. Steiner and H. Schwab, Recent advances in rational approaches for enzyme engineering, Comput. Struct. Biotechnol. J., 2012, 2, e201209010 Search PubMed.
  589. Y. Nov, Fitness loss and library size determination in saturation mutagenesis, PLoS One, 2013, 8, e68069 CAS.
  590. Nomenclature Committee of the International Union of Biochemistry (NC-IUB), Nomenclature for incompletely specified bases in nucleic acid sequences. Recommendations 1984, Eur. J. Biochem., 1985, 150, 1–5 CrossRef PubMed.
  591. M. A. Mena and P. S. Daugherty, Automated design of degenerate codon libraries, Protein Eng., Des. Sel., 2005, 18, 559–561 CrossRef CAS PubMed.
  592. L. Tang, X. Wang, B. Ru, H. Sun, J. Huang and H. Gao, MDC-Analyzer: A novel degenerate primer design tool for the construction of intelligent mutagenesis libraries with contiguous sites, BioTechniques, 2014, 56, 301–310 CAS.
  593. E. Muñoz and M. W. Deem, Amino acid alphabet size in protein evolution experiments: better to search a small library thoroughly or a large library sparsely?, Protein Eng., Des. Sel., 2008, 21, 311–317 CrossRef PubMed.
  594. K. L. Tee and T. S. Wong, Polishing the craft of genetic diversity creation in directed evolution, Biotechnol. Adv., 2013, 31, 1707–1721 CrossRef CAS PubMed.
  595. S. Kille, C. G. Acevedo-Rocha, L. P. Parra, Z. G. Zhang, D. J. Opperman, M. T. Reetz and J. P. Acevedo, Reducing Codon Redundancy and Screening Effort of Combinatorial Protein Libraries Created by Saturation Mutagenesis, ACS Synth. Biol., 2013, 2, 83–92 CrossRef CAS PubMed.
  596. M. D. Hughes, D. A. Nagel, A. F. Santos, A. J. Sutherland and A. V. Hine, Removing the redundancy from randomised gene libraries, J. Mol. Biol., 2003, 331, 973–979 CrossRef CAS.
  597. L. X. Tang, H. Gao, X. C. Zhu, X. Wang, M. Zhou and R. X. Jiang, Construction of “small-intelligent” focused mutagenesis libraries using well-designed combinatorial degenerate primers, BioTechniques, 2012, 52, 149–157 CAS.
  598. S. Akanuma, T. Kigawa and S. Yokoyama, Combinatorial mutagenesis to restrict amino acid usage in an enzyme to a reduced set, Proc. Natl. Acad. Sci. U. S. A., 2002, 99, 13549–13553 CrossRef CAS PubMed.
  599. L. H. Bradley, R. E. Kleiner, A. F. Wang, M. H. Hecht and D. W. Wood, An intein-based genetic selection allows the construction of a high-quality library of binary patterned de novo protein sequences, Protein Eng., Des. Sel., 2005, 18, 201–207 CrossRef CAS PubMed.
  600. J. F. Chaparro-Riggers, K. M. Polizzi and A. S. Bommarius, Better library design: data-driven protein engineering, Biotechnol. J., 2007, 2, 180–191 CrossRef PubMed.
  601. J. Tanaka, H. Yanagawa and N. Doi, Comparison of the frequency of functional SH3 domains with different limited sets of amino acids using mRNA display, PLoS One, 2011, 6, e18034 CAS.
  602. A. G. Sandström, Y. Wikmark, K. Engström, J. Nyhlén and J. E. Bäckvall, Combinatorial reshaping of the Candida antarctica lipase A substrate pocket for enantioselectivity using an extremely condensed library, Proc. Natl. Acad. Sci. U. S. A., 2012, 109, 78–83 CrossRef PubMed.
  603. J. Bacardit, M. Stout, J. D. Hirst, A. Valencia, R. E. Smith and N. Krasnogor, Automated alphabet reduction for protein datasets, BMC Bioinf., 2009, 10, 6 CrossRef PubMed.
  604. S. Zheng and I. Kwon, Manipulation of enzyme properties by noncanonical amino acid incorporation, Biotechnol. J., 2012, 7, 47–60 CrossRef CAS PubMed.
  605. M. G. Hoesl and N. Budisa, In vivo incorporation of multiple noncanonical amino acids into proteins, Angew. Chem., Int. Ed., 2011, 50, 2896–2902 CrossRef CAS PubMed.
  606. L. Wang, A. Brock, B. Herberich and P. G. Schultz, Expanding the genetic code of Escherichia coli, Science, 2001, 292, 498–500 CrossRef CAS PubMed.
  607. Q. Wang, A. R. Parrish and L. Wang, Expanding the genetic code for biological studies, Chem. Biol., 2009, 16, 323–336 CrossRef CAS PubMed.
  608. J. C. Jackson, S. P. Duffy, K. R. Hess and R. A. Mehl, Improving Nature's enzyme active site with genetically encoded unnatural amino acids, J. Am. Chem. Soc., 2006, 128, 11124–11127 CrossRef CAS PubMed.
  609. J. W. Chin, T. A. Cropp, J. C. Anderson, M. Mukherji, Z. Zhang and P. G. Schultz, An expanded eukaryotic genetic code, Science, 2003, 301, 964–967 CrossRef CAS PubMed.
  610. J. C. Anderson, N. Wu, S. W. Santoro, V. Lakshman, D. S. King and P. G. Schultz, An expanded genetic code with a functional quadruplet codon, Proc. Natl. Acad. Sci. U. S. A., 2004, 101, 7566–7571 CrossRef CAS PubMed.
  611. H. Neumann, K. Wang, L. Davis, M. Garcia-Alai and J. W. Chin, Encoding multiple unnatural amino acids via evolution of a quadruplet-decoding ribosome, Nature, 2010, 464, 441–444 CrossRef CAS PubMed.
  612. J. W. Chin, Reprogramming the genetic code, Science, 2012, 336, 428–429 CrossRef CAS PubMed.
  613. L. Davis and J. W. Chin, Designer proteins: applications of genetic code expansion in cell biology, Nat. Rev. Mol. Cell Biol., 2012, 13, 168–182 CAS.
  614. K. Wang, W. H. Schmied and J. W. Chin, Reprogramming the genetic code: from triplet to quadruplet codes, Angew. Chem., Int. Ed., 2012, 51, 2288–2297 CrossRef CAS PubMed.
  615. C. C. Liu and P. G. Schultz, Adding new chemistries to the genetic code, Annu. Rev. Biochem., 2010, 79, 413–444 CrossRef CAS PubMed.
  616. C. C. Liu, A. V. Mack, M. L. Tsao, J. H. Mills, H. S. Lee, H. Choe, M. Farzan, P. G. Schultz and V. V. Smider, Protein evolution with an expanded genetic code, Proc. Natl. Acad. Sci. U. S. A., 2008, 105, 17688–17693 CrossRef CAS PubMed.
  617. J. Xie and P. G. Schultz, A chemical toolkit for proteins—an expanded genetic code, Nat. Rev. Mol. Cell Biol., 2006, 7, 775–782 CrossRef CAS PubMed.
  618. L. Wang, J. Xie and P. G. Schultz, Expanding the genetic code, Annu. Rev. Biophys. Biomol. Struct., 2006, 35, 225–249 CrossRef CAS PubMed.
  619. K. M. Bradley and S. A. Benner, OligArch: A software tool to allow artificially expanded genetic information systems (AEGIS) to guide the autonomous self-assembly of long DNA constructs from multiple DNA single strands, Beilstein J. Org. Chem., 2014, 10, 1826–1833 CrossRef PubMed.
  620. J. Xie and P. G. Schultz, Adding amino acids to the genetic repertoire, Curr. Opin. Chem. Biol., 2005, 9, 548–554 CrossRef CAS PubMed.
  621. W. P. C. Stemmer, Rapid evolution of a protein in vivo by DNA shuffling, Nature, 1994, 370, 389–391 CrossRef CAS PubMed.
  622. W. P. C. Stemmer, DNA shuffling by random fragmentation and reassembly: in vitro recombination for molecular evolution., Proc. Natl. Acad. Sci. U. S. A., 1994, 91, 10747–10751 CrossRef CAS.
  623. H. Zhao and F. H. Arnold, Optimization of DNA shuffling for high fidelity recombination, Nucleic Acids Res., 1997, 25, 1307–1308 CrossRef CAS PubMed.
  624. W. M. Coco, W. E. Levinson, M. J. Crist, H. J. Hektor, A. Darzins, P. T. Pienkos, C. H. Squires and D. J. Monticello, DNA shuffling method for generating highly recombined genes and evolved enzymes, Nat. Biotechnol., 2001, 19, 354–359 CrossRef CAS PubMed.
  625. W. M. Coco, L. P. Encell, W. E. Levinson, M. J. Crist, A. K. Loomis, L. L. Licato, J. J. Arensdorf, N. Sica, P. T. Pienkos and D. J. Monticello, Growth factor engineering by degenerate homoduplex gene family recombination, Nat. Biotechnol., 2002, 20, 1246–1250 CrossRef CAS PubMed.
  626. M. Kikuchi, K. Ohnishi and S. Harayama, Novel family shuffling methods for the in vitro evolution of enzymes, Gene, 1999, 236, 159–167 CrossRef CAS.
  627. M. Kikuchi, K. Ohnishi and S. Harayama, An effective family shuffling method using single-stranded DNA, Gene, 2000, 243, 133–137 CrossRef CAS.
  628. K. Miyazaki, Random DNA fragmentation with endonuclease V: application to DNA shuffling, Nucleic Acids Res., 2002, 30, e139 CrossRef PubMed.
  629. Y. An, W. Wu and A. Lv, A convenient and robust method for construction of combinatorial and random mutant libraries, Biochimie, 2010, 92, 1081–1084 CrossRef CAS PubMed.
  630. A. Crameri, S. A. Raillard, E. Bermudez and W. P. C. Stemmer, DNA shuffling of a family of genes from diverse species accelerates directed evolution, Nature, 1998, 391, 288–291 CrossRef CAS PubMed.
  631. J. B. Y. H. Behrendorff, W. A. Johnston and E. M. J. Gillam, Restriction enzyme-mediated DNA family shuffling, Methods Mol. Biol., 2014, 1179, 175–187 Search PubMed.
  632. J. B. Y. H. Behrendorff, W. A. Johnston and E. M. J. Gillam, DNA shuffling of cytochrome P450 enzymes, Methods Mol. Biol., 2013, 987, 177–188 CAS.
  633. N. N. Rosic, W. Huang, W. A. Johnston, J. J. DeVoss and E. M. J. Gillam, Extending the diversity of cytochrome P450 enzymes by DNA family shuffling, Gene, 2007, 395, 40–48 CrossRef CAS PubMed.
  634. J. M. Joern, P. Meinhold and F. H. Arnold, Analysis of shuffled gene libraries, J. Mol. Biol., 2002, 316, 643–656 CrossRef CAS PubMed.
  635. J. F. Chaparro-Riggers, B. L. Loo, K. M. Polizzi, P. R. Gibbs, X. S. Tang, M. J. Nelson and A. S. Bommarius, Revealing biases inherent in recombination protocols, BMC Biotechnol., 2007, 7, 77 CrossRef PubMed.
  636. Y. Kawarasaki, K. E. Griswold, J. D. Stevenson, T. Selzer, S. J. Benkovic, B. L. Iverson and G. Georgiou, Enhanced crossover SCRATCHY: construction and high-throughput screening of a combinatorial library containing multiple non-homologous crossovers, Nucleic Acids Res., 2003, 31, e126 CrossRef PubMed.
  637. S. Lutz and M. Ostermeier, Preparation of SCRATCHY hybrid protein libraries: size- and in-frame selection of nucleic acid sequences, Methods Mol. Biol., 2003, 231, 143–151 CAS.
  638. M. Ostermeier and S. Lutz, The creation of ITCHY hybrid protein libraries, Methods Mol. Biol., 2003, 231, 129–141 CAS.
  639. W. M. Patrick and M. L. Gerth, ITCHY: Incremental Truncation for the Creation of Hybrid enzYmes, Methods Mol. Biol., 2014, 1179, 225–244 Search PubMed.
  640. W. M. Coco, RACHITT: Gene family shuffling by Random Chimeragenesis on Transient Templates, Methods Mol. Biol., 2003, 231, 111–127 CAS.
  641. S. H. Lee, E. J. Ryu, M. J. Kang, E. S. Wang, Z. Piao, Y. J. Choi, K. H. Jung, J. Y. J. Jeon and Y. C. Shin, A new approach to directed gene evolution by recombined extension on truncated templates (RETT), J. Mol. Catal., 2003, 26, 119–129 CrossRef CAS PubMed.
  642. A. Hidalgo, A. Schliessmann, R. Molina, J. Hermoso and U. T. Bornscheuer, A one-pot, simple methodology for cassette randomisation and recombination for focused directed evolution, Protein Eng., Des. Sel., 2008, 21, 567–576 CrossRef CAS PubMed.
  643. A. Hidalgo, A. Schliessmann and U. T. Bornscheuer, One-pot Simple methodology for CAssette Randomization and Recombination for focused directed evolution (OSCARR), Methods Mol. Biol., 2014, 1179, 207–212 Search PubMed.
  644. K. Kashiwagi, Y. Isogai, K. Nishiguchi and K. Shiba, Frame shuffling: a novel method for in vitro protein evolution, Protein Eng., Des. Sel., 2006, 19, 135–140 CrossRef CAS PubMed.
  645. J. E. Ness, S. Kim, A. Gottman, R. Pak, A. Krebber, T. V. Borchert, S. Govindarajan, E. C. Mundorff and J. Minshull, Synthetic shuffling expands functional protein diversity by allowing amino acids to recombine independently, Nat. Biotechnol., 2002, 20, 1251–1255 CrossRef CAS PubMed.
  646. P. L. Bergquist, R. A. Reeves and M. D. Gibbs, Degenerate oligonucleotide gene shuffling (DOGS) and random drift mutagenesis (RNDM): two complementary techniques for enzyme evolution, Biomol. Eng., 2005, 22, 63–72 CrossRef CAS PubMed.
  647. B. R. Villiers, V. Stein and F. Hollfelder, USER friendly DNA recombination (USERec): a simple and flexible near homology-independent method for gene library construction, Protein Eng., Des. Sel., 2010, 23, 1–8 CrossRef CAS PubMed.
  648. B. Villiers and F. Hollfelder, USER friendly DNA recombination (USERec): gene library construction requiring minimal sequence homology, Methods Mol. Biol., 2014, 1179, 213–224 Search PubMed.
  649. P. E. O'Maille, M. Bakhtina and M. D. Tsai, Structure-based combinatorial protein engineering (SCOPE), J. Mol. Biol., 2002, 321, 677–691 CrossRef.
  650. P. E. O'Maille, M. D. Tsai, B. T. Greenhagen, J. Chappell and J. P. Noel, Gene library synthesis by structure-based combinatorial protein engineering, Methods Enzymol., 2004, 388, 75–91 Search PubMed.
  651. M. Dokarry, C. Laurendon and P. E. O'Maille, Automating gene library synthesis by structure-based combinatorial protein engineering: examples from plant sesquiterpene synthases, Methods Enzymol., 2012, 515, 21–42 CAS.
  652. A. Herman and D. S. Tawfik, Incorporating Synthetic Oligonucleotides via Gene Reassembly (ISOR): a versatile tool for generating targeted libraries, Protein Eng., Des. Sel., 2007, 20, 219–226 CrossRef CAS PubMed.
  653. L. Rockah-Shmuel, D. S. Tawfik and M. Goldsmith, Generating targeted libraries by the combinatorial incorporation of synthetic oligonucleotides during gene shuffling (ISOR), Methods Mol. Biol., 2014, 1179, 129–137 Search PubMed.
  654. M. D. Hughes, Z. R. Zhang, A. J. Sutherland, A. F. Santos and A. V. Hine, Discovery of active proteins directly from combinatorial randomized protein libraries without display, purification or sequencing: identification of novel zinc finger proteins, Nucleic Acids Res., 2005, 33, e32 CrossRef PubMed.
  655. Z. Shao, H. Zhao and H. Zhao, DNA assembler, an in vivo genetic method for rapid construction of biochemical pathways, Nucleic Acids Res., 2009, 37, e16 CrossRef PubMed.
  656. Z. Shao, Y. Luo and H. Zhao, Rapid characterization and engineering of natural product biosynthetic pathways via DNA assembler, Mol. BioSyst., 2011, 7, 1056–1059 RSC.
  657. M. Z. Li and S. J. Elledge, Harnessing homologous recombination in vitro to generate recombinant DNA via SLIC, Nat. Methods, 2007, 4, 251–256 CrossRef CAS PubMed.
  658. J. A. Mosberg, C. J. Gregg, M. J. Lajoie, H. H. Wang and G. M. Church, Improving lambda red genome engineering in Escherichia colivia rational removal of endogenous nucleases, PLoS One, 2012, 7, e44638 CAS.
  659. E. M. Nordwald, A. Garst, R. T. Gill and J. L. Kaar, Accelerated protein engineering for chemical biotechnology via homologous recombination, Curr. Opin. Biotechnol., 2013, 24, 1017–1022 CrossRef CAS PubMed.
  660. Z. Qian and S. Lutz, Improving the catalytic activity of Candida antarctica lipase B by circular permutation, J. Am. Chem. Soc., 2005, 127, 13466–13467 CrossRef CAS PubMed.
  661. Z. Qian, C. J. Fields and S. Lutz, Investigating the structural and functional consequences of circular permutation on lipase B from Candida antarctica, ChemBioChem, 2007, 8, 1989–1996 CrossRef CAS PubMed.
  662. S. Lutz, A. B. Daugherty, Y. Yu and Z. Qian, Generating random circular permutation libraries, Methods Mol. Biol., 2014, 1179, 245–258 Search PubMed.
  663. E. Fischereder, D. Pressnitz, W. Kroutil and S. Lutz, Engineering strictosidine synthase: Rational design of a small, focused circular permutation library of the beta-propeller fold enzyme, Bioorg. Med. Chem., 2014, 22, 5633–5637 CrossRef CAS PubMed.
  664. A. B. Daugherty, S. Govindarajan and S. Lutz, Improved biocatalysts from a synthetic circular permutation library of the flavin-dependent oxidoreductase old yellow enzyme, J. Am. Chem. Soc., 2013, 135, 14425–14432 CrossRef CAS PubMed.
  665. Y. Yu and S. Lutz, Circular permutation: a different way to engineer enzyme structure and function, Trends Biotechnol., 2011, 29, 18–25 CrossRef CAS PubMed.
  666. E. Haglund, M. O. Lindberg and M. Oliveberg, Changes of protein folding pathways by circular permutation. Overlapping nuclei promote global cooperativity, J. Biol. Chem., 2008, 283, 27904–27915 CrossRef CAS PubMed.
  667. M. Lindberg, J. Tangrot and M. Oliveberg, Complete change of the protein folding transition state upon circular permutation, Nat. Struct. Biol., 2002, 9, 818–822 CAS.
  668. M. A. Smith, P. A. Romero, T. Wu, E. M. Brustad and F. H. Arnold, Chimeragenesis of distantly-related proteins by noncontiguous recombination, Protein Sci., 2013, 22, 231–238 CrossRef CAS PubMed.
  669. D. L. Trudeau, M. A. Smith and F. H. Arnold, Innovation by homologous recombination, Curr. Opin. Chem. Biol., 2013, 17, 902–909 CrossRef CAS PubMed.
  670. M. C. Saraf, A. Gupta and C. D. Maranas, Design of combinatorial protein libraries of optimal size, Proteins, 2005, 60, 769–777 CrossRef CAS PubMed.
  671. R. J. Pantazes, M. C. Saraf and C. D. Maranas, Optimal protein library design using recombination or point mutations based on sequence-based scoring functions, Protein Eng., Des. Sel., 2007, 20, 361–373 CrossRef CAS PubMed.
  672. J. B. Endelman, J. J. Silberg, Z. G. Wang and F. H. Arnold, Site-directed protein recombination as a shortest-path problem, Protein Eng., Des. Sel., 2004, 17, 589–594 CrossRef CAS PubMed.
  673. C. A. Voigt, C. Martinez, Z. G. Wang, S. L. Mayo and F. H. Arnold, Protein building blocks preserved by recombination, Nat. Struct. Biol., 2002, 9, 553–558 CAS.
  674. C. R. Otey, J. J. Silberg, C. A. Voigt, J. B. Endelman, G. Bandara and F. H. Arnold, Functional evolution and structural conservation in chimeric cytochromes p450: calibrating a structure-guided approach, Chem. Biol., 2004, 11, 309–318 CrossRef CAS PubMed.
  675. P. Heinzelman, C. D. Snow, M. A. Smith, X. Yu, A. Kannan, K. Boulware, A. Villalobos, S. Govindarajan, J. Minshull and F. H. Arnold, SCHEMA recombination of a fungal cellulase uncovers a single mutation that contributes markedly to stability, J. Biol. Chem., 2009, 284, 26229–26233 CrossRef CAS PubMed.
  676. P. Heinzelman, R. Komor, A. Kanaan, P. Romero, X. Yu, S. Mohler, C. Snow and F. Arnold, Efficient screening of fungal cellobiohydrolase class I enzymes for thermostabilizing sequence blocks by SCHEMA structure-guided recombination, Protein Eng., Des. Sel., 2010, 23, 871–880 CrossRef CAS PubMed.
  677. P. Heinzelman, P. A. Romero and F. H. Arnold, Efficient sampling of SCHEMA chimera families to identify useful sequence elements, Methods Enzymol., 2013, 523, 351–368 CAS.
  678. M. M. Meyer, J. J. Silberg, C. A. Voigt, J. B. Endelman, S. L. Mayo, Z. G. Wang and F. H. Arnold, Library analysis of SCHEMA-guided protein recombination, Protein Sci., 2003, 12, 1686–1693 CrossRef CAS PubMed.
  679. M. M. Meyer, L. Hochrein and F. H. Arnold, Structure-guided SCHEMA recombination of distantly related beta-lactamases, Protein Eng., Des. Sel., 2006, 19, 563–570 CrossRef CAS PubMed.
  680. P. A. Romero, E. Stone, C. Lamb, L. Chantranupong, A. Krause, A. E. Miklos, R. A. Hughes, B. Fechtel, A. D. Ellington, F. H. Arnold and G. Georgiou, SCHEMA-designed variants of human Arginase I and II reveal sequence elements important to stability and catalysis, ACS Synth. Biol., 2012, 1, 221–228 CrossRef CAS PubMed.
  681. J. J. Silberg, J. B. Endelman and F. H. Arnold, SCHEMA-guided protein recombination, Methods Enzymol., 2004, 388, 35–42 CAS.
  682. M. A. Smith and F. H. Arnold, Designing libraries of chimeric proteins using SCHEMA recombination and RASPP, Methods Mol. Biol., 2014, 1179, 335–343 Search PubMed.
  683. M. A. Smith and F. H. Arnold, Noncontiguous SCHEMA protein recombination, Methods Mol. Biol., 2014, 1179, 345–352 Search PubMed.
  684. A. S. Parker, K. E. Griswold and C. Bailey-Kellogg, Optimization of Combinatorial Mutagenesis, Research in Computational Molecular Biology, 2011, 6577, 321–335 Search PubMed , 580.
  685. H. Zhao, K. Blazanovic, Y. Choi, C. Bailey-Kellogg and K. E. Griswold, Gene and protein sequence optimization for high-level production of fully active and aglycosylated lysostaphin in Pichia pastoris, Appl. Environ. Microbiol., 2014, 80, 2746–2753 CrossRef CAS PubMed.
  686. L. He, A. M. Friedman and C. Bailey-Kellogg, Algorithms for optimizing cross-overs in DNA shuffling, BMC Bioinf., 2012, 13(suppl 3), S3 CrossRef PubMed.
  687. W. Zheng, A. M. Friedman and C. Bailey-Kellogg, Algorithms for joint optimization of stability and diversity in planning combinatorial libraries of chimeric proteins, J. Comput. Biol., 2009, 16, 1151–1168 CrossRef CAS PubMed.
  688. X. Ye, A. M. Friedman and C. Bailey-Kellogg, Optimizing Bayes error for protein structure model selection by stability mutagenesis, Comput. Syst. Bioinf., CSB2007 Conf. Proc., 6th, 2008, 7, 99–108 Search PubMed.
  689. W. Zheng, X. Ye, A. M. Friedman and C. Bailey-Kellogg, Algorithms for selecting breakpoint locations to optimize diversity in protein engineering by site-directed protein recombination, Comput. Syst. Bioinf., CSB2007 Conf. Proc., 6th, 2007, 6, 31–40 Search PubMed.
  690. X. Ye, A. M. Friedman and C. Bailey-Kellogg, Hypergraph model of multi-residue interactions in proteins: sequentially-constrained partitioning algorithms for optimization of site-directed protein recombination, J. Comput. Biol., 2007, 14, 777–790 CrossRef CAS PubMed.
  691. L. Saftalov, P. A. Smith, A. M. Friedman and C. Bailey-Kellogg, Site-directed combinatorial construction of chimaeric genes: general method for optimizing assembly of gene fragments, Proteins, 2006, 64, 629–642 CrossRef CAS PubMed.
  692. Y. Li, D. A. Drummond, A. M. Sawayama, C. D. Snow, J. D. Bloom and F. H. Arnold, A diverse family of thermostable cytochrome P450s created by recombination of stabilizing fragments, Nat. Biotechnol., 2007, 25, 1051–1056 CrossRef CAS PubMed.
  693. D. Lipovsek and A. Plückthun, In vitro protein evolution by ribosome display and mRNA display, J. Immunol. Methods, 2004, 290, 51–67 CrossRef CAS PubMed.
  694. M. Y. He, Cell-free protein synthesis: applications in proteomics and biotechnology, New Biotechnol., 2008, 25, 126–132 CrossRef CAS PubMed.
  695. T. Okano, T. Matsuura, H. Suzuki and T. Yomo, Cell-free Protein Synthesis in a Microchamber Revealed the Presence of an Optimum Compartment Volume for High-order Reactions, ACS Synth. Biol., 2014, 3, 347–352 CrossRef CAS PubMed.
  696. T. Nishikawa, T. Sunami, T. Matsuura and T. Yomo, Directed Evolution of Proteins through In Vitro Protein Synthesis in Liposomes, J. Nucleic Acids, 2012, 2012, 923214 Search PubMed.
  697. T. Okano, T. Matsuura, Y. Kazuta, H. Suzuki and T. Yomo, Cell-free protein synthesis from a single copy of DNA in a glass microchamber, Lab Chip, 2012, 12, 2704–2711 RSC.
  698. K. Nishimura, T. Matsuura, T. Sunami, H. Suzuki and T. Yomo, Cell-free protein synthesis inside giant unilamellar vesicles analyzed by flow cytometry, Langmuir, 2012, 28, 8426–8432 CrossRef CAS PubMed.
  699. H. Yanagida, T. Matsuura and T. Yomo, Ribosome display for rapid protein evolution by consecutive rounds of mutation and selection, Methods Mol. Biol., 2010, 634, 257–267 CAS.
  700. M. T. Smith, K. M. Wilding, J. M. Hunt, A. M. Bennett and B. C. Bundy, The emerging age of cell-free synthetic biology, FEBS Lett., 2014, 588, 2755–2761 CrossRef CAS PubMed.
  701. M. H. Caruthers, A. D. Barone, S. L. Beaucage, D. R. Dodds, E. F. Fisher, L. J. McBride, M. Matteucci, Z. Stabinsky and J. Y. Tang, Chemical synthesis of deoxyoligonucleotides by the phosphoramidite method, Methods Enzymol., 1987, 154, 287–313 CAS.
  702. J. D. Tian, K. S. Ma and I. Saaem, Advancing high-throughput gene synthesis technology, Mol. BioSyst., 2009, 5, 714–722 RSC.
  703. E. M. LeProust, B. J. Peck, K. Spirin, H. B. McCuen, B. Moore, E. Namsaraev and M. H. Caruthers, Synthesis of high-quality libraries of long (150mer) oligonucleotides by a novel depurination controlled process, Nucleic Acids Res., 2010, 38, 2522–2540 CrossRef CAS PubMed.
  704. K. E. Richmond, M. H. Li, M. J. Rodesch, M. Patel, A. M. Lowe, C. Kim, L. L. Chu, N. Venkataramaian, S. F. Flickinger, J. Kaysen, P. J. Belshaw, M. R. Sussman and F. Cerrina, Amplification and assembly of chip-eluted DNA (AACED): a method for high-throughput gene synthesis, Nucleic Acids Res., 2004, 32, 5011–5018 CrossRef CAS PubMed.
  705. A. Y. Borovkov, A. V. Loskutov, M. D. Robida, K. M. Day, J. A. Cano, T. Le Olson, H. Patel, K. Brown, P. D. Hunter and K. F. Sykes, High-quality gene assembly directly from unpurified mixtures of microarray-synthesized oligonucleotides, Nucleic Acids Res., 2010, 38, e180 CrossRef PubMed.
  706. M. Matzas, P. F. Stähler, N. Kefer, N. Siebelt, V. Boisguérin, J. T. Leonard, A. Keller, C. F. Stähler, P. Häberle, B. Gharizadeh, F. Babrzadeh and G. M. Church, High-fidelity gene synthesis by retrieval of sequence-verified DNA identified using high-throughput pyrosequencing, Nat. Biotechnol., 2010, 28, 1291–1294 CrossRef CAS PubMed.
  707. H. Kim, H. Han, J. Ahn, J. Lee, N. Cho, H. Jang, H. Kim, S. Kwon and D. Bang, 'Shotgun DNA synthesis' for the high-throughput construction of large DNA molecules, Nucleic Acids Res., 2012, 40, e140 CrossRef CAS PubMed.
  708. G. M. Church, M. B. Elowitz, C. D. Smolke, C. A. Voigt and R. Weiss, Realizing the potential of synthetic biology, Nat. Rev. Mol. Cell Biol., 2014, 15, 289–294 CrossRef CAS PubMed.
  709. I. Tabuchi, S. Soramoto, S. Ueno and Y. Husimi, Multi-line split DNA synthesis: a novel combinatorial method to make high quality peptide libraries, BMC Biotechnol., 2004, 4, 19 CrossRef PubMed.
  710. J. Liang, Y. Luo and H. Zhao, Synthetic biology: putting synthesis into biology, Wiley Interdiscip. Rev.: Syst. Biol. Med., 2011, 3, 7–20 CrossRef CAS PubMed.
  711. S. A. Lynch and R. T. Gill, Synthetic biology: New strategies for directing design, Metab. Eng., 2011, 14, 205–211 CrossRef PubMed.
  712. J. L. Foo, C. B. Ching, M. W. Chang and S. S. Leong, The imminent role of protein engineering in synthetic biology, Biotechnol. Adv., 2012, 30, 541–549 CrossRef CAS PubMed.
  713. D. W. Watkins, C. T. Armstrong and J. L. Anderson, De novo protein components for oxidoreductase assembly and biological integration, Curr. Opin. Chem. Biol., 2014, 19, 90–98 CrossRef CAS PubMed.
  714. S. Ma, N. Tang and J. Tian, DNA synthesis, assembly and applications in synthetic biology, Curr. Opin. Chem. Biol., 2012, 16, 260–267 CrossRef CAS PubMed.
  715. W. P. C. Stemmer, A. Crameri, K. D. Ha, T. M. Brennan and H. L. Heyneker, Single-step assembly of a gene and entire plasmid from large numbers of oligodeoxyribonucleotides, Gene, 1995, 164, 49–53 CrossRef CAS.
  716. A. S. Xiong, Q. H. Yao, R. H. Peng, X. Li, H. Q. Fan, Z. M. Cheng and Y. Li, A simple, rapid, high-fidelity and cost-effective PCR-based two-step DNA synthesis method for long gene sequences, Nucleic Acids Res., 2004, 32, e98 CrossRef PubMed.
  717. H. O. Smith, C. A. Hutchison, 3rd, C. Pfannkoch and J. C. Venter, Generating a synthetic genome by whole genome assembly: phiX, 174, bacteriophage from synthetic oligonucleotides, Proc. Natl. Acad. Sci. U. S. A., 2003, 100, 15440–15445 CrossRef CAS PubMed.
  718. G. H. Yang, S. Q. Wang, H. L. Wei, J. Ping, J. Liu, L. M. Xu and W. W. Zhang, Patch oligodeoxynucleotide synthesis (POS): a novel method for synthesis of long DNA sequences and full-length genes, Biotechnol. Lett., 2012, 34, 721–728 CrossRef CAS PubMed.
  719. S. de Kok, L. H. Stanton, T. Slaby, M. Durot, V. F. Holmes, K. G. Patel, D. Platt, E. B. Shapland, Z. Serber, J. Dean, J. D. Newman and S. S. Chandran, Rapid and reliable DNA assembly via ligase cycling reaction, ACS Synth. Biol., 2014, 3, 97–106 CrossRef CAS PubMed.
  720. T. L. Roth, L. Milenkovic and M. P. Scott, A rapid and simple method for DNA engineering using cycled ligation assembly, PLoS One, 2014, 9, e107329 Search PubMed.
  721. J. Cherry, B. W. Nieuwenhuijsen, E. J. Kaftan, J. D. Kennedy and P. K. Chanda, A modified method for PCR-directed gene synthesis from large number of overlapping oligodeoxyribonucleotides, J. Biochem. Biophys. Methods, 2008, 70, 820–822 CrossRef CAS PubMed.
  722. A. S. Xiong, R. H. Peng, J. Zhuang, F. Gao, Y. Li, Z. M. Cheng and Q. H. Yao, Chemical gene synthesis: strategies, softwares, error corrections, and applications, FEMS Microbiol. Rev., 2008, 32, 522–540 CrossRef CAS PubMed.
  723. B. F. Binkowski, K. E. Richmond, J. Kaysen, M. R. Sussman and P. J. Belshaw, Correcting errors in synthetic DNA through consensus shuffling, Nucleic Acids Res., 2005, 33, e55 CrossRef PubMed.
  724. A. S. Xiong, Q. H. Yao, R. H. Peng, H. Duan, X. Li, H. Q. Fan, Z. M. Cheng and Y. Li, PCR-based accurate synthesis of long DNA sequences, Nat. Protoc., 2006, 1, 791–797 CrossRef CAS PubMed.
  725. P. A. Carr, J. S. Park, Y. J. Lee, T. Yu, S. Zhang and J. M. Jacobson, Protein-mediated error correction for de novo DNA synthesis, Nucleic Acids Res., 2004, 32, e162 CrossRef PubMed.
  726. M. Fuhrmann, W. Oertel, P. Berthold and P. Hegemann, Removal of mismatched bases from synthetic genes by enzymatic mismatch cleavage, Nucleic Acids Res., 2005, 33, e58 CrossRef PubMed.
  727. I. Saaem, S. Ma, J. Quan and J. Tian, Error correction of microchip synthesized genes using Surveyor nuclease, Nucleic Acids Res., 2012, 40, e23 CrossRef CAS PubMed.
  728. A. Currin, N. Swainston, P. J. Day and D. B. Kell, SpeedyGenes: a novel approach for the efficient production of error-corrected, synthetic gene libraries, Protein Evol Design Sel, 2014, 27, 273–280 CrossRef CAS PubMed.
  729. N. Swainston, A. Currin, P. J. Day and D. B. Kell, GeneGenie: optimised oligomer design for directed evolution, Nucleic Acids Res., 2014, 12, W395–W400 CrossRef PubMed.
  730. H. Lin and V. W. Cornish, Screening and selection methods for large-scale analysis of protein function, Angew. Chem., Int. Ed., 2002, 41, 4402–4425 CrossRef CAS.
  731. Y. L. Boersma, M. J. Dröge and W. J. Quax, Selection strategies for improved biocatalysts, FEBS J., 2007, 274, 2181–2195 CrossRef CAS PubMed.
  732. C. Troll, D. Alexander, J. Allen, J. Marquette and M. Camps, Mutagenesis and functional selection protocols for directed evolution of proteins in E. coli, J. Visualized Exp., 2011, 49 Search PubMed.
  733. M. R. Parikh, D. N. Greene, K. K. Woods and I. Matsumura, Directed evolution of RuBisCO hypermorphs through genetic selection in engineered E.coli, Protein Eng., Des. Sel., 2006, 19, 113–119 CrossRef CAS PubMed.
  734. M. T. Reetz, H. Hobenreich, P. Soni and L. Fernandez, A genetic selection system for evolving enantioselectivity of enzymes, Chem. Commun., 2008, 5502–5504 RSC.
  735. C. G. Acevedo-Rocha, R. Agudo and M. T. Reetz, Directed evolution of stereoselective enzymes based on genetic selection as opposed to screening systems, J. Biotechnol., 2014, 191C, 3–10,  DOI:10.1016/j.jbiotec.2014.1004.1009.
  736. Y. L. Boersma, M. J. Dröge, A. M. van der Sloot, T. Pijning, R. H. Cool, B. W. Dijkstra and W. J. Quax, A novel genetic selection system for improved enantioselectivity of Bacillus subtilis lipase A, ChemBioChem, 2008, 9, 1110–1115 CrossRef CAS PubMed.
  737. P. Peralta-Yahya, B. T. Carter, H. Lin, H. Tao and V. W. Cornish, High-throughput selection for cellulase catalysts using chemical complementation, J. Am. Chem. Soc., 2008, 130, 17446–17452 CrossRef CAS PubMed.
  738. H. Tao, P. Peralta-Yahya, H. Lin and V. W. Cornish, Optimized design and synthesis of chemical dimerizer substrates for detection of glycosynthase activity via chemical complementation, Bioorg. Med. Chem., 2006, 14, 6940–6953 CrossRef CAS PubMed.
  739. S. K. Desai and J. P. Gallivan, Genetic screens and selections for small molecules based on a synthetic riboswitch that activates protein translation, J. Am. Chem. Soc., 2004, 126, 13247–13254 CrossRef CAS PubMed.
  740. W. C. Winkler and R. R. Breaker, Regulation of bacterial gene expression by riboswitches, Annu. Rev. Microbiol., 2005, 59, 487–517 CrossRef CAS PubMed.
  741. Y. Nomura and Y. Yokobayashi, Reengineering a natural riboswitch by dual genetic selection, J. Am. Chem. Soc., 2007, 129, 13814–13815 CrossRef CAS PubMed.
  742. N. Dixon, J. N. Duncan, T. Geerlings, M. S. Dunstan, J. E. McCarthy, D. Leys and J. Micklefield, Reengineering orthogonally selective riboswitches, Proc. Natl. Acad. Sci. U. S. A., 2010, 107, 2830–2835 CrossRef CAS PubMed.
  743. N. Dixon, C. J. Robinson, T. Geerlings, J. N. Duncan, S. P. Drummond and J. Micklefield, Orthogonal riboswitches for tuneable coexpression in bacteria, Angew. Chem., Int. Ed., 2012, 51, 3620–3624 CrossRef CAS PubMed.
  744. L. M. Wingler and V. W. Cornish, A library approach for the discovery of customized yeast three-hybrid counter selections, ChemBioChem, 2011, 12, 715–717 CrossRef CAS PubMed.
  745. Y. Yokobayashi and F. H. Arnold, A dual selection module for directed evolution of genetic circuits, Nat. Comput., 2005, 4, 245–254 CrossRef CAS.
  746. W. Besenmatter, P. Kast and D. Hilvert, New enzymes from combinatorial library modules, Methods Enzymol., 2004, 388, 91–102 CAS.
  747. E. G. Hibbert and P. A. Dalby, Directed evolution strategies for improved enzymatic performance, Microb. Cell Fact., 2005, 4, 29 CrossRef PubMed.
  748. M. M. Müller, H. Kries, E. Csuhai, P. Kast and D. Hilvert, Design, selection, and characterization of a split chorismate mutase, Protein Sci., 2010, 19, 1000–1010 Search PubMed.
  749. K. Lanthaler, E. Bilsland, P. Dobson, H. J. Moss, P. Pir, D. B. Kell and S. G. Oliver, Genome-wide assessment of the carriers involved in the cellular uptake of drugs: a model system in yeast, BMC Biol., 2011, 9, 70 CrossRef CAS PubMed.
  750. G. E. Winter, B. Radic, C. Mayor-Ruiz, V. A. Blomen, C. Trefzer, R. K. Kandasamy, K. V. M. Huber, M. Gridling, D. Chen, T. Klampfl, R. Kralovics, S. Kubicek, O. Fernandez-Capetillo, T. R. Brummelkamp and G. Superti-Furga, The solute carrier SLC35F2 enables YM155-mediated DNA damage toxicity, Nat. Chem. Biol., 2014, 10, 768–773 CrossRef CAS PubMed.
  751. M. L. Geddie, L. A. Rowe, O. B. Alexander and I. Matsumura, High throughput microplate screens for directed protein evolution, Methods Enzymol., 2004, 388, 134–145 CAS.
  752. G. An, J. Bielich, R. Auerbach and E. A. Johnson, Isolation and characterization of carotenoid hyperproducing mutants of yeast by flow cytometry and cell sorting, Bio/Technology, 1991, 9, 69–73 CrossRef PubMed.
  753. T. Azuma, G. I. Harrison and A. L. Demain, Isolation of gramicidin S hyperproducing strain of Bacillus brevis by use of a fluorescence activated cell sorting system., Appl. Microbiol. Biotechnol., 1992, 38, 173–178 CrossRef CAS.
  754. B. P. Cormack, R. H. Valdivia and S. Falkow, FACS-optimized mutants of the green fluorescent protein (GFP), Gene, 1996, 173, 33–38 CrossRef CAS.
  755. M. Reckermann, Flow sorting in aquatic ecology, Sci. Mar., 2000, 64, 235–246 Search PubMed.
  756. C. J. Hewitt and G. Nebe-Von-Caron, An industrial application of multiparameter flow cytometry: assessment of cell physiological state and its application to the study of microbial fermentations, Cytometry, 2001, 44, 179–187 CrossRef CAS.
  757. M. Rieseberg, C. Kasper, K. F. Reardon and T. Scheper, Flow cytometry in biotechnology, Appl. Microbiol. Biotechnol., 2001, 56, 350–360 CrossRef CAS.
  758. J. Vidal-Mas, P. Resina, E. Haba, J. Comas, A. Manresa and J. Vives-Rego, Rapid flow cytometry–Nile red assessment of PHA cellular content and heterogeneity in cultures of Pseudomonas aeruginosa 47T2 (NCIB 40044) grown in waste frying oil, Antonie van Leeuwenhoek, 2001, 80, 57–63 CrossRef CAS.
  759. S. W. Santoro and P. G. Schultz, Directed evolution of the site specificity of Cre recombinase, Proc. Natl. Acad. Sci. U. S. A., 2002, 99, 4185–4190 CrossRef CAS PubMed.
  760. K. Bernath, M. Hai, E. Mastrobattista, A. D. Griffiths, S. Magdassi and D. S. Tawfik, In vitro compartmentalization by double emulsions: sorting and gene enrichment by fluorescence activated cell sorting, Anal. Biochem., 2004, 325, 151–157 CrossRef CAS PubMed.
  761. A. R. Buskirk, Y. C. Ong, Z. J. Gartner and D. R. Liu, Directed evolution of ligand dependence: small-molecule-activated protein splicing, Proc. Natl. Acad. Sci. U. S. A., 2004, 101, 10505–10510 CrossRef CAS PubMed.
  762. C. J. Hewitt and G. Nebe-Von-Caron, The application of multi-parameter flow cytometry to monitor individual microbial cell physiological state, Adv. Biochem. Eng./Biotechnol., 2004, 89, 197–223 CrossRef CAS.
  763. A. Aharoni, K. Thieme, C. P. Chiu, S. Buchini, L. L. Lairson, H. Chen, N. C. Strynadka, W. W. Wakarchuk and S. G. Withers, High-throughput screening methodology for the directed evolution of glycosyltransferases, Nat. Methods, 2006, 3, 609–614 CrossRef CAS PubMed.
  764. H. M. Davey and D. B. Kell, Flow cytometry and cell sorting of heterogeneous microbial populations: the importance of single-cell analysis, Microbiol. Rev., 1996, 60, 641–696 CAS.
  765. D. Mattanovich and N. Borth, Applications of cell sorting in biotechnology, Microb. Cell Fact., 2006, 5, 12 CrossRef PubMed.
  766. O. J. Miller, K. Bernath, J. J. Agresti, G. Amitai, B. T. Kelly, E. Mastrobattista, V. Taly, S. Magdassi, D. S. Tawfik and A. D. Griffiths, Directed evolution by in vitro compartmentalization, Nat. Methods, 2006, 3, 561–570 CrossRef CAS PubMed.
  767. M. Valli, M. Sauer, P. Branduardi, N. Borth, D. Porro and D. Mattanovich, Improvement of lactic acid production in Saccharomyces cerevisiae by cell sorting for high intracellular pH, Appl. Environ. Microbiol., 2006, 72, 5492–5499 CrossRef CAS PubMed.
  768. S. Becker, H. Hobenreich, A. Vogel, J. Knorr, S. Wilhelm, F. Rosenau, K. E. Jaeger, M. T. Reetz and H. Kolmar, Single-cell high-throughput screening to identify enantioselective hydrolytic enzymes, Angew. Chem., Int. Ed., 2008, 47, 5085–5088 CrossRef CAS PubMed.
  769. N. Varadarajan, S. Rodriguez, B. Y. Hwang, G. Georgiou and B. L. Iverson, Highly active and selective endopeptidases with programmed substrate specificities, Nat. Chem. Biol., 2008, 4, 290–294 CrossRef CAS PubMed.
  770. L. Liu, Y. Li, D. Liotta and S. Lutz, Directed evolution of an orthogonal nucleoside analog kinase via fluorescence-activated cell sorting, Nucleic Acids Res., 2009, 37, 4472–4481 CrossRef CAS PubMed.
  771. G. Yang and S. G. Withers, Ultrahigh-throughput FACS-based screening for directed enzyme evolution, ChemBioChem, 2009, 10, 2704–2715 CrossRef CAS PubMed.
  772. M. Díaz, M. Herrero, L. A. García and C. Quirós, Application of flow cytometry to industrial microbial bioprocesses, Biochem. Eng. J., 2010, 48, 385–407 CrossRef PubMed.
  773. J. A. Dietrich, A. E. McKee and J. D. Keasling, High-throughput metabolic engineering: advances in small-molecule screening and selection, Annu. Rev. Biochem., 2010, 79, 563–590 CrossRef CAS PubMed.
  774. S. Iijima, Y. Shimomura, Y. Haba, F. Kawai, A. Tani and K. Kimbara, Flow cytometry-based method for isolating live bacteria with meta-cleavage activity on dihydroxy compounds of biphenyl, J. Biosci. Bioeng., 2010, 109, 645–651 CrossRef CAS PubMed.
  775. S. Ishii, K. Tago and K. Senoo, Single-cell analysis and isolation for microbiology and biotechnology: methods and applications, Appl. Microbiol. Biotechnol., 2010, 86, 1281–1292 CrossRef CAS PubMed.
  776. M. E. Lidstrom and M. C. Konopka, The role of physiological heterogeneity in microbial population behavior, Nat. Chem. Biol., 2010, 6, 705–712 CrossRef CAS PubMed.
  777. S. W. Lim and A. R. Abate, Ultrahigh-throughput sorting of microfluidic drops with flow cytometry, Lab Chip, 2013, 13, 4563–4572 RSC.
  778. S. Müller and G. Nebe-von-Caron, Functional single-cell analyses: flow cytometry and cell sorting of microbial populations and communities, FEMS Microbiol. Rev., 2010, 34, 554–587 Search PubMed.
  779. S. G. Rhee, T. S. Chang, W. Jeong and D. Kang, Methods for detection and measurement of hydrogen peroxide inside and outside of cells, Mol. Cells, 2010, 29, 539–549 CrossRef CAS PubMed.
  780. G. Stadlmayr, K. Benakovitsch, B. Gasser, D. Mattanovich and M. Sauer, Genome-scale analysis of library sorting (GALibSo): Isolation of secretion enhancing factors for recombinant protein production in Pichia pastoris, Biotechnol. Bioeng., 2010, 105, 543–555 CrossRef CAS PubMed.
  781. W. Throndset, S. Kim, B. Bower, S. Lantz, B. Kelemen, M. Pepsin, N. Chow, C. Mitchinson and M. Ward, Flow cytometric sorting of the filamentous fungus Trichoderma reesei for improved strains, Enzyme Microb. Technol., 2010, 47, 335–341 CrossRef CAS PubMed.
  782. W. Throndset, B. Bower, R. Caguiat, T. Baldwin and M. Ward, Isolation of a strain of Trichoderma reesei with improved glucoamylase secretion by flow cytometric sorting, Enzyme Microb. Technol., 2010, 47, 342–347 CrossRef CAS PubMed.
  783. B. P. Tracy, S. M. Gaida and E. T. Papoutsakis, Flow cytometry for bacteria: enabling metabolic engineering, synthetic biology and the elucidation of complex phenotypes, Curr. Opin. Biotechnol., 2010, 21, 85–99 CrossRef CAS PubMed.
  784. Y. J. Eun, A. S. Utada, M. F. Copeland, S. Takeuchi and D. B. Weibel, Encapsulating bacteria in agarose microparticles using microfluidics for high-throughput cell analysis and isolation, ACS Chem. Biol., 2011, 6, 260–266 CrossRef CAS PubMed.
  785. E. Fernández-Álvaro, R. Snajdrova, H. Jochens, T. Davids, D. Böttcher and U. T. Bornscheuer, A combination of in vivo selection and cell sorting for the identification of enantioselective biocatalysts, Angew. Chem., Int. Ed., 2011, 50, 8584–8587 CrossRef PubMed.
  786. T. T. Y. Doan, B. Sivaloganathan and J. P. Obbard, Screening of marine microalgae for biodiesel feedstock, Biomass Bioenergy, 2011, 35, 2534–2544 CrossRef CAS PubMed.
  787. S. Binder, G. Schendzielorz, N. Stäbler, K. Krumbach, K. Hoffmann, M. Bott and L. Eggeling, A high-throughput approach to identify genomic variants of bacterial metabolite producers at the single-cell level, Genome Biol., 2012, 13, R40 CrossRef CAS PubMed.
  788. R. Lönneborg, E. Varga and P. Brzezinski, Directed evolution of the transcriptional regulator DntR: isolation of mutants with improved DNT-response, PLoS One, 2012, 7, e29994 Search PubMed.
  789. T. H. Yoo, M. Pogson, B. L. Iverson and G. Georgiou, Directed Evolution of Highly Selective Proteases by Using a Novel FACS-Based Screen that Capitalizes on the p53 Regulator MDM2, ChemBioChem, 2012, 13, 649–653 CrossRef CAS PubMed.
  790. T. Lopes da Silva, J. C. Roseiro and A. Reis, Applications and perspectives of multi-parameter flow cytometry to microbial biofuels production processes, Trends Biotechnol., 2012, 30, 225–232 CrossRef CAS PubMed.
  791. M. Uttamchandani, X. Huang, G. Y. J. Chen and S. Q. Yao, Nanodroplet profiling of enzymatic activities in a microarray, Bioorg. Med. Chem. Lett., 2005, 15, 2135–2139 CrossRef CAS PubMed.
  792. A. A. Gordeev, T. R. Samatov, H. V. Chetverina and A. B. Chetverin, 2D format for screening bacterial cells at the throughput of flow cytometry, Biotechnol. Bioeng., 2011, 108, 2682–2690 CrossRef CAS PubMed.
  793. W. E. Huang, M. Li, R. M. Jarvis, R. Goodacre and S. A. Banwart, Shining light on the microbial world the application of Raman microspectroscopy, Adv. Appl. Microbiol., 2010, 70, 153–186 CAS.
  794. L. Peng, G. Wang, W. Liao, H. Yao, S. Huang and Y. Q. Li, Intracellular ethanol accumulation in yeast cells during aerobic fermentation: a Raman spectroscopic exploration, Lett. Appl. Microbiol., 2010, 51, 632–638 CrossRef CAS PubMed.
  795. J. G. Lees and R. W. Janes, Combining sequence-based prediction methods and circular dichroism and infrared spectroscopic data to improve protein secondary structure determinations, BMC Bioinf., 2008, 9, 24 CrossRef PubMed.
  796. T. van Rossum, S. W. Kengen and J. van der Oost, Reporter-based screening and selection of enzymes, FEBS J., 2013, 280, 2979–2996 CrossRef CAS PubMed.
  797. T. Uchiyama and K. Watanabe, The SIGEX scheme: high throughput screening of environmental metagenomes for the isolation of novel catabolic genes, Biotechnol. Genet. Eng. Rev., 2007, 24, 107–116 CrossRef CAS.
  798. T. Uchiyama and K. Watanabe, Substrate-induced gene expression (SIGEX) screening of metagenome libraries, Nat. Protoc., 2008, 3, 1202–1212 CrossRef CAS PubMed.
  799. T. Uchiyama and K. Miyazaki, Substrate-induced gene expression screening: a method for high-throughput screening of metagenome libraries, Methods Mol. Biol., 2010, 668, 153–168 CAS.
  800. T. Uchiyama and K. Miyazaki, Product-induced gene expression, a product-responsive reporter assay used to screen metagenomic libraries for enzyme-encoding genes, Appl. Environ. Microbiol., 2010, 76, 7029–7035 CrossRef CAS PubMed.
  801. J. Dictenberg, Genetic encoding of fluorescent RNA ensures a bright future for visualizing nucleic acid dynamics, Trends Biotechnol., 2012, 30, 621–626 CrossRef CAS PubMed.
  802. J. R. van der Meer and S. Belkin, Where microbiology meets microengineering: design and applications of reporter bacteria, Nat. Rev. Microbiol., 2010, 8, 511–522 CrossRef CAS PubMed.
  803. M. N. Stojanovic, P. de Prada and D. W. Landry, Aptamer-based folding fluorescent sensor for cocaine, J. Am. Chem. Soc., 2001, 123, 4928–4931 CrossRef CAS PubMed.
  804. M. N. Stojanovic and D. M. Kolpashchikov, Modular aptameric sensors, J. Am. Chem. Soc., 2004, 126, 9266–9270 CrossRef CAS PubMed.
  805. K. Kikuchi, Design, synthesis and biological application of chemical probes for bio-imaging, Chem. Soc. Rev., 2010, 39, 2048–2053 RSC.
  806. K. Kikuchi, Design, synthesis, and biological application of fluorescent sensor molecules for cellular imaging, Adv. Biochem. Eng./Biotechnol., 2010, 119, 63–78 CAS.
  807. S. Okumoto, A. Jones and W. B. Frommer, Quantitative Imaging with Fluorescent Biosensors: Advanced Tools for Spatiotemporal Analysis of Biodynamics in Cells, Annu. Rev. Plant Biol., 2012, 63, 663–706 CrossRef CAS PubMed.
  808. S. Okumoto, Quantitative imaging using genetically encoded sensors for small molecules in plants, Plant J., 2012, 70, 108–117 CrossRef CAS PubMed.
  809. J. S. Paige, K. Y. Wu and S. R. Jaffrey, RNA mimics of green fluorescent protein, Science, 2011, 333, 642–646 CrossRef CAS PubMed.
  810. J. S. Paige, T. Nguyen-Duc, W. Song and S. R. Jaffrey, Fluorescence imaging of cellular metabolites with RNA, Science, 2012, 335, 1194 CrossRef CAS PubMed.
  811. R. L. Strack and S. R. Jaffrey, New approaches for sensing metabolites and proteins in live cells using RNA, Curr. Opin. Chem. Biol., 2013, 17, 651–655 CrossRef CAS PubMed.
  812. X. Xu, J. Zhang, F. Yang and X. Yang, Colorimetric logic gates for small molecules using split/integrated aptamers and unmodified gold nanoparticles, Chem. Commun., 2011, 47, 9435–9437 RSC.
  813. R. Wombacher and V. W. Cornish, Chemical tags: applications in live cell fluorescence imaging, J. Biophotonics, 2011, 4, 391–402 CrossRef CAS PubMed.
  814. C. Jing and V. W. Cornish, Chemical tags for labeling proteins inside living cells, Acc. Chem. Res., 2011, 44, 784–792 CrossRef CAS PubMed.
  815. Chemical proteomics, ed. G. Drewes and M. Bantscheff, Springer, Berlin, 2012 Search PubMed.
  816. A. P. Arkin and D. C. Youvan, Digital imaging spectroscopy, in The photosynthetic reaction center, ed. J. Deisenhofer and J. R. Norris, Academic Press, New York, 1993, vol. 1, pp. 133–155 Search PubMed.
  817. E. R. Goldman and D. C. Youvan, An algorithmically optimized combinatorial library screened by digital imaging spectroscopy, Bio/Technology, 1992, 10, 1557–1561 CrossRef CAS.
  818. H. Joo, A. Arisawa, Z. L. Lin and F. H. Arnold, A high-throughput digital imaging screen for the discovery and directed evolution of oxygenases, Chem. Biol., 1999, 6, 699–706 CrossRef CAS.
  819. M. Alexeeva, A. Enright, M. J. Dawson, M. Mahmoudian and N. J. Turner, Deracemization of alpha-methylbenzylamine using an enzyme obtained by in vitro evolution, Angew. Chem., Int. Ed., 2002, 41, 3177–3180 CrossRef CAS.
  820. M. Alexeeva, R. Carr and N. J. Turner, Directed evolution of enzymes: new biocatalysts for asymmetric synthesis, Org. Biomol. Chem., 2003, 1, 4133–4137 CAS.
  821. S. Delagrave, D. J. Murphy, J. L. Pruss, A. M. Maffia, 3rd, B. L. Marrs, E. J. Bylina, W. J. Coleman, C. L. Grek, M. R. Dilworth, M. M. Yang and D. C. Youvan, Application of a very high-throughput digital imaging screen to evolve the enzyme galactose oxidase, Protein Eng., 2001, 14, 261–267 CrossRef CAS PubMed.
  822. J. C. Weaver, Gel Microdroplets for Microbial Measurement and Screening: Basic Principles, Biotechnol. Bioeng. Symp., 1987, 17, 185–195 Search PubMed.
  823. J. C. Weaver, G. B. Williams, A. Klibanov and A. L. Demain, Gel Microdroplets: Rapid Detection and Enumeration of Individual Microorganisms by their Metabolic Activity, Bio/Technology, 1988, 6, 1084–1089 CrossRef CAS PubMed.
  824. J. C. Weaver, J. G. Bliss, K. T. Powell, G. I. Harrison and G. B. Williams, Rapid Clonal Growth Measurements at the Single-Cell Level: Gel Microdroplets and Flow Cytometry, Bio/Technology, 1991, 9, 873–876 CrossRef CAS PubMed.
  825. J. C. Weaver, J. G. Bliss, G. I. Harrison, K. T. Powell and G. B. Williams, Microdrop Technology: A General Method for Separating Cells by Function and Composition, Methods, 1991, 2, 234–247 CrossRef CAS.
  826. D. S. Tawfik and A. D. Griffiths, Man-made cell-like compartments for molecular evolution, Nat. Biotechnol., 1998, 16, 652–656 CrossRef CAS PubMed.
  827. A. D. Griffiths and D. S. Tawfik, Man-made enzymes - from design to in vitro compartmentalisation, Curr. Opin. Biotechnol., 2000, 11, 338–353 CrossRef CAS.
  828. F. Courtois, L. F. Olguin, G. Whyte, A. B. Theberge, W. T. Huck, F. Hollfelder and C. Abell, Controlling the retention of small molecules in emulsion microdroplets for use in cell-based assays, Anal. Chem., 2009, 81, 3008–3016 CrossRef CAS PubMed.
  829. S. R. A. Devenish, M. Kaltenbach, M. Fischlechner and F. Hollfelder, Droplets as reaction compartments for protein nanotechnology, Methods Mol. Biol., 2013, 996, 269–286 CAS.
  830. W. C. Lu and A. D. Ellington, In vitro selection of proteins via emulsion compartments, Methods, 2013, 60, 75–80 CrossRef CAS PubMed.
  831. M. Fischlechner, Y. Schaerli, M. F. Mohamed, S. Patil, C. Abell and F. Hollfelder, Evolution of enzyme catalysts caged in biomimetic gel-shell beads, Nat. Chem., 2014, 6, 791–796 CrossRef CAS PubMed.
  832. A. Fallah-Araghi, J. C. Baret, M. Ryckelynck and A. D. Griffiths, A completely in vitro ultrahigh-throughput droplet-based microfluidic screening system for protein engineering and directed evolution, Lab Chip, 2012, 12, 882–891 RSC.
  833. A. Grünberger, N. Paczia, C. Probst, G. Schendzielorz, L. Eggeling, S. Noack, W. Wiechert and D. Kohlheyer, A disposable picolitre bioreactor for cultivation and investigation of industrially relevant bacteria on the single cell level, Lab Chip, 2012, 12, 2060–2068 RSC.
  834. I. Levin and A. Aharoni, Evolution in microfluidic droplet, Chem. Biol., 2012, 19, 929–931 CrossRef CAS PubMed.
  835. Y. P. Bai, E. Weibull, H. N. Joensson and H. Andersson-Svahn, Interfacing picoliter droplet microfluidics with addressable microliter compartments using fluorescence activated cell sorting, Sens. Actuators, B, 2014, 194, 249–254 CrossRef CAS PubMed.
  836. F. Ma, Y. Xie, C. Huang, Y. Feng and G. Yang, An improved single cell ultrahigh throughput screening method based on in vitro compartmentalization, PLoS One, 2014, 9, e89785 Search PubMed.
  837. L. Rosenfeld, T. Lin, R. Derda and S. K. Y. Tang, Review and analysis of performance metrics of droplet microfluidics systems, Microfluid. Nanofluid., 2014, 16, 921–939 CrossRef.
  838. S. L. Sjostrom, Y. P. Bai, M. T. Huang, Z. H. Liu, J. Nielsen, H. N. Joensson and H. A. Svahn, High-throughput screening for industrial enzyme production hosts by droplet microfluidics, Lab Chip, 2014, 14, 806–813 RSC.
  839. B. Kintses, L. D. van Vliet, S. R. A. Devenish and F. Hollfelder, Microfluidic droplets: new integrated workflows for biological experiments, Curr. Opin. Chem. Biol., 2010, 14, 548–555 CrossRef CAS PubMed.
  840. B. Kintses, C. Hein, M. F. Mohamed, M. Fischlechner, F. Courtois, C. Laine and F. Hollfelder, Picoliter cell lysate assays in microfluidic droplet compartments for directed enzyme evolution, Chem. Biol., 2012, 19, 1001–1009 CrossRef CAS PubMed.
  841. J. U. Shim, R. T. Ranasinghe, C. A. Smith, S. M. Ibrahim, F. Hollfelder, W. T. Huck, D. Klenerman and C. Abell, Ultrarapid generation of femtoliter microfluidic droplets for single-molecule-counting immunoassays, ACS Nano, 2013, 7, 5955–5964 CrossRef CAS PubMed.
  842. C. A. Smith, X. Li, T. H. Mize, T. D. Sharpe, E. I. Graziani, C. Abell and W. T. S. Huck, Sensitive, high throughput detection of proteins in individual, surfactant-stabilized picoliter droplets using nanoelectrospray ionization mass spectrometry, Anal. Chem., 2013, 85, 3812–3816 CrossRef CAS PubMed.
  843. A. Zinchenko, S. R. Devenish, B. Kintses, P. Y. Colin, M. Fischlechner and F. Hollfelder, One in a million: flow cytometric sorting of single cell-lysate assays in monodisperse picolitre double emulsion droplets for directed evolution, Anal. Chem., 2014, 86, 2526–2533 CrossRef CAS PubMed.
  844. J. J. Agresti, E. Antipov, A. R. Abate, K. Ahn, A. C. Rowat, J. C. Baret, M. Marquez, A. M. Klibanov, A. D. Griffiths and D. A. Weitz, Ultrahigh-throughput screening in drop-based microfluidics for directed evolution, Proc. Natl. Acad. Sci. U. S. A., 2010, 107, 4004–4009 CrossRef CAS PubMed.
  845. J. Sacks, W. Welch, T. Mitchell and H. Wynn, Design and analysis of computer experiments (with discussion), Statist Sci, 1989, 4, 409–435 CrossRef.
  846. J. R. Koza, Genetic programming: on the programming of computers by means of natural selection, MIT Press, Cambridge, Mass, 1992 Search PubMed.
  847. J. R. Koza, Genetic programming II: automatic discovery of reusable programs, MIT Press, Cambridge, Mass, 1994 Search PubMed.
  848. T. Bäck, Evolutionary algorithms in theory and practice, Oxford University Press, Oxford, 1996 Search PubMed.
  849. W. B. Langdon, Genetic programming and data structures: genetic programming+data structures=automatic programming!, Kluwer, Boston, 1998 Search PubMed.
  850. New ideas in optimization, ed. D. Corne, M. Dorigo and F. Glover, McGraw Hill, London, 1999 Search PubMed.
  851. J. R. Koza, F. H. Bennett, M. A. Keane and D. Andre, Genetic Programming III: Darwinian Invention and Problem Solving, Morgan Kaufmann, San Francisco, 1999 Search PubMed.
  852. Evolutionary Computation 1: basic algorithms and operators, ed. T. Bäck, D. B. Fogel and Z. Michalewicz, IOP Publishing, Bristol, 2000 Search PubMed.
  853. Evolutionary Computation 2: advanced algorithms and operators, ed. T. Bäck, D. B. Fogel and Z. Michalewicz, IOP Publishing, Bristol, 2000 Search PubMed.
  854. W. B. Langdon and R. Poli, Foundations of genetic programming, Springer-Verlag, Berlin, 2002 Search PubMed.
  855. J. R. Koza, M. A. Keane and M. J. Streeter, Evolving inventions, Sci Am, 2003, 288, 52–59 CrossRef PubMed.
  856. J. R. Koza, M. A. Keane, M. J. Streeter, W. Mydlowec, J. Yu and G. Lanza, Genetic programming: routine human-competitive machine intelligence, Kluwer, New York, 2003 Search PubMed.
  857. J. Handl and J. Knowles, An evolutionary approach to multiobjective clustering, IEEE Trans. Evol. Comput., 2007, 11, 56–76 CrossRef.
  858. R. Poli, W. B. Langdon and N. F. McPhee, A Field Guide to Genetic Programming, http://www.lulu.com/product/file-download/a-field-guide-to-genetic-programming/2502914, 2009.
  859. D. R. Jones, M. Schonlau and W. J. Welch, Efficient global optimization of expensive black-box functions, J. Global. Opt., 1998, 13, 455–492 CrossRef.
  860. W. M. Patrick, A. E. Firth and J. M. Blackburn, User-friendly algorithms for estimating completeness and diversity in randomized protein-encoding libraries, Protein Eng., 2003, 16, 451–457 CrossRef CAS PubMed.
  861. A. E. Firth and W. M. Patrick, Statistics of protein library construction, Bioinformatics, 2005, 21, 3314–3315 CrossRef CAS PubMed.
  862. W. M. Patrick and A. E. Firth, Strategies and computational tools for improving randomized protein libraries, Biomol. Eng., 2005, 22, 105–112 CrossRef CAS PubMed.
  863. A. E. Firth and W. M. Patrick, GLUE-IT and PEDEL-AA: new programmes for analyzing protein diversity in randomized libraries, Nucleic Acids Res., 2008, 36, W281–W285 CrossRef CAS PubMed.
  864. J. Starrfelt and H. Kokko, Bet-hedging-a triple trade-off between means, variances and correlations, Biol. Rev. Cambridge Philos. Soc., 2012, 87, 742–755 CrossRef PubMed.
  865. A. del Sol Mesa, F. Pazos and A. Valencia, Automatic methods for predicting functionally important residues, J. Mol. Biol., 2003, 326, 1289–1302 CrossRef CAS.
  866. S. Herrgard, S. A. Cammer, B. T. Hoffman, S. Knutson, M. Gallina, J. A. Speir, J. S. Fetrow and S. M. Baxter, Prediction of deleterious functional effects of amino acid mutations using a library of structure-based function descriptors, Proteins, 2003, 53, 806–816 CrossRef CAS PubMed.
  867. G. Lopez, P. Maietta, J. M. Rodriguez, A. Valencia and M. L. Tress, firestar—advances in the prediction of functionally important residues, J. Chem. Inf. Model., 2011, 39, W235–W241 CAS.
  868. T. A. Addington, R. W. Mertz, J. B. Siegel, J. M. Thompson, A. J. Fisher, V. Filkov, N. M. Fleischman, A. A. Suen, C. S. Zhang and M. D. Toney, Janus: Prediction and Ranking of Mutations Required for Functional Interconversion of Enzymes, J. Mol. Biol., 2013, 425, 1378–1389 CrossRef CAS PubMed.
  869. R. J. Fox and G. W. Huisman, Enzyme optimization: moving from blind evolution to statistical exploration of sequence-function space, Trends Biotechnol., 2008, 26, 132–138 CrossRef CAS PubMed.
  870. F. Liang, X. J. Feng, M. Lowry and H. Rabitz, Maximal use of minimal libraries through the adaptive substituent reordering algorithm, J. Phys. Chem. B, 2005, 109, 5842–5854 CrossRef CAS PubMed.
  871. S. R. McAllister, X. J. Feng, P. A. DiMaggio, Jr., C. A. Floudas, J. D. Rabinowitz and H. Rabitz, Descriptor-free molecular discovery in large libraries by adaptive substituent reordering, Bioorg. Med. Chem. Lett., 2008, 18, 5967–5970 CrossRef CAS PubMed.
  872. X. Feng, J. Sanchis, M. T. Reetz and H. Rabitz, Enhancing the efficiency of directed evolution in focused enzyme libraries by the adaptive substituent reordering algorithm, Chemistry, 2012, 18, 5646–5654 CrossRef CAS PubMed.
  873. C. L. Araya and D. M. Fowler, Deep mutational scanning: assessing protein function on a massive scale, Trends Biotechnol., 2011, 29, 435–442 CrossRef CAS PubMed.
  874. N. Fischer, Sequencing antibody repertoires: the next generation, MAbs, 2011, 3, 17–20 CrossRef.
  875. F. Luciani, R. A. Bull and A. R. Lloyd, Next generation deep sequencing and vaccine design: today and tomorrow, Trends Biotechnol., 2012, 30, 443–452 CrossRef CAS PubMed.
  876. T. A. Whitehead, A. Chevalier, Y. Song, C. Dreyfus, S. J. Fleishman, C. De Mattos, C. A. Myers, H. Kamisetty, P. Blair, I. A. Wilson and D. Baker, Optimization of affinity, specificity and function of designed influenza inhibitors using deep sequencing, Nat. Biotechnol., 2012, 30, 543–548 CrossRef CAS PubMed.
  877. P. Baldi and S. Brunak, Bioinformatics: the machine learning approach, MIT Press, Cambridge, MA, 1998 Search PubMed.
  878. Machine learning and data mining. Methods and applications, ed. R. S. Michalski, I. Bratko and M. Kubat, Wiley, Chichester, 1998 Search PubMed.
  879. T. M. Mitchell, Machine learning, McGraw Hill, New York, 1997 Search PubMed.
  880. B. G. Buchanan and E. A. Feigenbaum, DENDRAL and META-DENDRAL: their application dimensions, Artif. Intell. Mater. Process., Proc. Int. Symp., 1978, 11, 5–24 Search PubMed.
  881. B. G. Buchanan, E. A. Feigenbaum and J. Lederberg, On Gray interpretation of the DENDRAL project and programs - myth or mythunderstanding, Chemom. Intell. Lab. Syst., 1988, 5, 33–35 CrossRef CAS.
  882. E. A. Feigenbaum and B. G. Buchanan, DENDRAL and META-DENDRAL: Roots of knowledge systems and expert system applications, Artif. Intell., 1993, 59, 223–240 CrossRef.
  883. J. Lederberg, How DENDRAL was conceived and born, ACM Symp Hist Med Informatics, 1987, http://profiles.nlm.nih.gov/ps/access/BBALYP.pdf.
  884. R. K. Lindsay, B. G. Buchanan, E. A. Feigenbaum and J. Lederberg, DENDRAL – a Case study of the first expert system for scientific hypothesis formation, Artif. Intell. Mater. Process., Proc. Int. Symp., 1993, 61, 209–261 CrossRef.
  885. J. Jonsson, T. Norberg, L. Carlsson, C. Gustafsson and S. Wold, Quantitative sequence-activity models (QSAM)—tools for sequence design, J. Chem. Inf. Model., 1993, 21, 733–739 CAS.
  886. L. Breiman, Statistical modeling: The two cultures, Stat Sci, 2001, 16, 199–215 CrossRef.
  887. J. Liao, M. K. Warmuth, S. Govindarajan, J. E. Ness, R. P. Wang, C. Gustafsson and J. Minshull, Engineering proteinase K using machine learning and synthetic genes, BMC Biotechnol., 2007, 7, 16 CrossRef PubMed.
  888. G. Liang and Z. Li, Scores of generalized base properties for quantitative sequence-activity modelings for E. coli promoters based on support vector machine, J. Mol. Graphics Modell., 2007, 26, 269–281 CrossRef CAS PubMed.
  889. B. Petersen, T. N. Petersen, P. Andersen, M. Nielsen and C. Lundegaard, A generic method for assignment of reliability scores applied to solvent accessibility predictions, BMC Struct. Biol., 2009, 9, 51 CrossRef PubMed.
  890. P. Zhou, X. Chen, Y. Wu and Z. Shang, Gaussian process: an alternative approach for QSAM modeling of peptides, Amino Acids, 2010, 38, 199–212 CrossRef CAS PubMed.
  891. B. A. van den Berg, M. J. T. Reinders, M. Hulsman, L. Wu, H. J. Pel, J. A. Roubos and D. de Ridder, Exploring Sequence Characteristics Related to High-Level Production of Secreted Proteins in Aspergillus niger, PLoS One, 2012, 7, e45869 CAS.
  892. S. Vaidyanathan, D. I. Broadhurst, D. B. Kell and R. Goodacre, Explanatory optimisation of protein mass spectrometry via genetic search, Anal. Chem., 2003, 75, 6679–6686 CrossRef CAS PubMed.
  893. S. Vaidyanathan, D. B. Kell and R. Goodacre, Selective detection of proteins in mixtures using electrospray ionization mass spectrometry: influence of instrumental settings and implications for proteomics, Anal. Chem., 2004, 76, 5024–5032 CrossRef CAS PubMed.
  894. D. Wedge, S. J. Gaskell, S. Hubbard, D. B. Kell, K. W. Lau and C. Eyers, Peptide detectability following ESI mass spectrometry: prediction using genetic programming, in GECCO 2007, ed. D. Thierens et al., ACM, New York, 2007, pp. 2219–2225 Search PubMed.
  895. L. Breiman, Random forests, Machine Learning, 2001, 45, 5–32 CrossRef.
  896. D. M. Fowler, J. J. Stephany and S. Fields, Measuring the activity of protein variants on a large scale using deep mutational scanning, Nat. Protoc., 2014, 9, 2267–2284 CrossRef CAS PubMed.
  897. D. M. Fowler and S. Fields, Deep mutational scanning: a new style of protein science, Nat. Methods, 2014, 11, 801–807 CrossRef CAS PubMed.
  898. J. Cairns, J. Overbaugh and S. Miller, The origin of mutants, Nature, 1988, 335, 142–145 CrossRef CAS PubMed.
  899. P. A. Romero, A. Krause and F. H. Arnold, Navigating the protein fitness landscape with Gaussian processes, Proc. Natl. Acad. Sci. U. S. A., 2013, 110, E193–E201 CrossRef CAS PubMed.
  900. T. Keleti, Basic enzyme kinetics, Akadémiai Kiadó, Budapest, 1986 Search PubMed.
  901. A. Cornish-Bowden, Fundamentals of enzyme kinetics, Portland Press, London, 2nd edn, 1995 Search PubMed.
  902. A. Fersht, Structure and mechanism in protein science: a guide to enzyme catalysis and protein folding, W.H. Freeman, San Francisco, 1999 Search PubMed.
  903. A. R. Fersht, Catalysis, binding and enzyme-substrate complementarity, Proc. R. Soc. London, Ser. B, 1974, 187, 397–407 CrossRef CAS.
  904. W. P. Jencks, Binding energy, specificity, and enzymic catalysis: the Circe effect, Adv. Enzymol. Relat. Areas Mol. Biol., 1975, 43, 219–410 CAS.
  905. A. Whitty, C. A. Fierke and W. P. Jencks, Role of binding energy with coenzyme A in catalysis by 3-oxoacid coenzyme A transferase, Biochemistry, 1995, 34, 11678–11689 CrossRef CAS.
  906. K. Liebeton, A. Zonta, K. Schimossek, M. Nardini, D. Lang, B. W. Dijkstra, M. T. Reetz and K. E. Jaeger, Directed evolution of an enantioselective lipase, Chem. Biol., 2000, 7, 709–718 CrossRef CAS.
  907. L. Greiner, S. Laue, A. Liese and C. Wandrey, Continuous homogeneous asymmetric transfer hydrogenation of ketones: lessons from kinetics, Chemistry, 2006, 12, 1818–1823 CrossRef CAS PubMed.
  908. R. J. Fox and M. D. Clay, Catalytic effectiveness, a measure of enzyme proficiency for industrial applications, Trends Biotechnol., 2009, 27, 137–140 CrossRef CAS PubMed.
  909. D. Porro, B. Gasser, T. Fossati, M. Maurer, P. Branduardi, M. Sauer and D. Mattanovich, Production of recombinant proteins and metabolites in yeasts: when are these systems better than bacterial production systems?, Appl. Microbiol. Biotechnol., 2011, 89, 939–948 CrossRef CAS PubMed.
  910. J. Becker and C. Wittmann, Systems and synthetic metabolic engineering for amino acid production - the heartbeat of industrial strain development, Curr. Opin. Biotechnol., 2012, 23, 718–726 CrossRef CAS PubMed.
  911. R. Dach, J. H. J. Song, F. Roschangar, W. Samstag and C. H. Senanayake, The Eight Criteria Defining a Good Chemical Manufacturing Process, Org. Process Res. Dev., 2012, 16, 1697–1706 CrossRef CAS.
  912. J. R. Knowles and W. J. Albery, Perfection in enzyme catalysis – energetics of triosephosphate isomerase, Acc. Chem. Res., 1977, 10, 105–111 CrossRef CAS.
  913. B. G. Miller and R. Wolfenden, Catalytic proficiency: the unusual case of OMP decarboxylase, Annu. Rev. Biochem., 2002, 71, 847–885 CrossRef CAS PubMed.
  914. A. Radzicka and R. Wolfenden, A proficient enzyme, Science, 1995, 267, 90–93 CAS.
  915. B. G. Miller, A. M. Hassell, R. Wolfenden, M. V. Milburn and S. A. Short, Anatomy of a proficient enzyme: the structure of orotidine 5'-monophosphate decarboxylase in the presence and absence of a potential transition state analog, Proc. Natl. Acad. Sci. U. S. A., 2000, 97, 2011–2016 CrossRef CAS PubMed.
  916. A. Bar-Even, E. Noor, Y. Savir, W. Liebermeister, D. Davidi, D. S. Tawfik and R. Milo, The moderately efficient enzyme: evolutionary and physicochemical trends shaping enzyme parameters, Biochemistry, 2011, 50, 4402–4410 CrossRef CAS PubMed.
  917. R. Milo and R. L. Last, Achieving diversity in the face of constraints: lessons from metabolism, Science, 2012, 336, 1663–1667 CrossRef CAS PubMed.
  918. I. Schomburg, A. Chang, S. Placzek, C. Söhngen, M. Rother, M. Lang, C. Munaretto, S. Ulas, M. Stelzer, A. Grote, M. Scheer and D. Schomburg, BRENDA in 2013: integrated reactions, kinetic data, enzyme function data, improved disease classification: new options and contents in BRENDA, Nucleic Acids Res., 2013, 41, D764–D772 CrossRef CAS PubMed.
  919. U. Wittig, R. Kania, M. Golebiewski, M. Rey, L. Shi, L. Jong, E. Algaa, A. Weidemann, H. Sauer-Danzwith, S. Mir, O. Krebs, M. Bittkowski, E. Wetsch, I. Rojas and W. Müller, SABIO-RK–database for biochemical reaction kinetics, J. Chem. Inf. Model., 2012, 40, D790–D796 CAS.
  920. H. Kacser and J. A. Burns, The molecular basis of dominance., Genetics, 1981, 97, 639–666 CAS.
  921. H. Kacser and J. A. Burns, The control of flux, in Rate Control of Biological Processes. Symposium of the Society for Experimental Biology, ed. D. D. Davies, Cambridge University Press, Cambridge, 1973, vol. 27, pp. 65–104 Search PubMed.
  922. D. B. Kell and H. V. Westerhoff, Metabolic control theory: its role in microbiology and biotechnology., FEMS Microbiol. Rev., 1986, 39, 305–320 CrossRef CAS PubMed.
  923. D. B. Kell and H. V. Westerhoff, Towards a rational approach to the optimization of flux in microbial biotransformations., Trends Biotechnol., 1986, 4, 137–142 CrossRef CAS.
  924. D. A. Fell, Understanding the control of metabolism, Portland Press, London, 1996 Search PubMed.
  925. R. Heinrich and S. Schuster, The regulation of cellular systems., Chapman & Hall, New York, 1996 Search PubMed.
  926. C. K. Savile, J. M. Janey, E. C. Mundorff, J. C. Moore, S. Tam, W. R. Jarvis, J. C. Colbeck, A. Krebber, F. J. Fleitz, J. Brands, P. N. Devine, G. W. Huisman and G. J. Hughes, Biocatalytic asymmetric synthesis of chiral amines from ketones applied to sitagliptin manufacture, Science, 2010, 329, 305–309 CrossRef CAS PubMed.
  927. Y. Suzuki, K. Asada, J. Miyazaki, T. Tomita, T. Kuzuyama and M. Nishiyama, Enhancement of the latent 3-isopropylmalate dehydrogenase activity of promiscuous homoisocitrate dehydrogenase by directed evolution, Biochem. J., 2010, 431, 401–410 CAS.
  928. S. J. Benkovic and S. Hammes-Schiffer, A perspective on enzyme catalysis, Science, 2003, 301, 1196–1202 CrossRef CAS PubMed.
  929. M. Garcia-Viloca, J. Gao, M. Karplus and D. G. Truhlar, How enzymes work: analysis by modern rate theory and computer simulations, Science, 2004, 303, 186–195 CrossRef CAS PubMed.
  930. G. G. Hammes, S. J. Benkovic and S. Hammes-Schiffer, Flexibility, diversity, and cooperativity: pillars of enzyme catalysis, Biochemistry, 2011, 50, 10422–10430 CrossRef CAS PubMed.
  931. T. C. Bruice, Computational approaches: Reaction trajectories, structures, and atomic motions. Enzyme reactions and proficiency, Chem. Rev., 2006, 106, 3119–3139 CrossRef CAS PubMed.
  932. J. L. Gao, S. H. Ma, D. T. Major, K. Nam, J. Z. Pu and D. G. Truhlar, Mechanisms and free energies of enzymatic reactions, Chem. Rev., 2006, 106, 3188–3209 CrossRef CAS PubMed.
  933. M. H. Olsson, W. W. Parson and A. Warshel, Dynamical contributions to enzyme catalysis: critical tests of a popular hypothesis, Chem. Rev., 2006, 106, 1737–1756 CrossRef CAS PubMed.
  934. P. R. Carey, Spectroscopic characterization of distortion in enzyme complexes, Chem. Rev., 2006, 106, 3043–3054 CrossRef CAS PubMed.
  935. A. Warshel, P. K. Sharma, M. Kato, Y. Xiang, H. B. Liu and M. H. M. Olsson, Electrostatic basis for enzyme catalysis, Chem. Rev., 2006, 106, 3210–3235 CrossRef CAS PubMed.
  936. A. V. Pisliakov, J. Cao, S. C. Kamerlin and A. Warshel, Enzyme millisecond conformational dynamics do not catalyze the chemical step, Proc. Natl. Acad. Sci. U. S. A., 2009, 106, 17359–17364 CrossRef CAS PubMed.
  937. S. C. Kamerlin and A. Warshel, At the dawn of the 21st century: Is dynamics the missing link for understanding enzyme catalysis?, Proteins, 2010, 78, 1339–1375 CAS.
  938. A. J. Adamczyk, J. Cao, S. C. Kamerlin and A. Warshel, Catalysis by dihydrofolate reductase and other enzymes arises from electrostatic preorganization, not conformational motions, Proc. Natl. Acad. Sci. U. S. A., 2011, 108, 14115–14120 CrossRef CAS PubMed.
  939. M. P. Frushicheva, M. J. Mills, P. Schopf, M. K. Singh, R. B. Prasad and A. Warshel, Computer aided enzyme design and catalytic concepts, Curr. Opin. Chem. Biol., 2014, 21C, 56–62 CrossRef PubMed.
  940. Z. D. Nagel and J. P. Klinman, Tunneling and dynamics in enzymatic hydride transfer, Chem. Rev., 2006, 106, 3095–3118 CrossRef CAS PubMed.
  941. J. Z. Pu, J. L. Gao and D. G. Truhlar, Multidimensional tunneling, recrossing, and the transmission coefficient for enzymatic reactions, Chem. Rev., 2006, 106, 3140–3169 CrossRef CAS PubMed.
  942. S. Hay and N. S. Scrutton, Incorporation of hydrostatic pressure into models of hydrogen tunneling highlights a role for pressure-modulated promoting vibrations, Biochemistry, 2008, 47, 9880–9887 CrossRef CAS PubMed.
  943. S. Hay, L. O. Johannissen, M. J. Sutcliffe and N. S. Scrutton, Barrier compression and its contribution to both classical and quantum mechanical aspects of enzyme catalysis, Biophys. J., 2010, 98, 121–128 CrossRef CAS PubMed.
  944. C. R. Pudney, L. O. Johannissen, M. J. Sutcliffe, S. Hay and N. S. Scrutton, Direct analysis of donor-acceptor distance and relationship to isotope effects and the force constant for barrier compression in enzymatic H-tunneling reactions, J. Am. Chem. Soc., 2010, 132, 11329–11335 CrossRef CAS PubMed.
  945. S. Hay and N. S. Scrutton, Good vibrations in enzyme-catalysed reactions, Nat. Chem., 2012, 4, 161–168 CrossRef CAS PubMed.
  946. S. Hay, L. O. Johannissen, P. Hothi, M. J. Sutcliffe and N. S. Scrutton, Pressure effects on enzyme-catalyzed quantum tunneling events arise from protein-specific structural and dynamic changes, J. Am. Chem. Soc., 2012, 134, 9749–9754 CrossRef CAS PubMed.
  947. M. Widersten, Protein engineering for development of new hydrolytic biocatalysts, Curr. Opin. Chem. Biol., 2014, 21C, 42–47 CrossRef PubMed.
  948. M. Fuxreiter and L. Mones, The role of reorganization energy in rational enzyme design, Curr. Opin. Chem. Biol., 2014, 21C, 34–41 CrossRef PubMed.
  949. B. Gavish and M. M. Werber, Viscosity-dependent structural fluctuations in enzyme catalysis, Biochemistry, 1979, 18, 1269–1275 CrossRef CAS.
  950. B. Gavish, Position-dependent viscosity effects on rate coefficients, Phys. Rev. Lett., 1980, 44, 1160–1163 CrossRef CAS.
  951. D. Beece, L. Eisenstein, H. Frauenfelder, D. Good, M. C. Marden, L. Reinisch, A. H. Reynolds, L. B. Sorensen and K. T. Yue, Solvent viscosity and protein dynamics, Biochemistry, 1980, 19, 5147–5157 CrossRef CAS.
  952. D. B. Kell, Enzymes As Energy Funnels, Trends Biochem. Sci., 1982, 7, 349 CrossRef.
  953. G. R. Welch, B. Somogyi and S. Damjanovich, The role of protein fluctuations in enzyme action: a review, Prog. Biophys. Mol. Biol., 1982, 39, 109–146 CrossRef CAS.
  954. B. Somogyi, G. R. Welch and S. Damjanovich, The dynamic basis of energy transduction in enzymes, Biochim. Biophys. Acta, 1984, 768, 81–112 CrossRef CAS.
  955. D. Vitkup, D. Ringe, G. A. Petsko and M. Karplus, Solvent mobility and the protein 'glass' transition, Nat. Struct. Biol., 2000, 7, 34–38 CrossRef CAS PubMed.
  956. R. M. Daniel, R. V. Dunn, J. L. Finney and J. C. Smith, The role of dynamics in enzyme activity, Annu. Rev. Biophys. Biomol. Struct., 2003, 32, 69–92 CrossRef CAS PubMed.
  957. J. E. Basner and S. D. Schwartz, How enzyme dynamics helps catalyze a reaction in atomic detail: a transition path sampling study, J. Am. Chem. Soc., 2005, 127, 13822–13831 CrossRef CAS PubMed.
  958. D. Antoniou, J. Basner, S. Nunez and S. D. Schwartz, Computational and theoretical methods to explore the relation between enzyme dynamics and catalysis, Chem. Rev., 2006, 106, 3170–3187 CrossRef CAS PubMed.
  959. R. Callender and R. B. Dyer, Advances in time-resolved approaches to characterize the dynamical nature of enzymatic catalysis, Chem. Rev., 2006, 106, 3031–3042 CrossRef CAS PubMed.
  960. I. J. Finkelstein, A. M. Massari and M. D. Fayer, Viscosity-dependent protein dynamics, Biophys. J., 2007, 92, 3652–3662 CrossRef CAS PubMed.
  961. H. Frauenfelder, P. W. Fenimore and R. D. Young, Protein dynamics and function: Insights from the energy landscape and solvent slaving, IUBMB Life, 2007, 59, 506–512 CrossRef CAS PubMed.
  962. K. Henzler-Wildman and D. Kern, Dynamic personalities of proteins, Nature, 2007, 450, 964–972 CrossRef CAS PubMed.
  963. K. A. Henzler-Wildman, M. Lei, V. Thai, S. J. Kerns, M. Karplus and D. Kern, A hierarchy of timescales in protein dynamics is linked to enzyme catalysis, Nature, 2007, 450, 913–916 CrossRef CAS PubMed.
  964. K. A. Henzler-Wildman, V. Thai, M. Lei, M. Ott, M. Wolf-Watz, T. Fenn, E. Pozharski, M. A. Wilson, G. A. Petsko, M. Karplus, C. G. Hubner and D. Kern, Intrinsic motions along an enzymatic reaction trajectory, Nature, 2007, 450, 838–844 CrossRef CAS PubMed.
  965. H. Frauenfelder, G. Chen, J. Berendzen, P. W. Fenimore, H. Jansson, B. H. McMahon, I. R. Stroe, J. Swenson and R. D. Young, A unified model of protein dynamics, Proc. Natl. Acad. Sci. U. S. A., 2009, 106, 5129–5134 CrossRef CAS PubMed.
  966. S. J. Benkovic, G. G. Hammes and S. Hammes-Schiffer, Free-energy landscape of enzyme catalysis, Biochemistry, 2008, 47, 3317–3321 CrossRef CAS PubMed.
  967. R. K. Eppler, E. P. Hudson, S. D. Chase, J. S. Dordick, J. A. Reimer and D. S. Clark, Biocatalyst activity in nonaqueous environments correlates with centisecond-range protein motions, Proc. Natl. Acad. Sci. U. S. A., 2008, 105, 15672–15677 CrossRef CAS PubMed.
  968. D. D. Boehr, R. Nussinov and P. E. Wright, The role of dynamic conformational ensembles in biomolecular recognition, Nat. Chem. Biol., 2009, 5, 789–796 CrossRef CAS PubMed.
  969. S. D. Schwartz and V. L. Schramm, Enzymatic transition states and dynamic motion in barrier crossing, Nat. Chem. Biol., 2009, 5, 551–558 CrossRef CAS PubMed.
  970. I. Bahar, T. R. Lezon, L. W. Yang and E. Eyal, Global dynamics of proteins: bridging between structure and function, Annu. Rev. Biophys., 2010, 39, 23–42 CrossRef CAS PubMed.
  971. P. Csermely, R. Palotai and R. Nussinov, Induced fit, conformational selection and independent dynamic segments: an extended view of binding events, Trends Biochem. Sci., 2010, 35, 539–546 CrossRef CAS PubMed.
  972. A. E. Sitnitsky, Solvent viscosity dependence for enzymatic reactions, Phys. A, 2008, 387, 5483–5497 CrossRef CAS PubMed.
  973. A. E. Sitnitsky, Model for solvent viscosity effect on enzymatic reactions, Chem. Phys., 2010, 369, 37–42 CrossRef CAS PubMed.
  974. G. Bhabha, J. Lee, D. C. Ekiert, J. Gam, I. A. Wilson, H. J. Dyson, S. J. Benkovic and P. E. Wright, A Dynamic Knockout Reveals That Conformational Fluctuations Influence the Chemical Step of Enzyme Catalysis, Science, 2011, 332, 234–238 CrossRef CAS PubMed.
  975. M. Shushanyan, D. E. Khoshtariya, T. Tretyakova, M. Makharadze and R. van Eldik, Diverse role of conformational dynamics in carboxypeptidase A-driven peptide and ester hydrolyses: disclosing the “perfect induced fit” and “protein local unfolding” pathways by altering protein stability, Biopolymers, 2011, 95, 852–870 CrossRef CAS PubMed.
  976. A. R. Jones, C. Levy, S. Hay and N. S. Scrutton, Relating localized protein motions to the reaction coordinate in coenzyme B12-dependent enzymes, FEBS J., 2013, 280, 2997–3008 CrossRef CAS PubMed.
  977. Dynamics in enzyme catalysis, ed. J. P. Klinman and S. Hammes-Schiffer, Springer, Berlin, 2013 Search PubMed.
  978. K. Świderek, J. Javier Ruiz-Pernía, V. Moliner and I. Tuñon, Heavy enzymes-experimental and computational insights in enzyme dynamics, Curr. Opin. Chem. Biol., 2014, 21C, 11–18 CrossRef PubMed.
  979. A. S. Davydov, Solitons and energy transfer along protein molecules, J. Theor. Biol., 1977, 66, 379–387 CrossRef CAS.
  980. A. S. Davydov, Excitons and solitons in molecular systems, Int. Rev. Cytol., 1987, 106, 183–225 CAS.
  981. A. Ansari, J. Berendzen, S. F. Bowne, H. Frauenfelder, I. E. Iben, T. B. Sauke, E. Shyamsunder and R. D. Young, Protein states and proteinquakes, Proc. Natl. Acad. Sci. U. S. A., 1985, 82, 5000–5004 CrossRef CAS.
  982. G. Dadusc, J. P. Ogilvie, P. Schulenberg, U. Marvet and R. J. Miller, Diffractive optics-based heterodyne-detected four-wave mixing signals of protein motion: from “protein quakes” to ligand escape for myoglobin, Proc. Natl. Acad. Sci. U. S. A., 2001, 98, 6110–6115 CrossRef CAS PubMed.
  983. D. Arnlund, L. C. Johansson, C. Wickstrand, A. Barty, G. J. Williams, E. Malmerberg, J. Davidsson, D. Milathianaki, D. P. DePonte, R. L. Shoeman, D. Wang, D. James, G. Katona, S. Westenhoff, T. A. White, A. Aquila, S. Bari, P. Berntsen, M. Bogan, T. B. van Driel, R. B. Doak, K. S. Kjaer, M. Frank, R. Fromme, I. Grotjohann, R. Henning, M. S. Hunter, R. A. Kirian, I. Kosheleva, C. Kupitz, M. Liang, A. V. Martin, M. M. Nielsen, M. Messerschmidt, M. M. Seibert, J. Sjohamn, F. Stellato, U. Weierstall, N. A. Zatsepin, J. C. Spence, P. Fromme, I. Schlichting, S. Boutet, G. Groenhof, H. N. Chapman and R. Neutze, Visualizing a protein quake with time-resolved X-ray scattering at a free-electron laser, Nat. Methods, 2014, 11, 923–926 CrossRef CAS PubMed.
  984. E. Fuglebakk, J. Echave and N. Reuter, Measuring and comparing structural fluctuation patterns in large protein datasets, Bioinformatics, 2012, 28, 2431–2440 CrossRef CAS PubMed.
  985. J. E. Jimenez-Roldan, R. B. Freedman, R. A. Romer and S. A. Wells, Rapid simulation of protein motion: merging flexibility, rigidity and normal mode analyses, Phys. Biol., 2012, 9, 016008 CrossRef CAS PubMed.
  986. R. M. Pelis, W. M. Suhre and S. H. Wright, Functional influence of N-glycosylation in OCT2-mediated tetraethylammonium transport, Am. J. Physiol.: Renal, Fluid Electrolyte Physiol., 2006, 290, F1118–F1126 CrossRef CAS PubMed.
  987. G. M. Süel, S. W. Lockless, M. A. Wall and R. Ranganathan, Evolutionarily conserved networks of residues mediate allosteric communication in proteins, Nat. Struct. Biol., 2003, 10, 59–69 CrossRef PubMed.
  988. K. F. Wong, T. Selzer, S. J. Benkovic and S. Hammes-Schiffer, Impact of distal mutations on the network of coupled motions correlated to hydride transfer in dihydrofolate reductase, Proc. Natl. Acad. Sci. U. S. A., 2005, 102, 6807–6812 CrossRef CAS PubMed.
  989. K. L. Morley and R. J. Kazlauskas, Improving enzyme properties: when are closer mutations better?, Trends Biotechnol., 2005, 23, 231–237 CrossRef CAS PubMed.
  990. J. Paramesvaran, E. G. Hibbert, A. J. Russell and P. A. Dalby, Distributions of enzyme residues yielding mutants with improved substrate specificities from two different directed evolution strategies, Protein Eng., Des. Sel., 2009, 22, 401–411 CrossRef CAS PubMed.
  991. N. G. H. Leferink, S. V. Antonyuk, J. A. Houwman, N. S. Scrutton, R. R. Eady and S. S. Hasnain, Impact of residues remote from the catalytic centre on enzyme catalysis of copper nitrite reductase, Nat. Commun., 2014, 5, 4395 Search PubMed.
  992. R. Fasan, Y. T. Meharenna, C. D. Snow, T. L. Poulos and F. H. Arnold, Evolutionary history of a specialized P450 propane monooxygenase, J. Mol. Biol., 2008, 383, 1069–1080 CrossRef CAS PubMed.
  993. N. Preiswerk, T. Beck, J. D. Schulz, P. Milovnik, C. Mayer, J. B. Siegel, D. Baker and D. Hilvert, Impact of scaffold rigidity on the design and evolution of an artificial Diels-Alderase, Proc. Natl. Acad. Sci. U. S. A., 2014, 111, 8013–8018 CrossRef CAS PubMed.
  994. X. Qi, Y. Chen, K. Jiang, W. Zuo, Z. Luo, Y. Wei, L. Du, H. Wei, R. Huang and Q. Du, Saturation-mutagenesis in two positions distant from active site of a Klebsiella pneumoniae glycerol dehydratase identifies some highly active mutants, J. Biotechnol., 2009, 144, 43–50 CrossRef CAS PubMed.
  995. D. L. Siehl, L. A. Castle, R. Gorton and R. J. Keenan, The molecular basis of glyphosate resistance by an optimized microbial acetyltransferase, J. Biol. Chem., 2007, 282, 11446–11455 CrossRef CAS PubMed.
  996. C. M. Cho, A. Mulchandani and W. Chen, Bacterial cell surface display of organophosphorus hydrolase for selective screening of improved hydrolysis of organophosphate nerve agents, Appl. Environ. Microbiol., 2002, 68, 2026–2030 CrossRef CAS.
  997. S. Oue, A. Okamoto, T. Yano and H. Kagamiyama, Redesigning the substrate specificity of an enzyme by cumulative effects of the mutations of non-active site residues, J. Biol. Chem., 1999, 274, 2344–2349 CrossRef CAS PubMed.
  998. K. Lindorff-Larsen, S. Piana, R. O. Dror and D. E. Shaw, How fast-folding proteins fold, Science, 2011, 334, 517–520 CrossRef CAS PubMed.
  999. S. Piana, K. Sarkar, K. Lindorff-Larsen, M. Guo, M. Gruebele and D. E. Shaw, Computational design and experimental testing of the fastest-folding beta-sheet protein, J. Mol. Biol., 2011, 405, 43–48 CrossRef CAS PubMed.
  1000. S. Piana, K. Lindorff-Larsen and D. E. Shaw, Atomic-level description of ubiquitin folding, Proc. Natl. Acad. Sci. U. S. A., 2013, 110, 5915–5920 CrossRef CAS PubMed.
  1001. A. Raval, S. Piana, M. P. Eastwood, R. O. Dror and D. E. Shaw, Refinement of protein structure homology models via long, all-atom molecular dynamics simulations, Proteins, 2012, 80, 2071–2079 CAS.
  1002. D. E. Shaw, M. M. Deneroff, R. O. Dror, J. S. Kuskin, R. H. Larson, J. K. Salmon, C. Young, B. Batson, K. J. Bowers, J. C. Chao, M. P. Eastwood, J. Gagliardo, J. P. Grossman, C. R. Ho, D. J. Ierardi, I. Kolossvary, J. L. Klepeis, T. Layman, C. Mcleavey, M. A. Moraes, R. Mueller, E. C. Priest, Y. B. Shan, J. Spengler, M. Theobald, B. Towles and S. C. Wang, Anton, a special-purpose machine for molecular dynamics simulation, Commun. ACM, 2008, 51, 91–97 CrossRef.
  1003. T. Schwede, Protein modeling: what happened to the “protein structure gap”?, Structure, 2013, 21, 1531–1540 CrossRef CAS PubMed.
  1004. K. A. Dill and J. L. MacCallum, The protein-folding problem, 50 years on, Science, 2012, 338, 1042–1046 CrossRef CAS PubMed.
  1005. F. Khatib, F. Dimaio, S. Cooper, M. Kazmierczyk, M. Gilski, S. Krzywda, H. Zabranska, I. Pichova, J. Thompson, Z. Popovic, M. Jaskolski and D. Baker, Crystal structure of a monomeric retroviral protease solved by protein folding game players, Nat. Struct. Mol. Biol., 2011, 18, 1175–1177 CAS.
  1006. E. H. Kellogg, O. F. Lange and D. Baker, Evaluation and optimization of discrete state models of protein folding, J. Phys. Chem. B, 2012, 116, 11405–11413 CrossRef CAS PubMed.
  1007. D. S. Marks, T. A. Hopf and C. Sander, Protein structure prediction from sequence variation, Nat. Biotechnol., 2012, 30, 1072–1080 CrossRef CAS PubMed.
  1008. W. R. Taylor, D. T. Jones and M. I. Sadowski, Protein topology from predicted residue contacts, Protein Sci., 2012, 21, 299–305 CrossRef CAS PubMed.
  1009. D. T. Jones, D. W. Buchan, D. Cozzetto and M. Pontil, PSICOV: Precise structural contact prediction using sparse inverse covariance estimation on large multiple sequence alignments, Bioinformatics, 2012, 28, 184–190 CrossRef CAS PubMed.
  1010. T. Nugent and D. T. Jones, Accurate de novo structure prediction of large transmembrane protein domains using fragment-assembly and correlated mutation analysis, Proc. Natl. Acad. Sci. U. S. A., 2012, 109, E1540–E1547 CrossRef CAS PubMed.
  1011. T. Kosciolek and D. T. Jones, De novo structure prediction of globular proteins aided by sequence variation-derived contacts, PLoS One, 2014, 9, e92197 Search PubMed.
  1012. C. Andreini, I. Bertini, G. Cavallaro, G. L. Holliday and J. M. Thornton, Metal ions in biological catalysis: from enzyme databases to general principles, JBIC, J. Biol. Inorg. Chem., 2008, 13, 1205–1218 CrossRef CAS PubMed.
  1013. D. B. Kell, Iron behaving badly: inappropriate iron chelation as a major contributor to the aetiology of vascular and other progressive inflammatory and degenerative diseases, BMC Med. Genomics, 2009, 2, 2 CrossRef PubMed.
  1014. D. B. Kell, Towards a unifying, systems biology understanding of large-scale cellular death and destruction caused by poorly liganded iron: Parkinson’s, Huntington’s, Alzheimer’s, prions, bactericides, chemical toxicology and others as examples, Arch. Toxicol., 2010, 577, 825–889 CrossRef PubMed.
  1015. D. B. Kell and E. Pretorius, Serum ferritin is an important disease marker, and is mainly a leakage product from damaged cells, Metallomics, 2014, 6, 748–773 RSC.
  1016. M. T. Reetz, J. J. Peyralans, A. Maichele, Y. Fu and M. Maywald, Directed evolution of hybrid enzymes: Evolving enantioselectivity of an achiral Rh-complex anchored to a protein, Chem. Commun., 2006, 4318–4320 RSC.
  1017. M. T. Reetz, M. Rentzsch, A. Pletsch, M. Maywald, P. Maiwald, J. J. P. Peyralans, A. Maichele, Y. Fu, N. Jiao, F. Hollmann, R. Mondière and A. Taglieber, Directed evolution of enantioselective hybrid catalysts: a novel concept in asymmetric catalysis, Tetrahedron, 2007, 63, 6404–6414 CrossRef CAS PubMed.
  1018. M. T. Reetz, Directed evolution of selective enzymes and hybrid catalysts, Tetrahedron, 2002, 58, 6595–6602 CrossRef CAS.
  1019. M. T. Reetz, M. Rentzsch, A. Pletsch and M. Maywald, Towards the directed evolution of hybrid catalysts, Chimia, 2002, 56, 721–723 CrossRef CAS.
  1020. D. E. Benson, M. S. Wisz and H. W. Hellinga, Rational design of nascent metalloenzymes, Proc. Natl. Acad. Sci. U. S. A., 2000, 97, 6292–6297 CrossRef CAS.
  1021. J. D. Bridgewater, J. Lim and R. W. Vachet, Transition metal-peptide binding studied by metal-catalyzed oxidation reactions and mass spectrometry, Anal. Chem., 2006, 78, 2432–2438 CrossRef CAS PubMed.
  1022. A. K. Petros, A. R. Reddi, M. L. Kennedy, A. G. Hyslop and B. R. Gibney, Femtomolar Zn(II) affinity in a peptide-based ligand designed to model thiolate-rich metalloprotein active sites, Inorg. Chem., 2006, 45, 9941–9958 CrossRef CAS PubMed.
  1023. A. Mantion, L. Massuger, P. Rabu, C. Palivan, L. B. McCusker and A. Taubert, Metal-peptide frameworks (MPFs): “bioinspired” metal organic frameworks, J. Am. Chem. Soc., 2008, 130, 2517–2526 CrossRef CAS PubMed.
  1024. G. D. Pirngruber, L. Frunz and M. Lüchinger, The characterisation and catalytic properties of biomimetic metal-peptide complexes immobilised on mesoporous silica, Phys. Chem. Chem. Phys., 2009, 11, 2928–2938 RSC.
  1025. J. T. Pedersen, K. Teilum, N. H. Heegaard, J. Ostergaard, H. W. Adolph and L. Hemmingsen, Rapid formation of a preoligomeric peptide-metal-peptide complex following copper(II) binding to amyloid beta peptides, Angew. Chem., Int. Ed., 2011, 50, 2532–2535 CrossRef CAS PubMed.
  1026. T. Tanaka, T. Mizuno, S. Fukui, H. Hiroaki, J. Oku, K. Kanaori, K. Tajima and M. Shirakawa, Two-metal ion, Ni(II) and Cu(II), binding alpha-helical coiled coil peptide, J. Am. Chem. Soc., 2004, 126, 14023–14028 CrossRef CAS PubMed.
  1027. C. Tamerler, D. Khatayevich, M. Gungormus, T. Kacar, E. E. Oren, M. Hnilova and M. Sarikaya, Molecular biomimetics: GEPI-based biological routes to technology, Biopolymers, 2010, 94, 78–94 CrossRef CAS PubMed.
  1028. R. L. Koder, J. L. Anderson, L. A. Solomon, K. S. Reddy, C. C. Moser and P. L. Dutton, Design and engineering of an O2 transport protein, Nature, 2009, 458, 305–309 CrossRef CAS PubMed.
  1029. A. F. Peacock, O. Iranzo and V. L. Pecoraro, Harnessing natures ability to control metal ion coordination geometry using de novo designed peptides, Dalton Trans., 2009, 2271–2280 RSC.
  1030. O. Iranzo, S. Chakraborty, L. Hemmingsen and V. L. Pecoraro, Controlling and Fine Tuning the Physical Properties of Two Identical Metal Coordination Sites in De Novo Designed Three Stranded Coiled Coil Peptides, J. Am. Chem. Soc., 2011, 133, 239–251 CrossRef CAS PubMed.
  1031. P. Braun, E. Goldberg, C. Negron, M. von Jan, F. Xu, V. Nanda, R. L. Koder and D. Noy, Design principles for chlorophyll-binding sites in helical proteins, Proteins, 2010, 79, 463–476 CrossRef PubMed.
  1032. F. V. Cochran, S. P. Wu, W. Wang, V. Nanda, J. G. Saven, M. J. Therien and W. F. DeGrado, Computational de novo design and characterization of a four-helix bundle protein that selectively binds a nonbiological cofactor, J. Am. Chem. Soc., 2005, 127, 1346–1347 CrossRef CAS PubMed.
  1033. K. Kuroda and M. Ueda, Molecular design of the microbial cell surface toward the recovery of metal ions, Curr. Opin. Biotechnol., 2011, 22, 427–433 CrossRef CAS PubMed.
  1034. J. M. González, M. R. Meini, P. E. Tomatis, F. J. Medrano Martín, J. A. Cricco and A. J. Vila, Metallo-beta-lactamases withstand low Zn(II) conditions by tuning metal-ligand interactions, Nat. Chem. Biol., 2012, 8, 698–700 CrossRef PubMed.
  1035. I. Sóvágó, C. Kállay and K. Várnagy, Peptides as complexing agents: Factors influencing the structure and thermodynamic stability of peptide complexes, Coord. Chem. Rev., 2012, 256, 2225–2233 CrossRef PubMed.
  1036. Y. Lu, N. Yeung, N. Sieracki and N. M. Marshall, Design of functional metalloproteins, Nature, 2009, 460, 855–862 CrossRef CAS PubMed.
  1037. K. L. Harris, S. Lim and S. J. Franklin, Of folding and function: understanding active-site context through metalloenzyme design, Inorg. Chem., 2006, 45, 10002–10012 CrossRef CAS PubMed.
  1038. K. E. Sapsford, W. R. Algar, L. Berti, K. B. Gemmill, B. J. Casey, E. Oh, M. H. Stewart and I. L. Medintz, Functionalizing nanoparticles with biological molecules: developing chemistries that facilitate nanotechnology, Chem. Rev., 2013, 113, 1904–2074 CrossRef CAS PubMed.
  1039. T. Happe and A. Hemschemeier, Metalloprotein mimics - old tools in a new light, Trends Biotechnol., 2014, 32, 170–176 CrossRef CAS PubMed.
  1040. M. Dürrenberger and T. R. Ward, Recent achievments in the design and engineering of artificial metalloenzymes, Curr. Opin. Chem. Biol., 2014, 19, 99–106 CrossRef PubMed.
  1041. J. Bos and G. Roelfes, Artificial metalloenzymes for enantioselective catalysis, Curr. Opin. Chem. Biol., 2014, 19, 135–143 CrossRef CAS PubMed.
  1042. I. D. Petrik, J. Liu and Y. Lu, Metalloenzyme design and engineering through strategic modifications of native protein scaffolds, Curr. Opin. Chem. Biol., 2014, 19, 67–75 CrossRef CAS PubMed.
  1043. A. J. Hickman and M. S. Sanford, High-valent organometallic copper and palladium in catalysis, Nature, 2012, 484, 177–185 CrossRef CAS PubMed.
  1044. A. J. Reig, M. M. Pires, R. A. Snyder, Y. Wu, H. Jo, D. W. Kulp, S. E. Butch, J. R. Calhoun, T. Szyperski, E. I. Solomon and W. F. DeGrado, Alteration of the oxygen-dependent reactivity of de novo Due Ferri proteins, Nat. Chem., 2012, 4, 900–906 CrossRef CAS PubMed.
  1045. T. Mizuno, K. Murao, Y. Tanabe, M. Oda and T. Tanaka, Metal-ion-dependent GFP emission in vivo by combining a circularly permutated green fluorescent protein with an engineered metal-ion-binding coiled-coil, J. Am. Chem. Soc., 2007, 129, 11378–11383 CrossRef CAS PubMed.
  1046. W. Bae, A. Mulchandani and W. Chen, Cell surface display of synthetic phytochelatins using ice nucleation protein for enhanced heavy metal bioaccumulation, J. Inorg. Biochem., 2002, 88, 223–227 CrossRef CAS.
  1047. C. S. Cutler, H. M. Hennkens, N. Sisay, S. Huclier-Markai and S. S. Jurisson, Radiometals for combined imaging and therapy, Chem. Rev., 2012, 113, 858–883 CrossRef PubMed.
  1048. R. Ferreirós-Martínez, D. Esteban-Gómez, C. Platas-Iglesias, A. de Blas and T. Rodríguez-Blas, Zn(II), Cd(II) and Pb(II) complexation with pyridinecarboxylate containing ligands, Dalton Trans., 2008, 5754–5765 RSC.
  1049. E. Boros, C. L. Ferreira, J. F. Cawthray, E. W. Price, B. O. Patrick, D. W. Wester, M. J. Adam and C. Orvig, Acyclic chelate with ideal properties for 68Ga PET imaging agent elaboration, J. Am. Chem. Soc., 2010, 132, 15726–15733 CrossRef CAS PubMed.
  1050. S. R. Stürzenbaum, M. Hockner, A. Panneerselvam, J. Levitt, J. S. Bouillard, S. Taniguchi, L. A. Dailey, R. Ahmad Khanbeigi, E. V. Rosca, M. Thanou, K. Suhling, A. V. Zayats and M. Green, Biosynthesis of luminescent quantum dots in an earthworm, Nat. Nanotechnol., 2013, 8, 57–60 CrossRef PubMed.
  1051. C. Lo, M. R. Ringenberg, D. Gnandt, Y. Wilson and T. R. Ward, Artificial metalloenzymes for olefin metathesis based on the biotin-(strept)avidin technology, Chem. Commun., 2011, 47, 12065–12067 RSC.
  1052. T. R. Ward, Artificial metalloenzymes based on the biotin-avidin technology: enantioselective catalysis and beyond, Acc. Chem. Res., 2011, 44, 47–57 CrossRef CAS PubMed.
  1053. T. Heinisch and T. R. Ward, Design strategies for the creation of artificial metalloenzymes, Curr. Opin. Chem. Biol., 2010, 14, 184–199 CrossRef CAS PubMed.
  1054. Y. You, Phosphorescence bioimaging using cyclometalated Ir(III) complexes, Curr. Opin. Chem. Biol., 2013, 17, 699–707 CrossRef CAS PubMed.
  1055. T. K. Hyster, L. Knorr, T. R. Ward and T. Rovis, Biotinylated Rh(III) complexes in engineered streptavidin for accelerated asymmetric C–H activation, Science, 2012, 338, 500–503 CrossRef CAS PubMed.
  1056. S. V. Wegner, H. Boyaci, H. Chen, M. P. Jensen and C. He, Engineering a uranyl-specific binding protein from NikR, Angew. Chem., Int. Ed., 2009, 48, 2339–2341 CrossRef CAS PubMed.
  1057. L. Zhou, M. Bosscher, C. Zhang, S. özçubukçu, L. Zhang, W. Zhang, C. J. Li, J. Liu, M. P. Jensen, L. Lai and C. He, A protein engineered to bind uranyl selectively and with femtomolar affinity, Nat. Chem., 2014, 6, 236–241 CrossRef CAS PubMed.
  1058. P. Turner, G. Mamo and E. N. Karlsson, Potential and utilization of thermophiles and thermostable enzymes in biorefining, Microb. Cell Fact., 2007, 6, 9 CrossRef PubMed.
  1059. P. S. Low and G. N. Somero, Temperature adaptation of enzymes: a proposed molecular basis for the different catalytic efficiencies of enzymes from ectotherms and endotherms, Comp. Biochem. Physiol., Part B: Biochem. Mol. Biol., 1974, 49, 307–312 CrossRef CAS.
  1060. P. L. Wintrode and F. H. Arnold, Temperature adaptation of enzymes: lessons from laboratory evolution, Adv. Protein Chem., 2000, 55, 161–225 CrossRef CAS.
  1061. R. D. Socha and N. Tokuriki, Modulating protein stability: directed evolution strategies for improved protein function, FEBS J., 2013, 280, 5582–5595 CrossRef CAS PubMed.
  1062. Y. Gumulya and M. T. Reetz, Enhancing the thermal robustness of an enzyme by directed evolution: least favorable starting points and inferior mutants can map superior evolutionary pathways, ChemBioChem, 2011, 12, 2502–2510 CrossRef CAS PubMed.
  1063. R. L. Chang, K. Andrews, D. Kim, Z. Li, A. Godzik and B. Ø. Palsson, Structural systems biology evaluation of metabolic thermotolerance in Escherichia coli, Science, 2013, 340, 1220–1223 CrossRef CAS PubMed.
  1064. M. Lehmann, D. Kostrewa, M. Wyss, R. Brugger, A. D'Arcy, L. Pasamontes and A. van Loon, From DNA sequence to improved functionality: using protein sequence comparisons to rapidly design a thermostable consensus phytase, Protein Eng., 2000, 13, 49–57 CrossRef CAS PubMed.
  1065. M. Lehmann, L. Pasamontes, S. F. Lassen and M. Wyss, The consensus concept for thermostability engineering of proteins, Biochim. Biophys. Acta, 2000, 1543, 408–415 CrossRef CAS.
  1066. M. Lehmann and M. Wyss, Engineering proteins for thermostability: the use of sequence alignments versus rational design and directed evolution, Curr. Opin. Biotechnol., 2001, 12, 371–375 CrossRef CAS.
  1067. M. Lehmann, C. Loch, A. Middendorf, D. Studer, S. F. Lassen, L. Pasamontes, A. P. G. M. van Loon and M. Wyss, The consensus concept for thermostability engineering of proteins: further proof of concept, Protein Eng., 2002, 15, 403–411 CrossRef CAS PubMed.
  1068. R. G. Coleman and K. A. Sharp, Shape and evolution of thermostable protein structure, Proteins, 2010, 78, 420–433 CrossRef CAS PubMed.
  1069. C. A. Hokanson, G. Cappuccilli, T. Odineca, M. Bozic, C. A. Behnke, M. Mendez, W. J. Coleman and R. Crea, Engineering highly thermostable xylanase variants using an enhanced combinatorial library method, Protein Eng., Des. Sel., 2011, 24, 597–605 CrossRef CAS PubMed.
  1070. J. P. Aucamp, A. M. Cosme, G. J. Lye and P. A. Dalby, High-throughput measurement of protein stability in microtiter plates, Biotechnol. Bioeng., 2005, 89, 599–607 CrossRef CAS PubMed.
  1071. J. P. Aucamp, R. J. Martinez-Torres, E. G. Hibbert and P. A. Dalby, A microplate-based evaluation of complex denaturation pathways: structural stability of Escherichia coli transketolase, Biotechnol. Bioeng., 2008, 99, 1303–1310 CrossRef CAS PubMed.
  1072. T. Schwab and R. Sterner, Stabilization of a metabolic enzyme by library selection in Thermus thermophilus, ChemBioChem, 2011, 12, 1581–1588 CrossRef CAS PubMed.
  1073. C. Pfleger, S. Radestock, E. Schmidt and H. Gohlke, Global and local indices for characterizing biomolecular flexibility and rigidity, J. Comput. Chem., Jpn, 2013, 34, 220–233 CrossRef CAS PubMed.
  1074. P. L. Wintrode, D. Zhang, N. Vaidehi, F. H. Arnold and W. A. Goddard 3rd, Protein dynamics in a family of laboratory evolved thermophilic enzymes, J. Mol. Biol., 2003, 327, 745–757 CrossRef CAS.
  1075. T. J. Kamerzell and C. R. Middaugh, The complex inter-relationships between protein flexibility and stability, J. Pharm. Sci., 2008, 97, 3494–3517 CrossRef CAS PubMed.
  1076. E. Bae, R. M. Bannen and G. N. Phillips, Bioinformatic method for protein thermal stabilization by structural entropy optimization, Proc. Natl. Acad. Sci. U. S. A., 2008, 105, 9594–9597 CrossRef CAS PubMed.
  1077. F. X. Schmid, Lessons about protein stability from in vitro selections, ChemBioChem, 2011, 12, 1501–1507 CrossRef CAS PubMed.
  1078. E. Vázquez-Figueroa, J. Chaparro-Riggers and A. S. Bommarius, Development of a thermostable glucose dehydrogenase by a structure-guided consensus concept, ChemBioChem, 2007, 8, 2295–2301 CrossRef PubMed.
  1079. K. M. Polizzi, J. F. Chaparro-Riggers, E. Vazquez-Figueroa and A. S. Bommarius, Structure-guided consensus approach to create a more thermostable penicillin G acylase, Biotechnol. J., 2006, 1, 531–536 CrossRef CAS PubMed.
  1080. H. J. Wijma, R. J. Floor and D. B. Janssen, Structure- and sequence-analysis inspired engineering of proteins for enhanced thermostability, Curr. Opin. Struct. Biol., 2013, 23, 588–594 CrossRef CAS PubMed.
  1081. C. Vieille, D. S. Burdette and J. G. Zeikus, Thermozymes, Biotechnol. Annu. Rev., 1996, 2, 1–83 CAS.
  1082. J. G. Zeikus, C. Vieille and A. Savchenko, Thermozymes: biotechnology and structure-function relationships, Extremophiles, 1998, 2, 179–183 CrossRef CAS.
  1083. R. Maheshwari, G. Bharadwaj and M. K. Bhat, Thermophilic fungi: their physiology and enzymes, Microbiol. Mol. Biol. Rev., 2000, 64, 461–488 CrossRef CAS.
  1084. D. Sriprapundh, C. Vieille and J. G. Zeikus, Molecular determinants of xylose isomerase thermal stability and activity: analysis of thermozymes by site-directed mutagenesis, Protein Eng., 2000, 13, 259–265 CrossRef CAS PubMed.
  1085. C. Vieille and G. J. Zeikus, Hyperthermophilic enzymes: sources, uses, and molecular mechanisms for thermostability, Microbiol. Mol. Biol. Rev., 2001, 65, 1–43 CrossRef CAS PubMed.
  1086. M. E. Bruins, A. E. Janssen and R. M. Boom, Thermozymes and their applications: a review of recent literature and patents, Appl. Biochem. Biotechnol., 2001, 90, 155–186 CrossRef CAS.
  1087. W. F. Li, X. X. Zhou and P. Lu, Structural features of thermozymes, Biotechnol. Adv., 2005, 23, 271–281 CrossRef CAS PubMed.
  1088. L. C. Wu, J. X. Lee, H. D. Huang, B. J. Liu and J. T. Horng, An expert system to predict protein thermostability using decision trees, Expert Syst. Appl., 2009, 36, 668–674 CrossRef PubMed.
  1089. I. N. Berezovsky, The diversity of physical forces and mechanisms in intermolecular interactions, Phys. Biol., 2011, 8, 035002 CrossRef PubMed.
  1090. T. Imanaka, Molecular bases of thermophily in hyperthermophiles, Proc. Jpn. Acad., Ser. B, 2011, 87, 587–602 CrossRef CAS.
  1091. L. D. Unsworth, J. van der Oost and S. Koutsopoulos, Hyperthermophilic enzymes—stability, activity and implementation strategies for high temperature applications, FEBS J., 2007, 274, 4044–4056 CrossRef CAS PubMed.
  1092. H. Hashimoto, T. Inoue, M. Nishioka, S. Fujiwara, M. Takagi, T. Imanaka and Y. Kai, Hyperthermostable protein structure maintained by intra and inter-helix ion-pairs in archaeal O6-methylguanine-DNA methyltransferase, J. Mol. Biol., 1999, 292, 707–716 CrossRef CAS PubMed.
  1093. I. Matsui and K. Harata, Implication for buried polar contacts and ion pairs in hyperthermostable enzymes, FEBS J., 2007, 274, 4012–4022 CrossRef CAS PubMed.
  1094. C. H. Chan, H. K. Liang, N. W. Hsiao, M. T. Ko, P. C. Lyu and J. K. Hwang, Relationship between local structural entropy and protein thermostability, Proteins, 2004, 57, 684–691 CrossRef CAS PubMed.
  1095. R. B. Greaves and J. Warwicker, Stability and solubility of proteins from extremophiles, Biochem. Biophys. Res. Commun., 2009, 380, 581–585 CrossRef CAS PubMed.
  1096. P. C. Rathi, S. Radestock and H. Gohlke, Thermostabilizing mutations preferentially occur at structural weak spots with a high mutation ratio, J. Biotechnol., 2012, 159, 135–144 CrossRef CAS PubMed.
  1097. S. Radestock and H. Gohlke, Protein rigidity and thermophilic adaptation, Proteins, 2011, 79, 1089–1108 CrossRef CAS PubMed.
  1098. C. P. Lin, S. W. Huang, Y. L. Lai, S. C. Yen, C. H. Shih, C. H. Lu, C. C. Huang and J. K. Hwang, Deriving protein dynamical properties from weighted protein contact number, Proteins: Struct., Funct., Bioinf., 2008, 72, 929–935 CrossRef CAS PubMed.
  1099. J. K. Blum, M. D. Ricketts and A. S. Bommarius, Improved thermostability of AEH by combining B-FIT analysis and structure-guided consensus method, J. Biotechnol., 2012, 160, 214–221 CrossRef CAS PubMed.
  1100. J. R. Engen, Analysis of protein conformation and dynamics by hydrogen/deuterium exchange MS, Anal. Chem., 2009, 81, 7870–7875 CrossRef CAS PubMed.
  1101. I. A. Kaltashov, C. E. Bobst and R. R. Abzalimov, H/D exchange and mass spectrometry in the studies of protein conformation and dynamics: is there a need for a top-down approach?, Anal. Chem., 2009, 81, 7892–7899 CrossRef CAS PubMed.
  1102. S. T. Esswein, H. V. Florance, L. Baillie, J. Lippens and P. E. Barran, A comparison of mass spectrometry based hydrogen deuterium exchange methods for probing the cyclophilin A cyclosporin complex, J. Chromatogr. A, 2010, 1217, 6709–6717 CrossRef CAS PubMed.
  1103. L. Konermann, J. Pan and Y. H. Liu, Hydrogen exchange mass spectrometry for studying protein structure and dynamics, Chem. Soc. Rev., 2011, 40, 1224–1234 RSC.
  1104. E. Jurneczko, F. Cruickshank, M. Porrini, P. Nikolova, I. D. Campuzano, M. Morris and P. E. Barran, Intrinsic disorder in proteins: a challenge for (un)structural biology met by ion mobility-mass spectrometry, Biochem. Soc. Trans., 2012, 40, 1021–1026 CrossRef CAS PubMed.
  1105. S. Nakazawa, J. Ahn, N. Hashii, K. Hirose and N. Kawasaki, Analysis of the local dynamics of human insulin and a rapid-acting insulin analog by hydrogen/deuterium exchange mass spectrometry, Biochim. Biophys. Acta, 2013, 1834, 1210–1214 CrossRef CAS PubMed.
  1106. D. Resetca and D. J. Wilson, Characterizing rapid, activity-linked conformational transitions in proteins via sub-second hydrogen deuterium exchange mass spectrometry, FEBS J., 2013, 280, 5616–5625 CrossRef CAS PubMed.
  1107. D. Goswami, C. Callaway, B. D. Pascal, R. Kumar, D. P. Edwards and P. R. Griffin, Influence of domain interactions on conformational mobility of the progesterone receptor detected by hydrogen/deuterium exchange mass spectrometry, Structure, 2014, 22, 961–973 CrossRef CAS PubMed.
  1108. D. P. Marciano, V. Dharmarajan and P. R. Griffin, HDX-MS guided drug discovery: small molecules and biopharmaceuticals, Curr. Opin. Struct. Biol., 2014, 28C, 105–111 CrossRef PubMed.
  1109. C. Pfleger, P. C. Rathi, D. L. Klein, S. Radestock and H. Gohlke, Constraint Network Analysis (CNA): a Python software package for efficiently linking biomacromolecular structure, flexibility, (thermo-)stability, and function, J. Chem. Inf. Model., 2013, 53, 1007–1015 CrossRef CAS PubMed.
  1110. B. C. Buer, B. J. Levin and E. N. G. Marsh, Influence of Fluorination on the Thermodynamics of Protein Folding, J. Am. Chem. Soc., 2012, 134, 13027–13034 CrossRef CAS PubMed.
  1111. B. C. Buer and E. N. G. Marsh, Fluorine: a new element in protein design, Protein Sci., 2012, 21, 453–462 CrossRef CAS PubMed.
  1112. B. C. Buer, J. L. Meagher, J. A. Stuckey and E. N. G. Marsh, Structural basis for the enhanced stability of highly fluorinated proteins, Proc. Natl. Acad. Sci. U. S. A., 2012, 109, 4810–4815 CrossRef CAS PubMed.
  1113. K. H. Oh, S. H. Nam and H. S. Kim, Improvement of oxidative and thermostability of N-carbamyl-d-amino Acid amidohydrolase by directed evolution, Protein Eng., 2002, 15, 689–695 CrossRef CAS PubMed.
  1114. K. H. Oh, S. H. Nam and H. S. Kim, Directed evolution of N-carbamyl-D-amino acid amidohydrolase for simultaneous improvement of oxidative and thermal stability, Biotechnol. Prog., 2002, 18, 413–417 CrossRef CAS PubMed.
  1115. E. Vázquez-Figueroa, V. Yeh, J. M. Broering, J. F. Chaparro-Riggers and A. S. Bommarius, Thermostable variants constructed via the structure-guided consensus method also show increased stability in salts solutions and homogeneous aqueous-organic media, Protein Eng., Des. Sel., 2008, 21, 673–680 CrossRef PubMed.
  1116. P. D. Dobson and D. B. Kell, Carrier-mediated cellular uptake of pharmaceutical drugs: an exception or the rule?, Nat. Rev. Drug Discovery, 2008, 7, 205–220 CrossRef CAS PubMed.
  1117. P. Dobson, K. Lanthaler, S. G. Oliver and D. B. Kell, Implications of the dominant role of cellular transporters in drug uptake, Curr. Top. Med. Chem., 2009, 9, 163–184 CrossRef CAS.
  1118. D. B. Kell, P. D. Dobson and S. G. Oliver, Pharmaceutical drug transport: the issues and the implications that it is essentially carrier-mediated only., Drug Discovery Today, 2011, 16, 704–714 CrossRef CAS PubMed.
  1119. D. B. Kell and R. Goodacre, Metabolomics and systems pharmacology: why and how to model the human metabolic network for drug discovery, Drug Discovery Today, 2014, 19, 171–182 CrossRef CAS PubMed.
  1120. G. J. Salter and D. B. Kell, Solvent selection for whole cell biotransformations in organic media., CRC Crit. Rev. Biotechnol., 1995, 15, 139–177 CrossRef CAS PubMed.
  1121. C. R. Wescott and A. M. Klibanov, Predicting the solvent dependence of enzymatic substrate specificity using semiempirical thermodynamic calculations, J. Am. Chem. Soc., 1993, 115, 10362–10363 CrossRef CAS.
  1122. C. R. Wescott and A. M. Klibanov, The solvent dependence of enzyme specificity, Biochim. Biophys. Acta, 1994, 1206, 1–9 CrossRef CAS.
  1123. G. Carrea, G. Ottolina and S. Riva, Role of solvents in the control of enzyme selectivity in organic media, Trends Biotechnol., 1995, 13, 63–70 CrossRef CAS.
  1124. Y. Sardessai and S. Bhosle, Tolerance of bacteria to organic solvents, Res. Microbiol., 2002, 153, 263–268 CrossRef CAS.
  1125. Y. N. Sardessai and S. Bhosle, Industrial Potential of Organic Solvent Tolerant Bacteria, Biotechnol. Prog., 2004, 20, 655–660 CrossRef CAS PubMed.
  1126. C. Liu, G. Yang, L. Wu, G. Tian, Z. Zhang and Y. Feng, Switch of substrate specificity of hyperthermophilic acylaminoacyl peptidase by combination of protein and solvent engineering, Protein Cell, 2011, 2, 497–506 CrossRef CAS PubMed.
  1127. P. J. Halling, Solvent selection for biocatalysis in mainly organic systems: predictions of effects on equilibrium position, Biotechnol. Bioeng., 1990, 35, 691–701 CrossRef CAS PubMed.
  1128. K. Xu, K. Griebenow and A. M. Klibanov, Correlation between catalytic activity and secondary structure of subtilisin dissolved in organic solvents, Biotechnol. Bioeng., 1997, 56, 485–491 CrossRef CAS.
  1129. M. N. Gupta and I. Roy, Enzymes in organic media. Forms, functions and applications, Eur. J. Biochem., 2004, 271, 2575–2583 CrossRef CAS PubMed.
  1130. E. M. Nordwald and J. L. Kaar, Stabilization of Enzymes in Ionic Liquids Via Modification of Enzyme Charge, Biotechnol. Bioeng., 2013, 110, 2352–2360 CrossRef CAS PubMed.
  1131. E. M. Nordwald and J. L. Kaar, Mediating Electrostatic Binding of 1-Butyl-3-methylimidazolium Chloride to Enzyme Surfaces Improves Conformational Stability, J. Phys. Chem. B, 2013, 117, 8977–8986 CrossRef CAS PubMed.
  1132. Z. Maugeri, W. Leitner and P. D. de Maria, Practical separation of alcohol-ester mixtures using Deep-Eutectic-Solvents, Tetrahedron Lett., 2012, 53, 6968–6971 CrossRef CAS PubMed.
  1133. Z. Maugeri and P. D. de Maria, Novel choline-chloride-based deep-eutectic-solvents with renewable hydrogen bond donors: levulinic acid and sugar-based polyols, RSC Adv., 2012, 2, 421–425 RSC.
  1134. P. Domínguez de María and Z. Maugeri, Ionic liquids in biotransformations: from proof-of-concept to emerging deep-eutectic-solvents, Curr. Opin. Chem. Biol., 2011, 15, 220–225 CrossRef PubMed.
  1135. Q. H. Zhang, K. D. Vigier, S. Royer and F. Jérôme, Deep eutectic solvents: syntheses, properties and applications, Chem. Soc. Rev., 2012, 41, 7108–7146 RSC.
  1136. A. Cadeddu, E. K. Wylie, J. Jurczak, M. Wampler-Doty and B. A. Grzybowski, Organic chemistry as a language and the implications of chemical linguistics for structural and retrosynthetic analyses, Angew. Chem., Int. Ed., 2014, 53, 8108–8112 CrossRef CAS PubMed.
  1137. P. Carbonell, A. G. Planson, D. Fichera and J. L. Faulon, A retrosynthetic biology approach to metabolic pathway design for therapeutic production, BMC Syst. Biol., 2011, 5, 122 CrossRef PubMed.
  1138. P. Carbonell, A. G. Planson and J. L. Faulon, Retrosynthetic design of heterologous pathways, Methods Mol. Biol., 2013, 985, 149–173 CAS.
  1139. Q. Huang, L. L. Li and S. Y. Yang, RASA: a rapid retrosynthesis-based scoring method for the assessment of synthetic accessibility of drug-like molecules, J. Chem. Inf. Model., 2011, 51, 2768–2777 CrossRef CAS PubMed.
  1140. J. Law, Z. Zsoldos, A. Simon, D. Reid, Y. Liu, S. Y. Khew, A. P. Johnson, S. Major, R. A. Wade and H. Y. Ando, Route Designer: a retrosynthetic analysis tool utilizing automated retrosynthetic rule generation, J. Chem. Inf. Model., 2009, 49, 593–602 CrossRef CAS PubMed.
  1141. X. Q. Lewell, D. B. Judd, S. P. Watson and M. M. Hann, RECAP–retrosynthetic combinatorial analysis procedure: a powerful new technique for identifying privileged molecular fragments with useful applications in combinatorial chemistry, J. Chem. Inf. Comput. Sci., 1998, 38, 511–522 CrossRef CAS.
  1142. J. Gonzalez-Lergier, L. J. Broadbelt and V. Hatzimanikatis, Theoretical considerations and computational analysis of the complexity in polyketide synthesis pathways, J. Am. Chem. Soc., 2005, 127, 9930–9938 CrossRef CAS PubMed.
  1143. V. Hatzimanikatis, C. Li, J. A. Ionita, C. S. Henry, M. D. Jankowski and L. J. Broadbelt, Exploring the diversity of complex metabolic networks, Bioinformatics, 2005, 21, 1603–1609 CrossRef CAS PubMed.
  1144. C. S. Henry, L. J. Broadbelt and V. Hatzimanikatis, Discovery and analysis of novel metabolic pathways for the biosynthesis of industrial chemicals: 3-hydroxypropanoate, Biotechnol. Bioeng., 2010, 106, 462–473 CAS.
  1145. K. C. Soh and V. Hatzimanikatis, DREAMS of metabolism, Trends Biotechnol., 2010, 28, 501–508 CrossRef CAS PubMed.
  1146. A. G. Planson, P. Carbonell, I. Grigoras and J. L. Faulon, A retrosynthetic biology approach to therapeutics: from conception to delivery, Curr. Opin. Biotechnol., 2012, 23, 948–956 CrossRef CAS PubMed.
  1147. N. J. Turner and E. O'Reilly, Biocatalytic retrosynthesis, Nat. Chem. Biol., 2013, 9, 285–288 CrossRef CAS PubMed.
  1148. M. A. Campodonico, B. A. Andrews, J. A. Asenjo, B. O. Palsson and A. M. Feist, Generation of an atlas for commodity chemical production in Escherichia coli and a novel pathway prediction algorithm, GEM-Path, Metab Eng, 2014, 25, 140–158 CrossRef CAS PubMed.
  1149. K. J. Bishop, R. Klajn and B. A. Grzybowski, The core and most useful molecules in organic chemistry, Angew. Chem., Int. Ed., 2006, 45, 5348–5354 CrossRef CAS PubMed.
  1150. M. Fialkowski, K. J. Bishop, V. A. Chubukov, C. J. Campbell and B. A. Grzybowski, Architecture and evolution of organic chemistry, Angew. Chem., Int. Ed., 2005, 44, 7263–7269 CrossRef CAS PubMed.
  1151. D. Ghislieri, A. P. Green, M. Pontini, S. C. Willies, I. Rowles, A. Frank, G. Grogan and N. J. Turner, Engineering an enantioselective amine oxidase for the synthesis of pharmaceutical building blocks and alkaloid natural products, J. Am. Chem. Soc., 2013, 135, 10863–10869 CrossRef CAS PubMed.
  1152. G. M. Whited, F. J. Feher, D. A. Benko, M. A. Cervin, G. K. Chotani, J. C. McAuliffe, R. J. LaDuca, E. A. Ben-Shoshan and K. J. Sanford, Development of a gas-phase bioprocess for isoprene-monomer production using metabolic pathway engineering, Ind. Biotechnol., 2010, 6, 152–163 CrossRef CAS.
  1153. G. DeSantis, K. Wong, B. Farwell, K. Chatman, Z. Zhu, G. Tomlinson, H. Huang, X. Tan, L. Bibbs, P. Chen, K. Kretz and M. J. Burk, Creation of a productive, highly enantioselective nitrilase through gene site saturation mutagenesis (GSSM), J. Am. Chem. Soc., 2003, 125, 11476–11477 CrossRef CAS PubMed.
  1154. A. S. Bommarius, J. K. Blum and M. J. Abrahamson, Status of protein engineering for biocatalysts: how to design an industrially useful biocatalyst, Curr. Opin. Chem. Biol., 2011, 15, 194–200 CrossRef CAS PubMed.
  1155. K. Faber, Biotransformations in organic chemistry. A textbook., Springer, Berlin, 2011 Search PubMed.
  1156. R. N. Patel, Biocatalysis: Synthesis of Key Intermediates for Development of Pharmaceuticals, ACS Catal., 2011, 1, 1056–1074 CrossRef CAS.
  1157. M. Schrewe, M. K. Julsing, B. Buhler and A. Schmid, Whole-cell biocatalysis for selective and productive C-O functional group introduction and modification, Chem. Soc. Rev., 2013, 42, 6346–6377 RSC.
  1158. R. C. Simon, F. G. Mutti and W. Kroutil, Biocatalytic synthesis of enantiopure building blocks for pharmaceuticals, Drug Discovery Today: Technol., 2013, 10, e37–44 CrossRef PubMed.
  1159. M. Wang, T. Si and H. Zhao, Biocatalyst development by directed evolution, Bioresour. Technol., 2012, 115, 117–125 CrossRef CAS PubMed.
  1160. N. Bhan, P. Xu and M. A. G. Koffas, Pathway and protein engineering approaches to produce novel and commodity small molecules, Curr. Opin. Biotechnol., 2013, 24, 1137–1143 CrossRef CAS PubMed.
  1161. G. W. Huisman and S. J. Collier, On the development of new biocatalytic processes for practical pharmaceutical synthesis, Curr. Opin. Chem. Biol., 2013, 17, 284–292 CrossRef CAS PubMed.
  1162. B. M. Nestl, S. C. Hammer, B. A. Nebel and B. Hauer, New generation of biocatalysts for organic synthesis, Angew. Chem., Int. Ed., 2014, 53, 3070–3095 CrossRef CAS PubMed.
  1163. A. Bolt, A. Berry and A. Nelson, Directed evolution of aldolases for exploitation in synthetic organic chemistry, Arch. Biochem. Biophys., 2008, 474, 318–330 CrossRef CAS PubMed.
  1164. C. L. Windle, M. Muller, A. Nelson and A. Berry, Engineering aldolases as biocatalysts, Curr. Opin. Chem. Biol., 2014, 19, 25–33 CrossRef CAS PubMed.
  1165. K. Okrasa, C. Levy, M. Wilding, M. Goodall, N. Baudendistel, B. Hauer, D. Leys and J. Micklefield, Structure-guided directed evolution of alkenyl and arylmalonate decarboxylases, Angew. Chem., Int. Ed., 2009, 48, 7691–7694 CrossRef CAS PubMed.
  1166. M. J. Abrahamson, E. Vazquez-Figueroa, N. B. Woodall, J. C. Moore and A. S. Bommarius, Development of an Amine Dehydrogenase for Synthesis of Chiral Amines, Angew. Chem., Int. Ed., 2012, 51, 3969–3972 CrossRef CAS PubMed.
  1167. M. J. Abrahamson, J. W. Wong and A. S. Bommarius, The Evolution of an Amine Dehydrogenase Biocatalyst for the Asymmetric Production of Chiral Amines, Adv. Synth. Catal., 2013, 355, 1780–1786 CrossRef CAS.
  1168. J. H. Sattler, M. Fuchs, K. Tauber, F. G. Mutti, K. Faber, J. Pfeffer, T. Haas and W. Kroutil, Redox Self-Sufficient Biocatalyst Network for the Amination of Primary Alcohols, Angew. Chem., Int. Ed., 2012, 51, 9156–9159 CrossRef CAS PubMed.
  1169. H. Kohls, F. Steffen-Munsberg and M. Höhne, Recent achievements in developing the biocatalytic toolbox for chiral amine synthesis, Curr. Opin. Chem. Biol., 2014, 19, 180–192 CrossRef CAS PubMed.
  1170. K. Meister, S. Ebbinghaus, Y. Xu, J. G. Duman, A. Devries, M. Gruebele, D. M. Leitner and M. Havenith, Long-range protein-water dynamics in hyperactive insect antifreeze proteins, Proc. Natl. Acad. Sci. U. S. A., 2013, 110, 1617–1622 CrossRef CAS PubMed.
  1171. M. L. Matthews, W. C. Chang, A. P. Layne, L. A. Miles, C. Krebs and J. M. Bollinger, Jr., Direct nitration and azidation of aliphatic carbons by an iron-dependent halogenase, Nat. Chem. Biol., 2014, 10, 209–215 CrossRef CAS PubMed.
  1172. E. M. Brustad, C-H activation: New recipes for biocatalysis, Nat. Chem. Biol., 2014, 10, 170–171 CrossRef CAS PubMed.
  1173. Z. G. Zhang, L. P. Parra and M. T. Reetz, Protein engineering of stereoselective Baeyer-Villiger monooxygenases, Chemistry, 2012, 18, 10160–10172 CrossRef CAS PubMed.
  1174. I. Polyak, M. T. Reetz and W. Thiel, Quantum mechanical/molecular mechanical study on the enantioselectivity of the enzymatic Baeyer–Villiger reaction of 4-hydroxycyclohexanone, J. Phys. Chem. B, 2013, 117, 4993–5001 CrossRef CAS PubMed.
  1175. T. Wells, Jr. and A. J. Ragauskas, Biotechnological opportunities with the beta-ketoadipate pathway, Trends Biotechnol., 2012, 30, 627–637 CrossRef PubMed.
  1176. C. Schmidt-Dannert, D. Umeno and F. H. Arnold, Molecular breeding of carotenoid biosynthetic pathways, Nat. Biotechnol., 2000, 18, 750–753 CrossRef CAS PubMed.
  1177. A. Butler and M. Sandy, Mechanistic considerations of halogenating enzymes, Nature, 2009, 460, 848–854 CrossRef CAS PubMed.
  1178. H. Deng and D. O'Hagan, The fluorinase, the chlorinase and the duf-62 enzymes, Curr. Opin. Chem. Biol., 2008, 12, 582–592 CrossRef CAS PubMed.
  1179. G. W. Gribble, Biohalogenation, Prog. Chem. Org. Nat. Prod., 2010, 91, 349–365 CAS.
  1180. W. Runguphan, X. Qu and S. E. O'Connor, Integrating carbon-halogen bond formation into medicinal plant metabolism, Nature, 2010, 468, 461–464 CrossRef CAS PubMed.
  1181. R. Vázquez-Duhalt, M. Ayala and F. J. Márquez-Rocha, Biocatalytic chlorination of aromatic hydrocarbons by chloroperoxidase of Caldariomyces fumago, Phytochemistry, 2001, 58, 929–933 CrossRef.
  1182. R. De Mot, A. De Schrijver, G. Schoofs and A. H. Parret, The thiocarbamate-inducible Rhodococcus enzyme ThcF as a member of the family of alpha/beta hydrolases with haloperoxidative side activity, FEMS Microbiol. Lett., 2003, 224, 197–203 CrossRef CAS.
  1183. Z. Hasan, R. Renirie, R. Kerkman, H. J. Ruijssenaars, A. F. Hartog and R. Wever, Laboratory-evolved vanadium chloroperoxidase exhibits 100-fold higher halogenating activity at alkaline pH: catalytic effects from first and second coordination sphere mutations, J. Biol. Chem., 2006, 281, 9738–9744 CrossRef CAS PubMed.
  1184. M. Hofrichter and R. Ullrich, Heme-thiolate haloperoxidases: versatile biocatalysts with biotechnological and environmental significance, Appl. Microbiol. Biotechnol., 2006, 71, 276–288 CrossRef CAS PubMed.
  1185. J. M. Winter and B. S. Moore, Exploring the Chemistry and Biology of Vanadium-dependent Haloperoxidases, J. Biol. Chem., 2009, 284, 18577–18581 CrossRef CAS PubMed.
  1186. M. Hofrichter, R. Ullrich, M. J. Pecyna, C. Liers and T. Lundell, New and classic families of secreted fungal heme peroxidases, Appl. Microbiol. Biotechnol., 2010, 87, 871–897 CrossRef CAS PubMed.
  1187. P. Bernhardt, T. Okino, J. M. Winter, A. Miyanaga and B. S. Moore, A Stereoselective Vanadium-Dependent Chloroperoxidase in Bacterial Antibiotic Biosynthesis, J. Am. Chem. Soc., 2011, 133, 4268–4270 CrossRef CAS PubMed.
  1188. C. R. Otey, M. Landwehr, J. B. Endelman, K. Hiraga, J. D. Bloom and F. H. Arnold, Structure-guided recombination creates an artificial family of cytochromes P450, PLoS Biol., 2006, 4, e112 Search PubMed.
  1189. S. L. Kelly and D. E. Kelly, Microbial cytochromes P450: biodiversity and biotechnology. Where do cytochromes P450 come from, what do they do and what can they do for us?, Philos. Trans. R. Soc. London, Ser. B, 2013, 368, 20120476 CrossRef PubMed.
  1190. C. Gavira, R. Höfer, A. Lesot, F. Lambert, J. Zucca and D. Werck-Reichhart, Challenges and pitfalls of P450-dependent (+)-valencene bioconversion by Saccharomyces cerevisiae, Metab. Eng., 2013, 18, 25–35 CrossRef CAS PubMed.
  1191. D. C. Lamb and M. R. Waterman, Unusual properties of the cytochrome P450 superfamily, Philos. Trans. R. Soc. London, Ser. B, 2013, 368, 20120434 CrossRef PubMed.
  1192. J. M. Caswell, M. O'Neill, S. J. C. Taylor and T. S. Moody, Engineering and application of P450 monooxygenases in pharmaceutical and metabolite synthesis, Curr. Opin. Chem. Biol., 2013, 17, 271–275 CrossRef CAS PubMed.
  1193. G. A. Roberts, A. Celik, D. J. B. Hunter, T. W. B. Ost, J. H. White, S. K. Chapman, N. J. Turner and S. L. Flitsch, A self-sufficient cytochrome P450 with a primary structural organization that includes a flavin domain and a [2Fe-2S] redox center, J. Biol. Chem., 2003, 278, 48914–48920 CrossRef CAS PubMed.
  1194. D. J. B. Hunter, G. A. Roberts, T. W. B. Ost, J. H. White, S. Muller, N. J. Turner, S. L. Flitsch and S. K. Chapman, Analysis of the domain properties of the novel cytochrome P450RhF, FEBS Lett., 2005, 579, 2215–2220 CrossRef CAS PubMed.
  1195. E. O'Reilly, M. Corbett, S. Hussain, P. P. Kelly, D. Richardson, S. L. Flitsch and N. J. Turner, Substrate promiscuity of cytochrome P450 RhF, Catal. Sci. Technol., 2013, 3, 1490–1492 Search PubMed.
  1196. E. O'Reilly, S. J. Aitken, G. Grogan, P. P. Kelly, N. J. Turner and S. L. Flitsch, Regio- and stereoselective oxidation of unactivated C-H bonds with Rhodococcus rhodochrous, Beilstein J. Org. Chem., 2012, 8, 496–500 CrossRef PubMed.
  1197. A. Robin, V. Kohler, A. Jones, A. Ali, P. P. Kelly, E. O'Reilly, N. J. Turner and S. L. Flitsch, Chimeric self-sufficient P450cam-RhFRed biocatalysts with broad substrate scope, Beilstein J. Org. Chem., 2011, 7, 1494–1498 CrossRef CAS PubMed.
  1198. A. Robin, G. A. Roberts, J. Kisch, F. Sabbadin, G. Grogan, N. Bruce, N. J. Turner and S. L. Flitsch, Engineering and improvement of the efficiency of a chimeric [P450cam-RhFRed reductase domain] enzyme, Chem. Commun., 2009, 2478–2480 RSC.
  1199. R. Fasan, M. M. Chen, N. C. Crook and F. H. Arnold, Engineered alkane-hydroxylating cytochrome P450(BM3) exhibiting nativelike catalytic properties, Angew. Chem., Int. Ed., 2007, 46, 8414–8418 CrossRef CAS PubMed.
  1200. A. Trefzer, V. Jungmann, I. Molnar, A. Botejue, D. Buckel, G. Frey, D. S. Hill, M. Jorg, J. M. Ligon, D. Mason, D. Moore, J. P. Pachlatko, T. H. Richardson, P. Spangenberg, M. A. Wall, R. Zirkle and J. T. Stege, Biocatalytic conversion of avermectin to 4''-oxo-avermectin: improvement of cytochrome p450 monooxygenase specificity by directed evolution, Appl. Environ. Microbiol., 2007, 73, 4317–4325 CrossRef CAS PubMed.
  1201. J. C. Lewis, S. M. Mantovani, Y. Fu, C. D. Snow, R. S. Komor, C. H. Wong and F. H. Arnold, Combinatorial alanine substitution enables rapid optimization of cytochrome P450BM3 for selective hydroxylation of large substrates, ChemBioChem, 2010, 11, 2502–2505 CrossRef CAS PubMed.
  1202. V. B. Urlacher and M. Girhard, Cytochrome P450 monooxygenases: an update on perspectives for synthetic application, Trends Biotechnol., 2012, 30, 26–36 CrossRef CAS PubMed.
  1203. A. Seifert, M. Antonovici, B. Hauer and J. Pleiss, An efficient route to selective bio-oxidation catalysts: an iterative approach comprising modeling, diversification, and screening, based on CYP102A1, ChemBioChem, 2011, 12, 1346–1351 CrossRef CAS PubMed.
  1204. F. E. Zilly, J. P. Acevedo, W. Augustyniak, A. Deege, U. W. Hausig and M. T. Reetz, Tuning a p450 enzyme for methane oxidation, Angew. Chem., Int. Ed., 2011, 50, 2720–2724 CrossRef CAS PubMed.
  1205. S. T. Jung, R. Lauchli and F. H. Arnold, Cytochrome P450: taming a wild type enzyme, Curr. Opin. Biotechnol., 2011, 22, 809–817 CrossRef CAS PubMed.
  1206. G. D. Roiban and M. T. Reetz, Enzyme promiscuity: using a P450 enzyme as a carbene transfer catalyst, Angew. Chem., Int. Ed., 2013, 52, 5439–5440 CrossRef CAS PubMed.
  1207. D. Jiang, R. Tu, P. Bai and Q. Wang, Directed evolution of cytochrome P450 for sterol epoxidation, Biotechnol. Lett., 2013, 35, 1663–1668 CrossRef CAS PubMed.
  1208. G. D. Roiban, R. Agudo, A. Ilie, R. Lonsdale and M. T. Reetz, CH-activating oxidative hydroxylation of 1-tetralones and related compounds with high regio- and stereoselectivity, Chem. Commun., 2014, 50, 14310–14313 RSC.
  1209. H. J. Kim, M. W. Ruszczycky, S. H. Choi, Y. N. Liu and H. W. Liu, Enzyme-catalysed [4+2] cycloaddition is a key step in the biosynthesis of spinosyn A, Nature, 2011, 473, 109–112 CrossRef CAS PubMed.
  1210. C. A. Townsend, A “Diels-Alderase” at Last, ChemBioChem, 2011, 12, 2267–2269 CrossRef CAS PubMed.
  1211. S. Obeid, A. Schnur, C. Gloeckner, N. Blatter, W. Welte, K. Diederichs and A. Marx, Learning from directed evolution: Thermus aquaticus DNA polymerase mutants with translesion synthesis activity, ChemBioChem, 2011, 12, 1574–1580 CrossRef CAS PubMed.
  1212. D. Loakes, J. Gallego, V. B. Pinheiro, E. T. Kool and P. Holliger, Evolving a polymerase for hydrophobic base analogues, J. Am. Chem. Soc., 2009, 131, 14827–14837 CrossRef CAS PubMed.
  1213. S. Park, K. L. Morley, G. P. Horsman, M. Holmquist, K. Hult and R. J. Kazlauskas, Focusing mutations into the P. fluorescens esterase binding site increases enantioselectivity more effectively than distant mutations, Chem. Biol., 2005, 12, 45–54 CrossRef CAS PubMed.
  1214. Z. L. Fowler and M. A. G. Koffas, Biosynthesis and biotechnological production of flavanones: current state and perspectives, Appl. Microbiol. Biotechnol., 2009, 83, 799–808 CrossRef CAS PubMed.
  1215. Z. L. Fowler, K. Shah, J. C. Panepinto, A. Jacobs and M. A. G. Koffas, Development of non-natural flavanones as antimicrobial agents, PLoS One, 2011, 6, e25681 CAS.
  1216. Z. L. Fowler, W. W. Gikandi and M. A. G. Koffas, Increased malonyl coenzyme A biosynthesis by tuning the Escherichia coli metabolic network and its application to flavanone production, Appl. Environ. Microbiol., 2009, 75, 5831–5839 CrossRef CAS PubMed.
  1217. E. Leonard, Y. Yan, Z. L. Fowler, Z. Li, C. G. Lim, K. H. Lim and M. A. G. Koffas, Strain improvement of recombinant Escherichia coli for efficient production of plant flavonoids, Mol. Pharmaceutics, 2008, 5, 257–265 CrossRef CAS PubMed.
  1218. P. Xu, S. Ranganathan, Z. L. Fowler, C. D. Maranas and M. A. Koffas, Genome-scale metabolic network modeling results in minimal interventions that cooperatively force carbon flux towards malonyl-CoA, Metab. Eng., 2011, 13, 578–587 CrossRef CAS PubMed.
  1219. M. Mora-Pale, S. P. Sanchez-Rodriguez, R. J. Linhardt, J. S. Dordick and M. A. G. Koffas, Metabolic engineering and in vitro biosynthesis of phytochemicals and non-natural analogues, Plant Sci., 2013, 210, 10–24 CrossRef CAS PubMed.
  1220. C. X. Wu, R. Liu, M. Gao, G. Zhao, S. Wu, C. F. Wu and G. H. Du, Pinocembrin protects brain against ischemia/reperfusion injury by attenuating endoplasmic reticulum stress induced apoptosis, Neurosci. Lett., 2013, 546, 57–62 CrossRef CAS PubMed.
  1221. J. Wu, G. Du, J. Zhou and J. Chen, Metabolic engineering of Escherichia coli for (2S)-pinocembrin production from glucose by a modular metabolic strategy, Metab. Eng., 2013, 16, 48–55 CrossRef PubMed.
  1222. H. Deng, S. M. Cross, R. P. McGlinchey, J. T. Hamilton and D. O'Hagan, In vitro reconstituted biotransformation of 4-fluorothreonine from fluoride ion: application of the fluorinase, Chem. Biol., 2008, 15, 1268–1276 CrossRef CAS PubMed.
  1223. W. Liu, X. Huang, M. J. Cheng, R. J. Nielsen, W. A. Goddard, 3rd and J. T. Groves, Oxidative aliphatic C-H fluorination with fluoride ion catalyzed by a manganese porphyrin, Science, 2012, 337, 1322–1325 CrossRef CAS PubMed.
  1224. R. M. Lennen, M. G. Politz, M. A. Kruziki and B. F. Pfleger, Identification of transport proteins involved in free fatty acid efflux in Escherichia coli, J. Bacteriol., 2013, 195, 135–144 CrossRef CAS PubMed.
  1225. R. M. Lennen and B. F. Pfleger, Engineering Escherichia coli to synthesize free fatty acids, Trends Biotechnol., 2012, 30, 659–667 CrossRef CAS PubMed.
  1226. J. T. Youngquist, R. M. Lennen, D. R. Ranatunga, W. H. Bothfeld, W. D. Marner 2nd and B. F. Pfleger, Kinetic modeling of free fatty acid production in Escherichia coli based on continuous cultivation of a plasmid free strain, Biotechnol. Bioeng., 2012, 109, 1518–1527 CrossRef CAS PubMed.
  1227. L. A. Castle, D. L. Siehl, R. Gorton, P. A. Patten, Y. H. Chen, S. Bertain, H. J. Cho, N. Duck, J. Wong, D. Liu and M. W. Lassner, Discovery and directed evolution of a glyphosate tolerance gene, Science, 2004, 304, 1151–1154 CrossRef CAS PubMed.
  1228. D. L. Siehl, L. A. Castle, R. Gorton, Y. H. Chen, S. Bertain, H. J. Cho, R. Keenan, D. Liu and M. W. Lassner, Evolution of a microbial acetyltransferase for modification of glyphosate: a novel tolerance strategy, Pest Manage. Sci., 2005, 61, 235–240 CrossRef CAS PubMed.
  1229. L. Pollegioni, E. Schonbrunn and D. Siehl, Molecular basis of glyphosate resistance-different approaches through protein engineering, FEBS J., 2011, 278, 2753–2766 CrossRef CAS PubMed.
  1230. M. Pedotti, E. Rosini, G. Molla, T. Moschetti, C. Savino, B. Vallone and L. Pollegioni, Glyphosate resistance by engineering the flavoenzyme glycine oxidase, J. Biol. Chem., 2009, 284, 36415–36423 CrossRef CAS PubMed.
  1231. T. Zhan, K. Zhang, Y. Chen, Y. Lin, G. Wu, L. Zhang, P. Yao, Z. Shao and Z. Liu, Improving glyphosate oxidation activity of glycine oxidase from Bacillus cereus by directed evolution, PLoS One, 2013, 8, e79175 CAS.
  1232. Z. Prokop, Y. Sato, J. Brezovsky, T. Mozga, R. Chaloupkova, T. Koudelakova, P. Jerabek, V. Stepankova, R. Natsume, J. G. van Leeuwen, D. B. Janssen, J. Florian, Y. Nagata, T. Senda and J. Damborsky, Enantioselectivity of haloalkane dehalogenases and its modulation by surface loop engineering, Angew. Chem., Int. Ed., 2010, 49, 6111–6115 CrossRef CAS PubMed.
  1233. T. Koudelakova, E. Chovancova, J. Brezovsky, M. Monincova, A. Fortova, J. Jarkovsky and J. Damborsky, Substrate specificity of haloalkane dehalogenases, Biochem. J., 2011, 435, 345–354 CrossRef CAS PubMed.
  1234. J. G. E. van Leeuwen, H. J. Wijma, R. J. Floor, J. M. van der Laan and D. B. Janssen, Directed evolution strategies for enantiocomplementary haloalkane dehalogenases: from chemical waste to enantiopure building blocks, ChemBioChem, 2012, 13, 137–148 CrossRef CAS PubMed.
  1235. R. J. Floor, H. J. Wijma, D. I. Colpa, A. Ramos-Silva, P. A. Jekel, W. Szymanski, B. L. Feringa, S. J. Marrink and D. B. Janssen, Computational library design for increasing haloalkane dehalogenase stability, ChemBioChem, 2014, 15, 1660–1672 CrossRef CAS PubMed.
  1236. W. S. Glenn, E. Nims and S. E. O'Connor, Reengineering a tryptophan halogenase to preferentially chlorinate a direct alkaloid precursor, J. Am. Chem. Soc., 2011, 133, 19346–19349 CrossRef CAS PubMed.
  1237. L. N. Herrera-Rodriguez, H. P. Meyer, K. T. Robins and F. Khan, Perspectives on biotechnological halogenation Part II: Prospecting for future biohalogenases, Chim. Oggi, 2011, 29, 47–49 CAS.
  1238. L. N. Herrera-Rodriguez, F. Khan, K. T. Robins and H. P. Meyer, Perspectives on biotechnological halogenation Part I: Halogenated products and enzymatic halogenation, Chim Oggi, 2011, 29, 31–33 CAS.
  1239. S. D. Wong, M. Srnec, M. L. Matthews, L. V. Liu, Y. Kwak, K. Park, C. B. Bell 3rd, E. E. Alp, J. Zhao, Y. Yoda, S. Kitao, M. Seto, C. Krebs, J. M. Bollinger, Jr. and E. I. Solomon, Elucidation of the Fe(IV)[double bond, length as m-dash]O intermediate in the catalytic cycle of the halogenase SyrB2, Nature, 2013, 499, 320–323 CrossRef CAS PubMed.
  1240. K. Bernath-Levin, J. Shainsky, L. Sigawi and A. Fishman, Directed evolution of nitrobenzene dioxygenase for the synthesis of the antioxidant hydroxytyrosol, Appl. Microbiol. Biotechnol., 2014, 98, 4975–4985 CrossRef CAS PubMed.
  1241. O. Khersonsky, D. Röthlisberger, O. Dym, S. Albeck, C. J. Jackson, D. Baker and D. S. Tawfik, Evolutionary optimization of computationally designed enzymes: Kemp eliminases of the KE07 series, J. Mol. Biol., 2010, 396, 1025–1042 CrossRef CAS PubMed.
  1242. O. Khersonsky, D. Röthlisberger, A. M. Wollacott, P. Murphy, O. Dym, S. Albeck, G. Kiss, K. N. Houk, D. Baker and D. S. Tawfik, Optimization of the in-silico-designed Kemp eliminase KE70 by computational design and directed evolution, J. Mol. Biol., 2010, 407, 391–412 CrossRef PubMed.
  1243. R. Blomberg, H. Kries, D. M. Pinkas, P. R. Mittl, M. G. Grutter, H. K. Privett, S. L. Mayo and D. Hilvert, Precision is essential for efficient catalysis in an evolved Kemp eliminase, Nature, 2013, 503, 418–421 CrossRef CAS PubMed.
  1244. A. Labas, E. Szabo, L. Mones and M. Fuxreiter, Optimization of reorganization energy drives evolution of the designed Kemp eliminase KE07, Biochim. Biophys. Acta, 2013, 1834, 908–917 CrossRef CAS PubMed.
  1245. O. Khersonsky, G. Kiss, D. Rothlisberger, O. Dym, S. Albeck, K. N. Houk, D. Baker and D. S. Tawfik, Bridging the gaps in design methodologies by evolutionary optimization of the stability and proficiency of designed Kemp eliminase KE59, Proc. Natl. Acad. Sci. U. S. A., 2012, 109, 10358–10363 CrossRef CAS PubMed.
  1246. M. P. Frushicheva, J. Cao and A. Warshel, Challenges and advances in validating enzyme design proposals: the case of Kemp eliminase catalysis, Biochemistry, 2011, 50, 3849–3858 CrossRef CAS PubMed.
  1247. M. P. Frushicheva, J. Cao, Z. T. Chu and A. Warshel, Exploring challenges in rational enzyme design by simulating the catalysis in artificial kemp eliminase, Proc. Natl. Acad. Sci. U. S. A., 2010, 107, 16869–16874 CrossRef CAS PubMed.
  1248. J. C. Moore, D. J. Pollard, B. Kosjek and P. N. Devine, Advances in the enzymatic reduction of ketones, Acc. Chem. Res., 2007, 40, 1412–1419 CrossRef CAS PubMed.
  1249. R. Agudo, G. D. Roiban and M. T. Reetz, Induced axial chirality in biocatalytic asymmetric ketone reduction, J. Am. Chem. Soc., 2013, 135, 1665–1668 CrossRef CAS PubMed.
  1250. S. Camarero, I. Pardo, A. I. Cañas, P. Molina, E. Record, A. T. Martínez, M. J. Martínez and M. Alcalde, Engineering platforms for directed evolution of Laccase from Pycnoporus cinnabarinus, Appl. Environ. Microbiol., 2012, 78, 1370–1384 CrossRef CAS PubMed.
  1251. J. R. Jeon and Y. S. Chang, Laccase-mediated oxidation of small organics: bifunctional roles for versatile applications, Trends Biotechnol., 2013, 31, 335–341 CrossRef CAS PubMed.
  1252. D. M. Mate, D. Gonzalez-Perez, M. Falk, R. Kittl, M. Pita, A. L. De Lacey, R. Ludwig, S. Shleev and M. Alcalde, Blood tolerant laccase by directed evolution, Chem. Biol., 2013, 20, 223–231 CrossRef CAS PubMed.
  1253. Y. Miao, E. M. Geertsema, P. G. Tepper, E. Zandvoort and G. J. Poelarends, Promiscuous catalysis of asymmetric Michael-type additions of linear aldehydes to beta-nitrostyrene by the proline-based enzyme 4-oxalocrotonate tautomerase, ChemBioChem, 2013, 14, 191–194 CrossRef CAS PubMed.
  1254. K. E. Atkin, R. Reiss, V. Koehler, K. R. Bailey, S. Hart, J. P. Turkenburg, N. J. Turner, A. M. Brzozowski and G. Grogan, The structure of monoamine oxidase from Aspergillus niger provides a molecular context for improvements in activity obtained by directed evolution, J. Mol. Biol., 2008, 384, 1218–1231 CrossRef CAS PubMed.
  1255. K. R. Bailey, A. J. Ellis, R. Reiss, T. J. Snape and N. J. Turner, A template-based mnemonic for monoamine oxidase (MAO-N) catalyzed reactions and its application to the chemo-enzymatic deracemisation of the alkaloid (+/-)-crispine A, Chem. Commun., 2007, 3640–3642 RSC.
  1256. I. Rowles, K. J. Malone, L. L. Etchells, S. C. Willies and N. J. Turner, Directed evolution of the enzyme monoamine oxidase (MAO-N): highly efficient chemo-enzymatic deracemisation of the alkaloid (+/-)-Crispine A, ChemCatChem, 2012, 4, 1259–1261 CrossRef CAS.
  1257. J. S. Anderson, J. Rittle and J. C. Peters, Catalytic conversion of nitrogen to ammonia by an iron model complex, Nature, 2013, 501, 84–87 CrossRef CAS PubMed.
  1258. A. Pingoud and W. Wende, Generation of novel nucleases with extended specificity by rational and combinatorial strategies, ChemBioChem, 2011, 12, 1495–1500 CrossRef CAS PubMed.
  1259. H. S. Toogood and N. S. Scrutton, Enzyme engineering toolbox - a 'catalyst' for change, Catal. Sci. Technol., 2013, 3, 2182–2194 CAS.
  1260. H. S. Toogood and N. S. Scrutton, New developments in 'ene'-reductase catalysed biological hydrogenations, Curr. Opin. Chem. Biol., 2014, 19, 107–115 CrossRef CAS PubMed.
  1261. Y. Ashani, R. D. Gupta, M. Goldsmith, I. Silman, J. L. Sussman, D. S. Tawfik and H. Leader, Stereo-specific synthesis of analogs of nerve agents and their utilization for selection and characterization of paraoxonase (PON1) catalytic scavengers, Chem.-Biol. Interact., 2010, 187, 362–369 CrossRef CAS PubMed.
  1262. U. Alcolombri, M. Elias and D. S. Tawfik, Directed evolution of sulfotransferases and paraoxonases by ancestral libraries, J. Mol. Biol., 2011, 411, 837–853 CrossRef CAS PubMed.
  1263. R. D. Gupta, M. Goldsmith, Y. Ashani, Y. Simo, G. Mullokandov, H. Bar, M. Ben-David, H. Leader, R. Margalit, I. Silman, J. L. Sussman and D. S. Tawfik, Directed evolution of hydrolases for prevention of G-type nerve agent intoxication, Nat. Chem. Biol., 2011, 7, 120–125 CrossRef CAS PubMed.
  1264. E. Garcia-Ruiz, D. Gonzalez-Perez, F. J. Ruiz-Dueñas, A. T. Martínez and M. Alcalde, Directed evolution of a temperature-peroxide- and alkaline pH-tolerant versatile peroxidase, Biochem. J., 2012, 441, 487–498 CrossRef CAS PubMed.
  1265. S. C. Patel and M. H. Hecht, Directed evolution of the peroxidase activity of a de novo-designed protein, Protein Eng., Des. Sel., 2012, 25, 445–452 CrossRef CAS PubMed.
  1266. L. F. Olguin, S. E. Askew, A. C. O'Donoghue and F. Hollfelder, Efficient catalytic promiscuity in an enzyme superfamily: an arylsulfatase shows a rate acceleration of 1013 for phosphate monoester hydrolysis, J. Am. Chem. Soc., 2008, 130, 16547–16555 CrossRef CAS PubMed.
  1267. A. C. Babtie, S. Bandyopadhyay, L. F. Olguin and F. Hollfelder, Efficient catalytic promiscuity for chemically distinct reactions, Angew. Chem., Int. Ed., 2009, 48, 3692–3694 CrossRef CAS PubMed.
  1268. B. van Loo, S. Jonas, A. C. Babtie, A. Benjdia, O. Berteau, M. Hyvonen and F. Hollfelder, An efficient, multiply promiscuous hydrolase in the alkaline phosphatase superfamily, Proc. Natl. Acad. Sci. U. S. A., 2010, 107, 2740–2745 CrossRef CAS PubMed.
  1269. L. Afriat-Jurnou, C. J. Jackson and D. S. Tawfik, Reconstructing a missing link in the evolution of a recently diverged phosphotriesterase by active-site loop remodeling, Biochemistry, 2012, 51, 6047–6055 CrossRef CAS PubMed.
  1270. M. F. Mohamed and F. Hollfelder, Efficient, crosswise catalytic promiscuity among enzymes that catalyze phosphoryl transfer, Biochim. Biophys. Acta, 2013, 1834, 417–424 CrossRef CAS PubMed.
  1271. H. Wiersma-Koch, F. Sunden and D. Herschlag, Site-directed mutagenesis maps interactions that enhance cognate and limit promiscuous catalysis by an alkaline phosphatase superfamily phosphodiesterase, Biochemistry, 2013, 52, 9167–9176 CrossRef CAS PubMed.
  1272. H. Fu and C. Khosla, Antibiotic activity of polyketide products derived from combinatorial biosynthesis: implications for directed evolution, Mol. Diversity, 1996, 1, 121–124 CrossRef CAS.
  1273. S. M. Ma, J. W. Li, J. W. Choi, H. Zhou, K. K. Lee, V. A. Moorthie, X. Xie, J. T. Kealey, N. A. Da Silva, J. C. Vederas and Y. Tang, Complete reconstitution of a highly reducing iterative polyketide synthase, Science, 2009, 326, 589–592 CrossRef CAS PubMed.
  1274. W. Zha, S. B. Rubin-Pitel and H. Zhao, Exploiting genetic diversity by directed evolution: molecular breeding of type III polyketide synthases improves productivity, Mol. BioSyst., 2008, 4, 246–248 RSC.
  1275. H. Y. Lee, C. J. Harvey, D. E. Cane and C. Khosla, Improved precursor-directed biosynthesis in E. colivia directed evolution, J. Antibiot., 2011, 64, 59–64 CrossRef CAS PubMed.
  1276. T. H. Yang, T. W. Kim, H. O. Kang, S. H. Lee, E. J. Lee, S. C. Lim, S. O. Oh, A. J. Song, S. J. Park and S. Y. Lee, Biosynthesis of polylactic acid and its copolymers using evolved propionate CoA transferase and PHA synthase, Biotechnol. Bioeng., 2010, 105, 150–160 CrossRef CAS PubMed.
  1277. F. Hollmann, I. W. C. E. Arends and K. Buehler, Biocatalytic Redox Reactions for Organic Synthesis: Nonconventional Regeneration Methods, ChemCatChem, 2010, 2, 762–782 CrossRef CAS.
  1278. F. Hollmann, I. W. C. E. Arends, K. Buehler, A. Schallmey and B. Buhler, Enzyme-mediated oxidations for the chemist, Green Chem., 2011, 13, 226–265 RSC.
  1279. F. Hollmann, I. W. C. E. Arends and D. Holtmann, Enzymatic reductions for the chemist, Green Chem., 2011, 13, 2285–2314 RSC.
  1280. F. Geu-Flores, N. H. Sherden, V. Courdavault, V. Burlat, W. S. Glenn, C. Wu, E. Nims, Y. Cui and S. E. O'Connor, An alternative route to cyclic terpenes by reductive cyclization in iridoid biosynthesis, Nature, 2012, 492, 138–142 CrossRef CAS PubMed.
  1281. M. Schöpfel, A. Tziridis, U. Arnold and M. T. Stubbs, Towards a restriction proteinase: construction of a self-activating enzyme, ChemBioChem, 2011, 12, 1523–1527 CrossRef PubMed.
  1282. R. Obexer, S. Studer, L. Giger, D. M. Pinkas, M. G. Grütter, D. Baker and D. Hilvert, Active Site Plasticity of a Computationally Designed RetroAldolase Enzyme, ChemCatChem, 2014, 6, 1043–1050 CrossRef CAS.
  1283. T. Wymore, B. Y. Chen, H. B. Nicholas, A. J. Ropelewski and C. L. Brooks, A Mechanism for Evolving Novel Plant Sesquiterpene Synthase Function, Mol. Inf., 2011, 30, 896–906 CrossRef CAS.
  1284. B. J. Baas, E. Zandvoort, E. M. Geertsema and G. J. Poelarends, Recent advances in the study of enzyme promiscuity in the tautomerase superfamily, ChemBioChem, 2013, 14, 917–926 CrossRef CAS PubMed.
  1285. J. E. Diaz, C. S. Lin, K. Kunishiro, B. K. Feld, S. K. Avrantinis, J. Bronson, J. Greaves, J. G. Saven and G. A. Weiss, Computational design and selections for an engineered, thermostable terpene synthase, Protein Sci., 2011, 20, 1597–1606 CrossRef CAS PubMed.
  1286. R. Lauchli, K. S. Rabe, K. Z. Kalbarczyk, A. Tata, T. Heel, R. Z. Kitto and F. H. Arnold, High-throughput screening for terpene-synthase-cyclization activity and directed evolution of a terpene synthase, Angew. Chem., Int. Ed., 2013, 52, 5571–5574 CrossRef CAS PubMed.
  1287. D. T. Major, Y. Freud and M. Weitman, Catalytic control in terpenoid cyclases: multiscale modeling of thermodynamic, kinetic, and dynamic effects, Curr. Opin. Chem. Biol., 2014, 21C, 25–33 CrossRef PubMed.
  1288. S. H. Chen, D. R. Hwang, G. H. Chen, N. S. Hsu, Y. T. Wu, T. L. Li and C. H. Wong, Engineering transaldolase in Pichia stipitis to improve bioethanol production, ACS Chem. Biol., 2012, 7, 481–486 CrossRef CAS PubMed.
  1289. A. K. Samland, M. Rale, G. A. Sprenger and W. D. Fessner, The transaldolase family: new synthetic opportunities from an ancient enzyme scaffold, ChemBioChem, 2011, 12, 1454–1474 CrossRef CAS PubMed.
  1290. E. G. Hibbert, T. Senussi, S. J. Costelloe, W. Lei, M. E. B. Smith, J. M. Ward, H. C. Hailes and P. A. Dalby, Directed evolution of transketolase activity on non-phosphorylated substrates, J. Biotechnol., 2007, 131, 425–432 CrossRef CAS PubMed.
  1291. E. G. Hibbert, T. Senussi, M. E. B. Smith, S. J. Costelloe, J. M. Ward, H. C. Hailes and P. A. Dalby, Directed evolution of transketolase substrate specificity towards an aliphatic aldehyde, J. Biotechnol., 2008, 134, 240–245 CrossRef CAS PubMed.
  1292. A. Cázares, J. L. Galman, L. G. Crago, M. E. B. Smith, J. Strafford, L. Ríos-Solís, G. J. Lye, P. A. Dalby and H. C. Hailes, Non-alpha-hydroxylated aldehydes with evolved transketolase enzymes, Org. Biomol. Chem., 2010, 8, 1301–1309 Search PubMed.
  1293. P. Payongsri, D. Steadman, J. Strafford, A. MacMurray, H. C. Hailes and P. A. Dalby, Rational substrate and enzyme engineering of transketolase for aromatics, Org. Biomol. Chem., 2012, 10, 9021–9029 CAS.
  1294. M. Minczuk, P. Kolasinska-Zwierz, M. P. Murphy and M. A. Papworth, Construction and testing of engineered zinc-finger proteins for sequence-specific modification of mtDNA, Nat. Protoc., 2010, 5, 342–356 CrossRef CAS PubMed.
  1295. M. Papworth, P. Kolasinska and M. Minczuk, Designer zinc-finger proteins and their applications, Gene, 2006, 366, 27–38 CrossRef CAS PubMed.
  1296. H. J. Wijma, S. J. Marrink and D. B. Janssen, Computationally efficient and accurate enantioselectivity modeling by clusters of molecular dynamics simulations, J. Chem. Inf. Model., 2014, 54, 2079–2092 CrossRef CAS PubMed.
  1297. H. J. Wijma and D. B. Janssen, Computational design gains momentum in enzyme catalysis engineering, FEBS J., 2013, 280, 2948–2960 CrossRef CAS PubMed.
  1298. A. Á. Rauscher, Z. Simon, G. J. Szöllösi, L. Gráf, I. Derényi and A. Málnási-Csizmadia, Temperature dependence of internal friction in enzyme reactions, FASEB J., 2011, 25, 2804–2813 CrossRef CAS PubMed.
  1299. A. Rauscher, I. Derényi, L. Gráf and A. Málnási-Csizmadia, Internal friction in enzyme reactions, IUBMB Life, 2013, 65, 35–42 CrossRef CAS PubMed.
  1300. D. Tapscott and A. Williams, Wikinomics: how mass collaboration changes everything, New Paradigm, 2007 Search PubMed.
  1301. A. Rinaldi, Science wikinomics. Mass networking through the web creates new forms of scientific collaboration, EMBO Rep., 2009, 10, 439–443 CrossRef CAS PubMed.
  1302. D. Corne and J. Knowles, No free lunch and free leftovers theorems for multiobjecitve optimisation problems., in Evolutionary Multi-criterion Optimization (EMO 2003), LNCS, ed. C. Fonseca et al., Springer, Berlin, 2003, vol. 2632, pp. 327–341 Search PubMed.
  1303. J. C. Culberson, On the futility of blind search: an algorithmic view of 'no free lunch', Evol. Comput., 1998, 6, 109–127 CrossRef CAS.
  1304. N. J. Radcliffe and P. D. Surry, Fundamental limitations on search algorithms: evolutionary computing in perspective, Computer Science Today, 1995, 1995, 275–291 Search PubMed.
  1305. D. H. Wolpert and W. G. Macready, No Free Lunch theorems for optimization, IEEE Trans. Evol. Comput., 1997, 1, 67–82 CrossRef.
  1306. J. E. Rowe, M. D. Vose and A. H. Wright, Reinterpreting no free lunch, Evol. Comput., 2009, 17, 117–129 CrossRef PubMed.
  1307. J. G. Zalatan and D. Herschlag, The far reaches of enzymology, Nat. Chem. Biol., 2009, 5, 516–520 CrossRef CAS PubMed.
  1308. Computational approaches in cheminformatics and bioinformatics, ed. R. Guha and A. Bender, Wiley, Hoboken, NJ, 2012 Search PubMed.
  1309. S. Ananiadou, D. B. Kell and J.-i. Tsujii, Text Mining and its potential applications in Systems Biology, Trends Biotechnol., 2006, 24, 571–579 CrossRef CAS PubMed.
  1310. S. Ananiadou, S. Pyysalo, J. i. Tsujii and D. B. Kell, Event extraction for systems biology by text mining the literature, Trends Biotechnol., 2010, 28, 381–390 CrossRef CAS PubMed.
  1311. S. Ananiadou, P. Thompson, R. Nawaz, J. McNaught and D. B. Kell, Event Based Text Mining for Biology and Functional Genomics, Briefings Funct. Genomics, 2014 DOI:10.1093/bfgp/elu1015.
  1312. M. J. Herrgård, N. Swainston, P. Dobson, W. B. Dunn, K. Y. Arga, M. Arvas, N. Blüthgen, S. Borger, R. Costenoble, M. Heinemann, M. Hucka, N. Le Novère, P. Li, W. Liebermeister, M. L. Mo, A. P. Oliveira, D. Petranovic, S. Pettifer, E. Simeonidis, K. Smallbone, I. Spasić, D. Weichart, R. Brent, D. S. Broomhead, H. V. Westerhoff, B. Kırdar, M. Penttilä, E. Klipp, B. Ø. Palsson, U. Sauer, S. G. Oliver, P. Mendes, J. Nielsen and D. B. Kell, A consensus yeast metabolic network obtained from a community approach to systems biology, Nat. Biotechnol., 2008, 26, 1155–1160 CrossRef PubMed.
  1313. N. Swainston, P. Mendes and D. B. Kell, An analysis of a ‘community-driven’ reconstruction of the human metabolic network, Metabolomics, 2013, 9, 757–764 CrossRef CAS PubMed.
  1314. P. Seeman, The membrane actions of anesthetics and tranquilizers, Pharmacol. Rev., 1972, 24, 583–655 CAS.
  1315. D. B. Kell and P. D. Dobson, The cellular uptake of pharmaceutical drugs is mainly carrier-mediated and is thus an issue not so much of biophysics but of systems biology, in Proc. Int. Beilstein Symposium on Systems Chemistry, ed. M. G. Hicks and C. Kettner, Logos Verlag, Berlin, 2009, pp. 149–168, http://www.beilstein-institut.de/Bozen2008/Proceedings/Kell/Kell.pdf Search PubMed.
  1316. R. Doshi, T. Nguyen and G. Chang, Transporter-mediated biofuel secretion, Proc. Natl. Acad. Sci. U. S. A., 2013, 110, 7642–7647 CrossRef CAS PubMed.
  1317. H. Ling, B. Chen, A. Kang, J. M. Lee and M. W. Chang, Transcriptome response to alkane biofuels in Saccharomyces cerevisiae: identification of efflux pumps involved in alkane tolerance, Biotechnol. Biofuels, 2013, 6, 95 CrossRef CAS PubMed.
  1318. B. Chen, H. Ling and M. W. Chang, Transporter engineering for improved tolerance against alkane biofuels in Saccharomyces cerevisiae, Biotechnol. Biofuels, 2013, 6, 21 CrossRef CAS PubMed.
  1319. J. L. Foo and S. S. J. Leong, Directed evolution of an E. coli inner membrane transporter for improved efflux of biofuel molecules, Biotechnol. Biofuels, 2013, 6, 81 CrossRef CAS PubMed.
  1320. N. Nishida, N. Ozato, K. Matsui, K. Kuroda and M. Ueda, ABC transporters and cell wall proteins involved in organic solvent tolerance in Saccharomyces cerevisiae, J. Biotechnol., 2013, 165, 145–152 CrossRef CAS PubMed.
  1321. C. Grant, D. Deszcz, Y. C. Wei, R. J. Martinez-Torres, P. Morris, T. Folliard, R. Sreenivasan, J. Ward, P. Dalby, J. M. Woodley and F. Baganz, Identification and use of an alkane transporter plug-in for applications in biocatalysis and whole-cell biosensing of alkanes, Sci. Rep., 2014, 4, 5844 Search PubMed.
  1322. M. Jasiński, Y. Stukkens, H. Degand, B. Purnelle, J. Marchand-Brynaert and M. Boutry, A plant plasma membrane ATP binding cassette-type transporter is involved in antifungal terpenoid secretion, Plant Cell, 2001, 13, 1095–1107 CrossRef.
  1323. K. Yazaki, ABC transporters involved in the transport of plant secondary metabolites, FEBS Lett., 2006, 580, 1183–1191 CrossRef CAS PubMed.
  1324. Q. Wu, M. Kazantzis, H. Doege, A. M. Ortegon, B. Tsang, A. Falcon and A. Stahl, Fatty acid transport protein 1 is required for nonshivering thermogenesis in brown adipose tissue, Diabetes, 2006, 55, 3229–3237 CrossRef CAS PubMed.
  1325. Q. Wu, A. M. Ortegon, B. Tsang, H. Doege, K. R. Feingold and A. Stahl, FATP1 is an insulin-sensitive fatty acid transporter involved in diet-induced obesity, J. Mol. Cell Biol., 2006, 26, 3455–3467 CrossRef CAS PubMed.
  1326. D. Khnykin, J. H. Miner and F. Jahnsen, Role of fatty acid transporters in epidermis: Implications for health and disease, Dermatoendocrinol, 2011, 3, 53–61 CrossRef CAS PubMed.
  1327. M. H. Lin and D. Khnykin, Fatty acid transporters in skin development, function and disease, Biochim. Biophys. Acta, 2014, 1841, 362–368 CrossRef CAS PubMed.
  1328. M. S. Villalba and H. M. Alvarez, Identification of a novel ATP-binding cassette transporter involved in long-chain fatty acid import and its role in triacylglycerol accumulation in Rhodococcus jostii RHA1, Microbiology, 2014, 160, 1523–1532 CrossRef CAS PubMed.
  1329. R. Gimenez, M. F. Nuñez, J. Badia, J. Aguilar and L. Baldoma, The gene yjcG, cotranscribed with the gene acs, encodes an acetate permease in Escherichia coli, J. Bacteriol., 2003, 185, 6448–6455 CrossRef CAS.
  1330. R. Islam, N. Anzai, N. Ahmed, B. Ellapan, C. J. Jin, S. Srivastava, D. Miura, T. Fukutomi, Y. Kanai and H. Endou, Mouse organic anion transporter 2 (mOat2) mediates the transport of short chain fatty acid propionate, J. Pharmacol. Sci., 2008, 106, 525–528 CrossRef CAS.
  1331. I. Moschen, A. Bröer, S. Galić, F. Lang and S. Bröer, Significance of short chain fatty acid transport by members of the monocarboxylate transporter family (MCT), Neurochem. Res., 2012, 37, 2562–2568 CrossRef CAS PubMed.
  1332. J. Sá-Pessoa, S. Paiva, D. Ribas, I. J. Silva, S. C. Viegas, C. M. Arraiano and M. Casal, SATP (YaaH), a succinate-acetate transporter protein in Escherichia coli, Biochem. J., 2013, 454, 585–595 CrossRef PubMed.
  1333. R. Kaldenhoff, L. Kai and N. Uehlein, Aquaporins and membrane diffusion of CO2 in living organisms, Biochim. Biophys. Acta, 2014, 1840, 1592–1595 CrossRef CAS PubMed.
  1334. L. Kai and R. Kaldenhoff, A refined model of water and CO2 membrane diffusion: Effects and contribution of sterols and proteins, Sci. Rep., 2014, 6665 CrossRef PubMed.
  1335. M. Galdzicki, K. P. Clancy, E. Oberortner, M. Pocock, J. Y. Quinn, C. A. Rodriguez, N. Roehner, M. L. Wilson, L. Adam, J. C. Anderson, B. A. Bartley, J. Beal, D. Chandran, J. Chen, D. Densmore, D. Endy, R. Grunberg, J. Hallinan, N. J. Hillson, J. D. Johnson, A. Kuchinsky, M. Lux, G. Misirli, J. Peccoud, H. A. Plahar, E. Sirin, G. B. Stan, A. Villalobos, A. Wipat, J. H. Gennari, C. J. Myers and H. M. Sauro, The Synthetic Biology Open Language (SBOL) provides a community standard for communicating designs in synthetic biology, Nat. Biotechnol., 2014, 32, 545–550 CrossRef CAS PubMed.
  1336. N. Roehner, E. Oberortner, M. Pocock, J. Beal, K. Clancy, C. Madsen, G. Misirli, A. Wipat, H. Sauro and C. J. Myers, A Proposed Data Model for the Next Version of the Synthetic Biology Open Language, ACS Synth. Biol., 2014 DOI:10.1021/sb500176h.
  1337. M. Hucka, A. Finney, H. M. Sauro, H. Bolouri, J. C. Doyle, H. Kitano, A. P. Arkin, B. J. Bornstein, D. Bray, A. Cornish-Bowden, A. A. Cuellar, S. Dronov, E. D. Gilles, M. Ginkel, V. Gor, I. I. Goryanin, W. J. Hedley, T. C. Hodgman, J. H. Hofmeyr, P. J. Hunter, N. S. Juty, J. L. Kasberger, A. Kremling, U. Kummer, N. Le Novere, L. M. Loew, D. Lucio, P. Mendes, E. Minch, E. D. Mjolsness, Y. Nakayama, M. R. Nelson, P. F. Nielsen, T. Sakurada, J. C. Schaff, B. E. Shapiro, T. S. Shimizu, H. D. Spence, J. Stelling, K. Takahashi, M. Tomita, J. Wagner and J. Wang, The systems biology markup language (SBML): a medium for representation and exchange of biochemical network models, Bioinformatics, 2003, 19, 524–531 CrossRef CAS PubMed.
  1338. N. Le Novère, M. Hucka, H. Mi, S. Moodie, F. Schreiber, A. Sorokin, E. Demir, K. Wegner, M. Aladjem, S. M. Wimalaratne, F. T. Bergman, R. Gauges, P. Ghazal, K. Hideya, L. Li, Y. Matsuoka, A. Villéger, S. E. Boyd, L. Calzone, M. Courtot, U. Dogrusoz, T. Freeman, A. Funahashi, S. Ghosh, A. Jouraku, S. Kim, F. Kolpakov, A. Luna, S. Sahle, E. Schmidt, S. Watterson, G. Wu, I. Goryanin, D. B. Kell, C. Sander, H. Sauro, J. L. Snoep, K. Kohn and H. Kitano, The systems biology graphical notation, Nat. Biotechnol., 2009, 27, 735–741 CrossRef PubMed.
  1339. N. Le Novère, A. Finney, M. Hucka, U. S. Bhalla, F. Campagne, J. Collado-Vides, E. J. Crampin, M. Halstead, E. Klipp, P. Mendes, P. Nielsen, H. Sauro, B. Shapiro, J. L. Snoep, H. D. Spence and B. L. Wanner, Minimum information requested in the annotation of biochemical models (MIRIAM), Nat. Biotechnol., 2005, 23, 1509–1515 CrossRef PubMed.
  1340. M. Courtot, N. Juty, C. Knüpfer, D. Waltemath, A. Zhukova, A. Dräger, M. Dumontier, A. Finney, M. Golebiewski, J. Hastings, S. Hoops, S. Keating, D. B. Kell, S. Kerrien, J. Lawson, A. Lister, J. Lu, R. Machne, P. Mendes, M. Pocock, N. Rodriguez, A. Villéger, D. J. Wilkinson, S. Wimalaratne, C. Laibe, M. Hucka and N. Le Novère, Controlled vocabularies and semantics in Systems Biology, Mol. Syst. Biol., 2011, 7, 543 CrossRef PubMed.
  1341. R. Goodacre, D. Broadhurst, A. Smilde, B. S. Kristal, J. D. Baker, R. Beger, C. Bessant, S. Connor, G. Capuani, A. Craig, T. Ebbels, D. B. Kell, C. Manetti, J. Newton, G. Paternostro, R. Somorjai, M. Sjöström, J. Trygg and F. Wulfert, Proposed minimum reporting standards for data analysis in metabolomics, Metabolomics, 2007, 3, 231–241 CrossRef CAS.
  1342. C. B. Anfinsen, E. Haber, M. Sela and F. H. White, The kinetics of formation of native ribonuclease during oxidation of the reduced polypeptide chain, Proc. Natl. Acad. Sci. U. S. A., 1961, 47, 1309–1314 CrossRef CAS.
  1343. C. B. Anfinsen, Principles that govern the folding of protein chains, Science, 1973, 181, 223–230 CAS.
  1344. T. Buzan, How to mind map, Thorsons, London, 2002 Search PubMed.
  1345. D. B. Kell, Genotype:phenotype mapping: genes as computer programs, Trends Genet., 2002, 18, 555–559 CrossRef CAS.
  1346. D. Broadhurst and D. B. Kell, Statistical strategies for avoiding false discoveries in metabolomics and related experiments, Metabolomics, 2006, 2, 171–196 CrossRef CAS.

This journal is © The Royal Society of Chemistry 2015