Next Article in Journal
Copper-Enhanced NiMo/TiO2 Catalysts for Bifunctional Green Hydrogen Production and Pharmaceutical Pollutant Removal
Previous Article in Journal
Theoretical Modeling of Pathways of Transformation of Fructose and Xylose to Levulinic and Formic Acids over Single Na Site in BEA Zeolite
Previous Article in Special Issue
Optimized Tandem Affinity Purification Strategy Enables High-Yield Isolation and Functional Characterization of Native COMPASS in Saccharomyces cerevisiae
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

How Mechanistic Enzymology Helps Industrial Biocatalysis: The Case for Kinetic Solvent Viscosity Effects

by
Gabriel Atampugre Atampugbire
1,
Joanna Afokai Quaye
1 and
Giovanni Gadda
1,2,3,*
1
Department of Chemistry, Georgia State University, Atlanta, GA 30302-3965, USA
2
Department of Biology, Georgia State University, Atlanta, GA 30302-3965, USA
3
The Center for Diagnostics and Therapeutics, Georgia State University, Atlanta, GA 30302-3965, USA
*
Author to whom correspondence should be addressed.
Catalysts 2025, 15(8), 736; https://doi.org/10.3390/catal15080736 (registering DOI)
Submission received: 30 June 2025 / Revised: 29 July 2025 / Accepted: 30 July 2025 / Published: 1 August 2025
(This article belongs to the Special Issue Enzyme Engineering—the Core of Biocatalysis)

Abstract

Biocatalysis is one of the oldest fields that has been used in industrial applications, with one of the earliest purposeful examples being the mass production of acetic acid from an immobilized Acinetobacter strain in the year 1815. Efficiency, specificity, reduced reaction times, lower overall costs, and environmental friendliness are some advantages biocatalysis has over conventional chemical synthesis, which has made biocatalysis increasingly used in industry. We highlight three necessary fields that are fundamental to advancing industrial biocatalysis, including biocatalyst engineering, solvent engineering, and mechanistic engineering. However, the fundamental mechanism of enzyme function is often overlooked or given less attention, which can limit the engineering process. In this review, we describe how mechanistic enzymology benefits industrial biocatalysis by elucidating key fundamental principles, including the kcat and kcat/Km parameters. Mechanistic enzymology presents a unique field that provides in-depth insights into the molecular mechanisms of enzyme activity and includes areas such as reaction kinetics, catalytic mechanisms, structural analysis, substrate specificity, and protein dynamics. In line with the objective of protein engineering to optimize enzyme activity, we summarize a range of strategies reported in the literature aimed at improving the product release rate, the chemical step of catalysis, and the overall catalytic efficiency of enzymes. Further into this review, we delineate kinetic solvent viscosity effects (KSVEs) as a very efficient, cost-effective, and easy-to-perform method to probe different aspects of enzyme reaction mechanisms, including diffusion-dependent kinetic steps and rate-limiting steps. KSVEs are cost-effective because simple kinetic enzyme assays, such as the Michaelis–Menten kinetic approach, can be combined with them without the need for specialized and costly equipment. Other techniques in protein engineering and genetic engineering are also covered in this review. Additionally, we provide information on solvent systems in enzymatic reactions, details on immobilized biocatalysts, and common misconceptions that misguide enzyme design and optimization processes.

Graphical Abstract

1. Introduction

Hudges and Lewis defined biocatalysis as the application of enzymes to transform molecules into desired products [1] (Figure 1). Biocatalysis has been identified as an alternative to conventional chemical synthesis because it uses natural or engineered enzymes and microorganisms to accelerate chemical reactions [2,3]. Several thousand years of history can be traced in industrial biocatalysis, with the earliest applications around 6000 BC or earlier [4,5]. One of the oldest purposeful applications of biocatalysis is the large-scale production of acetic acid from ethanol by an Acinetobacter strain in 1815 [6]. Biocatalysis has since been a rapidly developing and expanding field in industry because of its benefits, which include improved efficiency and specificity, shorter reaction times, lower overall costs, and environmental friendliness [7]. The production of several pharmaceutical compounds, including those used to treat heart disease, cancer, and bacterial and viral infections, has been achieved through biocatalysis [8,9].
The primary focus of industrial protein engineering is to develop enzymes with optimized catalytic activity, substrate affinity and specificity, binding affinity, and stability by adopting a rational design or directed evolution approach [10,11] (Figure 2). We noticed a trend in some biocatalytic studies, especially in industrial research, where, in the quest to modify enzymes for optimal turnover rates, the underlying mechanisms of enzyme function were overlooked or given less priority. When there is a limited understanding of the mechanistic changes occurring in a modified enzyme, the likelihood of encountering challenges in predicting the stability and scalability of the enzyme is high [10,12]. Additionally, the potential for further optimization may be reduced by a lack of mechanistic information. To illustrate, by identifying the rate-limiting steps in a reaction, approaches that target modifications of such steps in the catalytic pathway can be prioritized. Another disadvantage of a results-oriented approach is the lack of control over tradeoffs during enzyme modifications since the absence of thorough knowledge of the mechanistic pathways makes it challenging to monitor and determine the effect of such changes. A prime example is a study that sought to modify the activity of formate dehydrogenase (FDH) from Candida methylica (CmFDH) [13]. Following site-directed mutagenesis, the mutants showed a 6.5-fold increase in the kcat value but also an increased Km. The authors described the observed higher Km value as a necessary compensatory mechanism in the enzyme’s active site to make room for the substrate when the bulky residues, Tyr and His, were substituted at Asn119 [13]. Therefore, the precise engineering of enzymes requires the integration of mechanistic studies.
Protein engineering has made significant progress, with a number of noteworthy advancements in rational design and directed evolution techniques. The rational design of enzymes requires an in-depth understanding of enzyme structures and mechanisms and utilizes a moderately sized library [14] (Figure 3). On the other hand, directed evolution only needs a basic understanding of the structure and function of enzymes and requires a larger library size [14] (Figure 3). Regardless of which approach is used in enzyme design, the desired functionality can be achieved by addressing tradeoffs between activity and stability. In protein engineering, the two common techniques of rational design include site-directed mutagenesis and domain swapping. Site-directed mutagenesis is a technique that is used to introduce specific, targeted changes to the DNA sequence of a protein, whereas domain swapping involves the exchange of structural elements between two or more identical protein monomers to form dimers or multimers with desired novel protein properties [15,16]. Merz et al. used a rational approach to enhance the catalytic activity of the thermophilic enzyme indoleglycerol phosphate synthase (IGPS) at low temperatures by mutating the trpC gene, which encodes IGPS [17]. Further investigation of the IGPS mechanistic pathway showed that the low activity of the enzyme was caused by its rigid structure. Specific amino acid substitutions were conducted to increase flexibility, which significantly increased catalytic turnover at low temperatures. The Nobel Prize-winning technique, directed evolution, aims to accelerate the process of natural evolution in biological molecules and systems in a test tube through repetitive rounds of gene variation and screening [18,19,20,21]. The development of continuous directed evolution systems, like yeast OrthoRep, is a recent advancement in directed evolution. Such systems allow for simultaneous hypermutation and selection in vivo at rates that are roughly 100,000 times faster than the host genome [22]. Protein engineering processes have been integrated with computational design and machine learning techniques to improve the prediction of protein properties and to facilitate the rapid exploration of vast sequence spaces [23]. ProteinMPNN has demonstrated high success rates in de novo enzyme design [24], while deep learning models, such as AlphaFold and RoseTTAFold, have greatly enhanced protein structure prediction [25]. Ancestral sequence reconstruction (ASR) is another trend in enzyme engineering design and optimization, which involves inferring and resurrecting ancestral protein sequences to create enzymes with enhanced properties [26]. Thermophilic enzymes that maintain high catalytic activity at lower temperatures have been engineered using ASR [26]. Historical sequence changes in the structure and function of enzymes can also be identified via ASR, which makes it possible to reconstruct a protein family’s evolutionary history and provide causal insights into functional diversities [27].
Biocatalysis has benefited from genetic engineering, which is a closely related field to protein engineering, differing in focus and techniques used. Genetic engineering is a process that involves the direct manipulation of an organism’s genes to enhance, modify, or remove specific properties pertaining to the organism [28]. In the context of biocatalysis, genetic engineering utilizes techniques such as Clustered Regularly Interspaced Short Palindromic Repeats (CRISPRs), cloning, and transformation to create new DNA constructs, commonly referred to as recombinant DNA [29]. Genetic engineering was used to construct a novel chimeric lysin (Ply187N-V12C) by fusing the catalytic domain of the bacteriophage lysin Ply187 with the cell-binding domain of the lysin PlyV12, which resulted in a hybrid that showed a broader lytic spectrum or host range [30]. Similarly, a soluble recombinant chimeric protein (CmAP/MBL-AJ) was created by fusing the genes encoding the marine bacterium Cobetia marina alkaline phosphatase (CmAP) and Far Eastern holothurian Apostichopus japonicus mannan-binding C-type lectin (MBL-AJ) [31]. The bifunctional hybrid demonstrated higher alkaline phosphatase activity and specificity relative to the wild-type. Table 1 below summarizes some genetically engineered proteins with their corresponding modified structural properties and industrial applications.
While enzyme engineering has been significantly improved by developments such as directed evolution and computational design, mechanistic enzymology holds a unique and crucial role by focusing on the molecular basis of enzyme-catalyzed reactions. Mechanistic enzymology examines detailed aspects of enzymes, including reaction kinetics, catalytic mechanisms, structural analysis, substrate specificity, protein dynamics, and inhibition studies [41,42]. Knowledge obtained from a detailed mechanistic study is relevant in the rational design and optimization of enzymes for specific industrial purposes. Mechanistic studies enable the successful identification and alteration of key residues and motifs that are essential to catalytic pathways [14,43]. Mechanistic studies also promote sustainable chemical synthesis by expanding the repertoire of available biocatalytic transformations [44,45]. Furthermore, through mechanistic enzymology, conserved residue networks have been discovered to connect distal regions of enzymes to their active sites [46,47]. This discovery implies that the conventional paradigm of “structure encodes function” should be expanded to incorporate the role of protein dynamics and energy flow in enzyme activity [3].
In Figure 4, we present a conceptual framework that highlights three important areas of industrial biocatalysis: biocatalyst engineering, solvent engineering, and mechanistic enzymology. Biocatalyst engineering includes those processes that directly modify the intrinsic properties of the enzyme, such as protein engineering and genetic engineering. Processes that involve changes to the enzyme’s surrounding environment (extrinsic), including the choice of solvent system, are classified under solvent engineering. Mechanistic enzymology, although sometimes overlooked, represents a core area that provides insights into enzyme behavior and will be the focus of this review.

2. The Significance of Mechanistic Studies in Enzyme Engineering

Different techniques are implemented by mechanistic enzymology to obtain in-depth details of the structural and functional properties of enzymes [48]. One of the most widely used techniques in enzymology, site-directed mutagenesis, is used to probe the potential roles of specific amino acid residues by comparing the function and structure of wild-type and mutant enzymes [49,50,51]. Specific mutagenesis experiments not only complement structural and kinetic data but also confirm the catalytic role of active-site side residues of an enzyme [52]. For instance, research on alkaline phosphatase has shown that enzyme activity is reduced by 88,000 times when the metal-coordinating catalytic residue R166 is changed to a serine, suggesting that R166 plays a crucial role in Mg2+ coordination and transition-state stabilization [53]. In addition to single-site mutations, combinatorial mutagenesis of multiple residues can be used to assess cooperative interactions between spatially distant residues within a protein’s structure. In the same alkaline phosphatase study, mutants of five structurally linked residues, namely, D101, D153, R166, E322, and K328, were individually restored, and the results showed that the enzyme’s catalytic efficiency depended on the combined presence of all five residues, hence indicating synergistic interactions among them [53].
In addition to understanding the unique roles of amino acid residues in enzyme dynamics, information about enzyme function can be provided by solvent-sensitive steps in catalysis. Kinetic solvent viscosity effects (KSVEs) are a powerful technique that analyzes how changes in solvent viscosity affect the kinetic parameters of an enzymatic reaction to provide unique perspectives on the factors influencing enzyme function [48]. The theoretical basis of KSVEs, Kramers’ theory, lies in the principle that increased solvent viscosity creates friction against proteins in solution, decreasing motion and potentially inhibiting catalysis in motile enzymes [54]. The KSVE technique is considered a highly efficient and cost-effective technique because simple kinetic enzyme assays, such as the Michaelis–Menten kinetic approach, can be combined with it without requiring specialized and costly equipment. Techniques such as site-directed mutagenesis, pH effects, and kinetic isotope effects have been combined with KSVEs in recent research to provide detailed insights into the mechanisms of action of several enzymes, including D-2-hydroxyglutarate dehydrogenase, NADH:quinone oxidoreductase, and nitronate monooxygenase [55,56,57]. In a study on Pseudomonas aeruginosa NADH:quinone oxidoreductase, KSVEs with glucose and sucrose were used to probe the presence of a solvent-sensitive isomerization within the enzyme–substrate complex [56]. Further analysis revealed that flavin reduction was the rate-limiting step for the substrate benzoquinone, whereas hydride transfer from NADH to the FMN and quinol product release were partially rate-limiting for the substrate juglone [56]. Further details on the methods used to study KSVEs are covered in a review by Gadda and Sobrado [48]. Different aspects of enzyme reaction mechanisms can be probed using KSVEs, including diffusion-controlled substrate binding and rate-limiting steps, which could be one of or a combination of isomerization steps, the product release step, or the chemical step of catalysis [48]. According to the Stokes–Einstein equation, a molecule’s diffusion coefficient (D) is inversely proportional to solvent viscosity (η) [58]. An increase in solvent viscosity raises the frictional resistance experienced by enzymes and substrates in solution [59]. Consequently, any kinetic step that is diffusion-dependent, including substrate association, substrate dissociation, or product release, will slow down in direct proportion to the viscosity change [59]. In KSVE experiments, this is reflected as a linear decrease in the kcat/Km and kcat values, with slopes approaching 1 for fully diffusion-controlled steps, intermediate slopes (0 < m < 1) for mixed control, and a slope nearing zero when the chemical step of catalysis is rate-limiting [48]. Diffusion controls how rapidly enzyme and substrate molecules encounter each other and how quickly products exit the active site, establishing an upper limit to catalytic efficiency (kcat/Km ≈ 108–109 M−1·s−1 for large enzymes) [48]. The KSVE methodology essentially involves setting up protocols to measure initial velocity data in the presence of varying concentrations of an additive, usually polyhydroxylated compounds. When two substrates are required for complete enzyme activity, the concentration of one substrate may be fixed, while the concentration of the other is varied [11]. The same initial velocity measurement is repeated with the addition of different concentrations of a viscosigen, i.e., 0%, 9%, 15%, 30%, and 40%. Solvent microviscosigens, such as glucose, glycerol, and sucrose, create resistance to the motion of molecules in solution and are associated with the internal diffusion of a substrate into the active site of an enzyme [48] (Figure 5). Solvent macroviscosigens, on the other hand, include polymeric additives, such as Ficoll-400 and polyethylene glycol (PEG), and determine the free volume available to molecules moving in solution [48]. Increasing solvent macroviscosity can be used as a control for KSVEs to distinguish whether the observed viscosity effect arises from diffusion limitations rather than macromolecular crowding [48] (Figure 5). Table 2 below lists some microviscosigens and macroviscosigens, along with their respective typical molecular weights.
The data are plotted using the Michaelis–Menten equation to establish a hyperbolic dependence of the rate on substrate concentration. Further analysis may be followed by making a plot of normalized kinetic parameters, such as the kcat or kcat/Km parameters (thus, “(kcat)o/(kcat)η or (kcat/Km)o/(kcat/Km)η)” against relative viscosity (ηrel) using Equation (1). The kinetic parameters kcat/Km and kcat are often preferred because they can easily be measured, and significant mechanistic information can be obtained [48,60]. In-depth details and references on KSVEs were summarized in a review by Gadda and Sobrado [48].
( k ) o ( k ) η = m η r e l 1 + 1
In Equation (1), the slope of the line, m, indicates how much the normalized kinetic parameter depends on viscosity [61]. A linear correlation between the relative kcat ((kcat)o/(kcat)η) and the relative viscosity (η/η0) is indicative of the presence of viscosity-dependent or solvent-sensitive steps in the enzyme catalytic cycle [59]. Analysis of the resulting trends can provide insights into the reaction mechanism, for instance, the slope of the line that fits the data (Equation (1)) for linear dependence of the data in a plot of (kcat/Km)o/(kcat/Km)η versus ηrel (Figure 6) is consistent with the capture of the substrate in enzyme−substrate complexes that is fully (m = 1) or partially (0 < m < 1) committed to catalysis in the forward direction [48]. Similarly, the slope of the line that fits the data (Equation (1)) for linear dependence of the data in a plot of (kcat)o/(kcat)η versus ηrel (Figure 7) is consistent with the overall turnover of the enzyme being fully (m = 1) or partially (0 < m < 1) limited by the product release rate [48,56]. A zero slope, for both plots, indicates a reaction that is consistent with the absence of forward and reverse commitments to catalysis on the kcat or kcat/Km values [48]. Additionally, a solvent-sensitive internal isomerization of the enzyme−substrate or enzyme−product complex is consistent with a hyperbolic trend in a plot of (kcat)o/(kcat)η or (kcat/Km)o/(kcat/Km)η versus ηrel during enzyme catalysis [48].

3. Experimental Methods in Enzymology

The broader implications of using the technique of KSVEs were discussed in the previous section; however, the knowledge obtained from KSVEs can also be complemented by combining several other important techniques in mechanistic enzymology. Nonetheless, these techniques may require expertise and resources that are not readily available to most laboratories [11]. The X-ray crystallography technique can be used to determine the 3D structure of a protein by producing an electron density map of the atoms within the protein [62]. The detailed architecture of the active site of an enzyme and the structures of intermediates occurring in the reaction pathway can also be elucidated using X-ray crystallography [63,64]. Hydrogen/deuterium exchange mass spectrometry (HDX-MS) is a solution-state structural technique that can be used for the direct monitoring of enzyme behavior by linking structure, conformational dynamics, and function under native conditions [65,66,67]. HDX-MS provides information on protein conformational dynamics [68], protein–membrane interactions [69], and folding dynamics [70] by analyzing the mass differences resulting from isotopic exchange between the protein backbone amide hydrogens and the surrounding solvent. Nuclear magnetic resonance (NMR) is another technique that is used to elucidate the structure and dynamics of an enzyme. NMR determines the 3D structure, monitors conformational changes during catalysis, and identifies important catalytic residues in the reaction pathway by using the magnetic properties of the enzyme’s atomic nuclei [71,72,73]. NMR can also monitor the interaction between an enzyme and its substrate throughout a reaction process, and the ionization states of catalytic residues can be determined using NMR [74,75]. However, interpreting data from structural techniques, such as NMR, without kinetic data can lead to errors, as enzymes typically complete their catalytic cycles in milliseconds, whereas NMR data collection usually takes hours [76,77,78]. Reaction by-products produced at rates slower than the enzyme turnover, for example, could be mistakenly identified as important intermediates in the reaction pathway when relying only on NMR data [76,79,80].
Reaction pathways are monitored using enzyme kinetic techniques, from substrate binding at the active site of an enzyme to product release. The transient-state kinetic technique provides valuable insights into the pathways of reactions catalyzed by enzymes, including the direct observation of intermediates and products [81]. The individual, rapid steps occurring in a reaction can also be directly observed and quantified in real time via transient kinetic analysis [76,82]. Another technique, kinetic isotope effects (KIEs), has been used for decades to investigate mechanisms of reactions by exploiting the differences in reaction rates when atoms of reactant molecules are replaced by their respective isotopes [83]. In addition to predicting whether two or more pathways of reactions share a similar transition state [83,84,85], KIEs can reveal which bond is broken or formed at each stage of a catalytic reaction [85,86]. The KIE method can also be used to determine protein dynamics, hybridization states, rate-limiting steps, and binding effects [48,87]. The ionization states of substrates or active site amino acids involved in enzyme activity can also be deduced by the method of pH-dependent enzyme kinetics [48,88].
A critical component to the stability and efficiency of enzymes is the reaction medium used for the enzymatic reaction. Solvent engineering is a tool that presents numerous opportunities for optimizing enzyme properties and involves the strategic design and manipulation of solvent systems to improve the solubility of reactants and products, enzyme stability, activity, and selectivity [89]. In biocatalysis, water is the most frequently used solvent system in biocatalytic transformations; however, the high polarity of water creates a drawback for hydrophobic reactants, which are commonly used in synthetic reactions [90]. The use of organic solvent systems for enzymatic reactions has become a trend because of the increased solubility of hydrophobic substances, especially when water-dependent side reactions need to be suppressed [91]. Ionic liquids have also been used as reaction media for the past two decades because of their wide thermal stability, low flammability, and low vapor pressure [92,93,94,95]. Butyl imidazolium chloride has demonstrated superior catalytic performance compared to other ionic solvents in the synthesis of the bio-based bis-epoxide compound, bis(oxiran-2-ylmethyl) furan-2,5-dicarboxylate, offering a simpler, more scalable, and cost-effective solution suitable for large-scale applications [96]. Other ionic solvents that were tested were summarized in a study by Chícharo et al. [96]. However, the usage of ionic solvents is often limited due to low sustainability, high cost, poor biodegradability, and poor biocompatibility. Deep eutectic solvents (DESs), as alternatives to ionic liquids, were first introduced by Abbott et al., 2003 [97] when they observed that mixtures of choline chloride and urea possessed significantly lower melting points relative to their pure components [98,99]. DESs are eutectics formed by mixing a hydrogen bond acceptor (HBA) and a hydrogen bond donor (HBD) with simultaneous heating and stirring until a liquid is formed [99]. This results in a self-association to create a new eutectic phase that is characterized by a significantly reduced melting point compared to the individual components [100]. An example of DESs is choline chloride–ethylene glycol (1:2) with a Tm of 66 °C. DESs are increasingly used as solvent systems in biocatalysis because they are nontoxic, nonvolatile, nonflammable, biodegradable, and considered “green” solvents [90]. Other advantages include their ease of preparation and lower cost [90]. The efficacy of different organic solvent systems, such as isopropylamine, oxalic acid, formic acid, and ethyl acetate, was studied and reported by He et al. [101]. Other approaches in solvent engineering have utilized surfactants as various reaction media due to their unique amphiphilic nature to enhance the solubility of hydrophobic substances. An advanced approach known as Pickering emulsion uses solid particles, called Pickering particles, which possess unique properties that enable them to adsorb onto and stabilize the interface between two immiscible components, thereby preventing separation [102]. A study by Wang et al. utilized certain bacteria as Pickering particles, taking advantage of their size and surface hydrophobicity to stabilize Pickering emulsions [103]. A repertoire of diverse Pickering particles was provided in a review by Omar and Sadeghi [104]. Table 3 below contains a list of DESs with their corresponding HBAs and HBDs, mole ratios, and melting temperatures. An extensive database of DESs was summarized in a review by Omar and Sadeghi [97].
While solvent systems can be considered for enzyme stability reasons, the state of the enzyme is also vital to its function. Immobilizing enzymes onto fixed support systems can overcome certain drawbacks of enzymes in the free state, including limited reusability, especially for industrial applications like waste removal, and susceptibility to denaturation [115]. Enzymes can be anchored via adsorption, covalent binding, encapsulation, or cross-linking in order to maintain high catalytic activity over extended periods, operate across a broader pH and temperature range, and withstand other harsh conditions [116]. In order to retain high catalytic activity, laccase from Pleurotus florida NCIM 1243 was immobilized by adsorption onto cellulose nanofibers, and the study results showed that the enzyme was capable of removing micropollutants, over 90% for Remazol Brilliant Blue R and 50% for Remazol Black 5 [117]. In a different study, laccase was adsorbed onto polyacrylonitrile–biochar composite nanofibrous membrane, and it was observed that the immobilized enzyme showed over 60% of chlortetracycline removal efficiency from simulated wastewater [118]. Both studies on laccase demonstrated that immobilizing enzymes maintain significant activity rates and that the enzymes can be reused repeatedly.

4. The kcat Parameter—Enzyme Turnover Number

For an enzyme that catalyzes a reversible reaction with a single substrate yielding a single product, Scheme 1 below illustrates the simplest, but realistic, kinetic model. Note that, unlike in most biochemistry textbooks, the chemical step of catalysis and the physical step of product release are treated independently here.
We begin by explaining the kinetic parameter kcat, first by stating the concept of steady-state kinetics introduced by Briggs and Haldane in 1925 [119]. Steady-state kinetics is described by the condition of an enzyme-catalyzed reaction, where the concentration of the enzyme–substrate (ES) complex remains constant during the catalytic process [119,120]. Under steady-state conditions, the reaction rate for an enzyme (E) with a single substrate (S) can be given by the Michaelis–Menten equation (Equation (2)):
v = k c a t [ E ] T [ S ] K m + [ S ]
When the substrate concentration is much greater than Km, Equation (2) is simplified as
v = V m a x = k c a t [ E ] T
where [E]T is the total concentration of enzyme, v is the velocity of reaction, and Vmax is the maximum reaction velocity [121]. kcat represents the maximum velocity at which substrate molecules are converted by an enzyme into a product per unit time when the substrate fully saturates the enzyme [122]. Thus, kcat sets a lower limit on the rate of any step following substrate binding and is a function of the time it takes for the enzyme to bind the substrate through to the release of the product into solution by the enzyme [123]. Under saturating substrate conditions, kcat reflects the rate-limiting step(s) in the reaction pathway [121].
The conventional definition of the rate-limiting step of an enzyme-catalyzed reaction, i.e., the slowest step in the catalytic pathway that determines the overall rate of reaction, misrepresents the operation of a sequence of enzyme-catalyzed reactions [124]. Several studies have shown that reactions may have two or more steps, influencing the total reaction rate to varying extents [3]. Thus, the overall reaction rate can proceed only as fast as the combination of all rate-limiting steps, implying that the identification of the rate-limiting step(s) in enzyme-catalyzed reactions through mechanistic enzymology is important. Depending on the enzyme system, an isomerization step, the chemical step of catalysis (k3), and/or the product release step (k5) can be the rate-limiting step(s) [125]. In some enzymes, the rates of bond formation and cleavage occur faster than the rate of conformational change, suggesting that the conformational change is the rate-limiting step in the catalytic process [126,127,128]. It is essential to note that substrate binding is not considered a rate-limiting step, as the enzyme turnover number (kcat) is defined under conditions of substrate saturation [11]. A further explanation will be provided later in this section.
In Scheme 1, k2, k3, k4, and k5 are first-order rate constants, whereas k1 is a second-order rate constant for substrate association [129]. The rate equation for Scheme 1 can be derived using the steady-state approximation as [60,130]
v = k 1 k 3 k 5 [ S ] k 2 k 4 + k 2 k 5 + k 3 k 5 + k 1 ( k 3 + k 4 + k 5 ) [ S ] [ E ] T
When the substrate concentration is much greater than Km, the rate is determined by Equation (3), and Equation (4) is simplified as
V m a x = k c a t [ E ] T = k 3 k 5 k 3 + k 4 + k 5 [ E ] T
From Equation (5),
k c a t = k 3 k 5 k 3 + k 4 + k 5
It is clear from Equation (6) that kcat may be dependent on the catalytic step and product release, but it is not affected by substrate association and dissociation (k1 and k2, respectively). Equation (6) simplifies to Equation (7) when the chemical step of catalysis is irreversible (k4 ≈ 0) or when the enzyme–product complex tends to partition forward and release the product rather than reverting to the enzyme–substrate complex (k5 >> k4) [48].
k c a t = k 3 k 5 k 3 + k 5
Three scenarios will be described to illustrate how the rate-limiting step(s) of a reaction determine the overall turnover number (kcat) and how identifying such steps can increase the likelihood of successful targeted attempts to improve enzyme activity. For an enzyme-catalyzed reaction with a single substrate and product and an irreversible catalytic step, suppose k3 and k5 are 10 s−1 and 10 s−1, respectively; the estimated kcat from Equation (7) would be 5 s−1.
Scenario 1.
If the chemical step of catalysis is the only rate-limiting step, i.e., k5 >> k3, with k5 = 5 s−1 and k3 = 0.20 s−1, then kcat = 0.19 ≈ k3, implying that any efforts to improve the product release step (k5) will not increase the overall turnover number of the enzyme.
Scenario 2.
If the product dissociation step is the only rate-limiting step, i.e., k3 >> k5, with k3 = 5 s−1 and k5 = 0.20 s−1, then kcat = 0.19 ≈ k5, implying that any efforts to improve the chemical step of catalysis (k3) will not increase the overall turnover number of the enzyme.
Scenario 3.
If both the chemical step of catalysis and the product dissociation step are rate-limiting, i.e., k3 ≈ k5, with k5 = 0.20 s−1 and k3 = 0.20 s−1, then kcat = 0.1 s−1, implying that two rate-limiting steps generally give lower turnover numbers relative to a single rate-limiting step; however, any efforts to improve either the product release step (k5) or chemical step of catalysis (k3) will increase the overall turnover number of the enzyme.
A clear example of scenario 1 is found in a study by Krooshof et al. [130], where the role of Asp260 in the catalytic triad of haloalkane dehalogenase (DhlaA) was investigated by the method of site-directed mutagenesis [131]. A 12-fold increase in bromide release (the product), a 220-fold reduction in carbon–bromine bond cleavage (the chemical step of catalysis), and a 10-fold decrease in kcat were observed. Further analysis revealed that the chemical step of catalysis was rate-limiting, explaining why the increased product release did not enhance the overall turnover rate [131]. From the three scenarios above, performing mechanistic studies first to determine the rate-limiting step(s) of a reaction will guide approaches aimed at increasing the turnover number of an enzyme by targeting such steps that are relevant to kcat.
We mentioned earlier in this section that substrate binding is not considered a rate-limiting step because enzyme turnover is defined under conditions of substrate saturation. Thus, when the substrate concentration is much less than Km, the rate is determined by kcat/Km, and Equation (2) is simplified as
v = k c a t K m [ S ] [ E ] T = k 1 k 3 k 5 k 2 k 4 + k 2 k 5 + k 3 k 5 [ E ] T
From Equation (8),
k c a t K m = k 1 k 3 k 5 k 2 k 4 + k 2 k 5 + k 3 k 5
Despite the simplicity of the kinetic model in Scheme 1, it remains challenging to determine from Equation (9) how each step in the kinetic model influences the kcat/Km parameter as compared to Equation (7) [129]. As a result, enzyme turnover is better defined under conditions of substrate saturation, where kcat is dependent on kinetic steps beyond the substrate binding step, i.e., the isomerization step, the chemical step of catalysis, or the product release step.

5. Strategies for Enhancing the Product Release Step

For the past few decades, approaches such as mutagenesis, nanobiocatalysis, modulation of loop dynamics, allosteric modulation, and engineering of transport tunnels have been used to enhance the product release step in enzyme catalysis.
A widely used technique, nanobiocatalysis, involves immobilizing enzymes on nanoparticles to improve enzyme catalytic efficiency, durability, and recyclability [132]. Nanoparticles have high surface areas and unique properties, which will enhance the turnover rate of enzymes via mechanisms including metal ion activation [133], electron transfer [134], morphology effects [134], reducing mass transfer limitations [135], conformation changes [136], increasing enzyme stability [137], and increasing substrate affinity [135].
Product release tunnels have been directly targeted by some studies through the modification of specific residues involved in product delivery or by the de novo synthesis of transport tunnels to improve product release [138,139,140]. Residue modification involves widening tunnels to influence product movement by altering the sizes of tunnel-lining residues or by adjusting the hydrophobicity/hydrophilicity content of such tunnels. For example, in a study on toluene-ortho-xylene monooxygenase, it was found that the size of the residue at position 314 was inversely proportional to enzyme activity [141]. The glutamine residue at this position was mutated to smaller residues, which resulted in an increased product efflux rate and a higher enzyme activity [141]. Conversely, de novo tunnel synthesis involves identifying potential locations for new tunnels through computational modeling and introducing mutations to create openings or channels in the protein structure. A notable example is a study on haloalkane dehalogenase, in which a transport tunnel was designed and tested computationally to show an increased product release rate and enhanced activity of the enzyme [138].
The effects of enzyme motion, such as loop dynamics, on catalysis have been a focus in protein engineering. Research on enhancing or reducing the flexibility of specific enzyme loop regions has led to increased product release and enzyme turnover rates [142,143]. Allosteric modulation is another approach used to enhance product release rates by using specific molecules that influence conformational dynamics, thereby increasing enzyme turnover rates [144,145,146].
In Table 4 below, we present a summary of diverse studies, briefly describing the experimental approach that the researchers used to increase the enzyme turnover, mainly by improving the rate of product release from the active site of enzymes. A significant increase in kcat, often by several folds, can be observed across all studies, indicating that accelerating product release rates is not only a theoretical approach but a practical way of significantly increasing enzyme turnover rates.

6. Strategies for Enhancing the Chemical Step of Catalysis

Enhancing the rate of the chemical step of catalysis is mostly the main focus in engineering enzymes, especially when it is rate-limiting in the enzyme-catalyzed reaction. This step records events from the enzyme–substrate complex to the enzyme–product complex, except for any isomerization steps following the enzyme–substrate complex. Computational approaches have been used in diverse studies to optimize enzymes by targeting certain active site properties, including catalytic motifs [165], as well as virtually simulating newly designed active sites for improved catalytic rates [166].
The unique roles of certain non-canonical amino acids in the active sites of enzymes have been investigated in some studies, where residues such as Nδ-methylhistidine have been shown to act as catalytic nucleophiles in certain hydrolytic enzymes, resulting in catalytic rate enhancements exceeding 9000-fold [167]. Similarly, canonical amino acids of specific properties have been substituted into enzyme active sites via site-directed mutagenesis to increase the kcat value by stabilizing the transition state, enhancing nucleophilicity, altering steric repulsion, modulating electronic effects, or changing protein dynamics [168].
Several studies that have used various strategies to improve the chemical step of catalysis (k3) and enzyme turnover rate are examined and recorded in Table 5. Table 5 below summarizes these studies, briefly highlighting the methodology and key findings obtained. These methods center on improving the bond-breaking and bond-forming reactions within the catalytic pathway, and in most cases, the overall turnover rate of the enzymes involved was either fully or partially rate-limiting by the chemical step of catalysis. A significant fold increase in kcat was observed across these studies when specific alterations affecting the catalytic step were performed.

7. The kcat/Km Parameter—Enzyme Catalytic Efficiency

kcat/Km measures how efficiently an enzyme converts substrates into products and is one of the parameters used to compare the level of preference an enzyme has for different substrates [182]. This steady-state kinetic parameter is often used to measure and compare the catalytic usefulness of engineered or mutant enzymes relative to wild-type or other mutants of the enzyme [11]. The parameter combines two important kinetic constants: kcat, which represents the turnover number, and Km, known as the Michaelis constant. From Equation (2) in the kcat section, when the substrate concentration is much less than the Km value, Equation (2) is simplified to Equation (10) below.
v = k c a t K m S [ E ] T
Thus, under the condition where [S] << Km, kcat/Km is a second-order rate constant with units of M−1s−1. Under this condition, can the physical meaning of kcat/Km as a rate constant be interpreted? Cleland demonstrated that kcat/Km represents the “net” rate constant for the substrate association step (E + S → ES) and that this parameter is determined by all the elementary rate constants up to and including the first irreversible step [60,129]. In order to report the kinetics of the production of the enzyme–substrate complex (ES) that “are destined to yield the product,” Northrop suggested using “kcap” rather than kcat/Km, as the latter is the apparent rate constant for the “capture” of the substrate [129]. The parameter kcat/Km should not be mistaken for merely the result of dividing kcat by Km, as students often incorrectly use it to evaluate catalytic efficiency and substrate affinity simultaneously, which is inaccurate in many cases [183]. kcat/Km is indicative of how effectively the free state of an enzyme captures the substrate for catalytic turnover. Thus, kcat/Km defines the rate of productive substrate binding, which governs specificity [123]. The Michaelis constant, Km, is the ratio of kcat to kcat/Km and corresponds to the concentration of the substrate at which the enzyme converts the substrate into product at half the maximum reaction rate. Km is a pseudo-dissociation constant and does not report on the chemical affinity of enzymes unless rigorously determined through advanced methods, which are beyond the scope of most enzyme kinetic analyses. Km may be considered a measure of substrate affinity only under specific conditions, which will be discussed later in this paper.

8. Strategies for Enhancing the Catalytic Efficiency of an Enzyme

The kcat/Km parameter is another focus of enzyme engineering and optimization, and it is particularly relevant under low substrate conditions where most of the enzyme is in its free form, which makes the reaction rate dependent on how efficiently the enzyme binds the substrate and initiates the catalytic process. Nanoparticle technology has been applied in enzyme engineering, where enzymes are immobilized on nanoparticle surfaces to influence substrate binding and catalysis. The catalytic efficiency of phosphotriesterase has been increased by nanoparticles, such as cadmium selenide (CdSe) and zinc sulfide (ZnS), through the alteration of the microenvironment surrounding the hydration layer of the enzyme–nanoparticle bioconjugate [148].
Understanding loop dynamics can be useful when optimizing enzymes because the conformationally dynamic nature of enzymes has been shown to be essential in regulating specificity and biological function. For instance, residues at the interface of the mobile loop of the NS2B-NS3pro complex, a two-component viral protease, were mutated, resulting in a 1.5-fold increase in the kcat/Km value [143]. Studies have incorporated non-canonical amino acids into the active site of enzymes to increase the kcat/Km value as a result of their unique properties, including transition state stabilization and enhanced nucleophilicity [171].
Selected studies, including their methodologies and key findings, that have used various strategies to enhance the catalytic efficiency (kcat/Km) of enzymes are summarized in Table 6. In these studies, the approaches used affected either the chemical step of catalysis, the product release step, or both, which resulted in a significant fold increase in the kcat/Km value. It is worth noting that some approaches increased the kcat/Km value despite a decrease in kcat, due to a more substantial reduction in Km relative to kcat.

9. Revisiting Misconceptions in Enzyme Design and Optimization

We define “misconception” as an incorrect or false theory, concept, principle, or claim resulting from inaccurate information, oversimplification, and/or incorrect inferences or interpretations of data. In this section, we highlight common fallacies made in enzyme kinetic studies and provide accurate information and formulas to refute them.
Misconception 1: Km is a measure of binding affinity and is equivalent to the dissociation constant Kd.
Refuting misconception: The relationship between Km and Kd is often misinterpreted. As Km is represented by the substrate concentration at half the maximum reaction velocity, substrate affinity is sometimes interpreted as Km. Km, however, is not a direct measure of affinity but rather is considered an indicator of how efficiently the substrate is bound and converted into product by the enzyme. In contrast, the equilibrium between the free and bound states of a substrate is quantified by Kd, which makes it the proper kinetic parameter for describing substrate affinity. Assuming Km is equivalent to Kd is incorrect. Although related to each other, Km is only equal to Kd under the following three conditions for an irreversible reaction: (1) when the substrate is non-sticky—the substrate quickly dissociates after binding the active site, (2) when the product is non-sticky—the product quickly dissociates after its formation, and (3) when the product release step is not rate-limiting. Consider Scheme 1, described earlier, which involves a reaction with a single substrate and product. The kcat and kcat/Km given in Equations (7) and (9) can be used to derive an expression for the Km parameter as follows:
K m = k c a t ( k c a t K m ) = k 3 k 5 k 3 +   k 4 +   k 5 . k 2 k 4 +   k 2 k 5 +   k 3 k 5   k 1 k 3 k 5
Equation (11) is simplified to Equation (12) below.
K m = k 2 k 4 + k 2 k 5 + k 3 k 5   k 1 ( k 3 + k 4 + k 5 )  
Condition 1: For a non-sticky substrate, where k2 >> k3, Equation (12) further simplifies as:
K m = k 2 k 4 +   k 5 ( k 2 +   k 3 )   k 1 k 3 +   k 1 k 4 +   k 1 k 5 =   k 2 k 4 +   k 2 k 5   k 1 k 3 +   k 1 k 4 +   k 1 k 5
Condition 2: For a non-sticky product, where k5 >> k4, Equation (13) further simplifies as:
K m = k 2 ( k 4 +   k 5 )   k 1 k 3 +   k 1 ( k 4 +   k 5 ) =   k 2 k 5   k 1 k 3 +   k 1 k 5
Condition 3: Given that the product release step is not rate-limiting where k5 >> k3, Km from Equation (14) then simplifies as:
K m = k 2 k 5   k 1 ( k 3 +   k 5 ) =   k 2 k 5   k 1 k 5
Therefore, under the three conditions mentioned above, Km is equivalent to Kd, as shown in Equation (16) below.
K m = k 2   k 1 = K d  
The conditions become even more complex when considering an enzyme-catalyzed reaction with two or more substrates, for which the analytical formulations for the kcat and kcat/Km parameters are considerably more complex than Equations (7) and (9).
Misconception 2: Increasing the rate of substrate binding increases enzyme turnover (kcat).
Refuting misconception: This claim may arise from studies where alterations or modifications that enhance substrate binding in the active site also led to increased enzyme turnover. The improved substrate binding might be correlated with enzyme turnover in such experiments. Nonetheless, it is important to underline that kcat is defined under saturating substrate conditions, more especially following the binding of the substrate to the enzyme [120]. When substrate levels are relatively low ([S] << Km), the rate of the reaction is determined by kcat/Km (Equation (8)), which presents a complex model that makes it challenging to determine how each step influences the reaction rate [128]. In contrast, under substrate-saturating conditions, Equation (5) provides a mechanistically feasible understanding of the impact of each kinetic step on enzyme turnover. Therefore, if kcat increases in the modified enzyme, the factor causing this increase likely affects the rate-limiting step(s), such as an isomerization step, the chemical step of catalysis, or the product release step, rather than substrate binding itself (Equation (7) derived from Equation (4) under saturating substrate conditions). The enzyme turnover number is dependent on all reaction steps, excluding the substrate association and dissociation steps.
Misconception 3: An alteration that affects the kcat value implies that the chemical step of catalysis is affected.
Refuting misconception: This statement is often made in studies that seek to optimize enzyme activity by causing an alteration, which may be in the form of a mutation of specific residues in the enzyme active site, and then measuring kinetic parameters, such as kcat, Km, and kcat/Km. An alteration that affects enzyme turnover does not necessarily imply that catalysis is affected. We established earlier from Equation (7) that kcat is a composite rate constant that can be influenced by either the isomerization step, the chemical step of catalysis, or the product release step, whichever is rate-limiting or a combination thereof. Without detailed mechanistic studies, it is impossible to attribute changes in enzyme turnover solely to the chemical step of catalysis. An alteration that speeds up the chemical step of catalysis may not affect overall turnover if product release is already the rate-limiting step. Conversely, an alteration that affects product release could change turnover without altering the chemical catalytic step if product release is rate-limiting.

10. Conclusions

Mechanistic enzymology is a strong foundation for understanding the molecular mechanisms of the catalytic activity of enzymes, which is essential for engineering and optimizing enzymes in industrial biocatalysis (Figure 8). Enzyme mechanisms can be elucidated by combining techniques such as X-ray crystallography, NMR spectroscopy, transient-state kinetics, kinetic isotope effects, site-directed mutagenesis, and kinetic solvent viscosity effects [48,62,81]. The information obtained from enzyme mechanisms is considered useful for improving industrial yield and productivity, as well as for other applications, including the design of new enzymes for food processing, drug development, and the production of biofuels, detergents, and pulp and paper [190]. The solvent systems of enzymatic reactions are critical to the stability, activity, and selectivity of enzymes. Deep eutectic solvents (DESs) have emerged as alternatives to most solvent systems because they are nontoxic, nonvolatile, nonflammable, biodegradable, and considered “green” solvents [90]. The state of the enzyme is also vital, as studies have shown that immobilized enzymes tend to exhibit greater stability, retain much of their activity, and are especially useful for various industrial applications, such as waste disposal, where enzymes can be reused.
The technique of kinetic solvent viscosity effects (KSVEs) is regarded as a highly efficient and cost-effective method that can be applied to simple enzyme kinetic assays, such as the Michaelis–Menten kinetic approach, without the need for advanced instruments and expertise. In KSVEs, microviscosigens, such as glucose and sucrose, can be used to increase the frictional resistance experienced by enzymes and substrates in solution, whereas macroviscogens, such as Ficoll-400 and polyethylene glycol, can be used to set up controls to distinguish whether the observed viscosity effect arises from diffusion limitations rather than macromolecular crowding. KSVEs can be used to probe different aspects of enzyme reaction mechanisms, including diffusion-controlled substrate binding and rate-limiting steps, which may involve one or a combination of isomerization steps, the product release step, or the chemical step of catalysis [48] Determining the effect of diffusion-dependent steps is important in biocatalysis, as it can guide targeted approaches in enzyme optimization.
We have demonstrated in this review that, beyond optimizing enzymes through approaches such as rational design and directed evolution, the integration of mechanistic enzymology techniques, particularly those involving KSVEs, enables the investigation of diffusion-dependent kinetic steps and the precise identification of rate-limiting steps. Knowledge of rate-limiting steps in an enzyme-catalyzed reaction can be used to guide targeted approaches in enzyme design and optimization by prioritizing such steps in the pathway. Moreover, challenges such as limited control over tradeoffs, difficulty in predicting enzyme behavior under different conditions, and issues with industrial scalability can all be addressed through mechanistic enzymology.

Author Contributions

Conceptualization, G.G. and J.A.Q.; investigation, G.A.A.; writing—original draft preparation, G.A.A.; writing—review and editing, G.G. and J.A.Q.; supervision, G.G. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

All data are contained within the manuscript.

Acknowledgments

We acknowledge the support of the Georgia State University Departments of Chemistry and Biology, and the Center for Diagnostics and Therapeutics. We appreciate members of the Gadda lab, as well as all conference attendees whose comments were insightful for this review. We are also thankful to the reviewers for their insightful suggestions.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Hughes, G.; Lewis, J.C. Introduction: Biocatalysis in Industry. Chem. Rev. 2018, 118, 1–3. [Google Scholar] [CrossRef]
  2. Cooper, G.M. The Central Role of Enzymes as Biological Catalysts. Available online: https://www.ncbi.nlm.nih.gov/books/NBK9921/ (accessed on 23 December 2024).
  3. Agarwal, P.K. A Biophysical Perspective on Enzyme Catalysis. Biochemistry 2018, 58, 438. [Google Scholar] [CrossRef]
  4. Fernandes, P. Enzymes in Food Processing: A Condensed Overview on Strategies for Better Biocatalysts. Enzym. Res. 2010, 2010, 862537. [Google Scholar] [CrossRef]
  5. Poulsen: History of Enzymology with Emphasis on Food. Available online: https://scholar.google.com/scholar_lookup?title=Handbook%20of%20Food%20Enzymology&author=PB%20Poulsen&author=H%20Klaus%20Buchholz&publication_year=2003& (accessed on 27 May 2025).
  6. Wandrey, C.; Liese, A.; Kihumbu, D. Industrial Biocatalysis: Past, Present, and Future. Org. Process Res. Dev. 2000, 4, 286–290. [Google Scholar] [CrossRef]
  7. Rosche, B.; Li, X.Z.; Hauer, B.; Schmid, A.; Buehler, K. Microbial Biofilms: A Concept for Industrial Catalysis? Trends Biotechnol 2009, 27, 636–643. [Google Scholar] [CrossRef] [PubMed]
  8. Jiang, W.; Fang, B. Synthesizing Chiral Drug Intermediates by Biocatalysis. Appl. Biochem. Biotechnol. 2020, 192, 146–179. [Google Scholar] [CrossRef]
  9. Patel, R.N. Synthesis of Chiral Pharmaceutical Intermediates by Biocatalysis. Coord. Chem. Rev. 2008, 252, 659–701. [Google Scholar] [CrossRef]
  10. Ndochinwa, O.G.; Wang, Q.Y.; Amadi, O.C.; Nwagu, T.N.; Nnamchi, C.I.; Okeke, E.S.; Moneke, A.N. Current Status and Emerging Frontiers in Enzyme Engineering: An Industrial Perspective. Heliyon 2024, 10, e32673. [Google Scholar] [CrossRef]
  11. Edmondson, D.E.; Gadda, G. Guidelines for the Functional Analysis of Engineered and Mutant Enzymes. Protein Eng. Handb. 2011, 1, 1–13. [Google Scholar] [CrossRef]
  12. Mao, S.; Jiang, J.; Xiong, K.; Chen, Y.; Yao, Y.; Liu, L.; Liu, H.; Li, X.; Mao, S.; Jiang, J.; et al. Enzyme Engineering: Performance Optimization, Novel Sources, and Applications in the Food Industry. Foods 2024, 13, 3846. [Google Scholar] [CrossRef]
  13. Pala, U.U.; Yelmazer, B.; ÇorbacloÄY Lu, M.; Ruupunen, J.; Valjakka, J.; Turunen, O.; Binay, B.Y. Functional Effects of Active Site Mutations in NAD+-Dependent Formate Dehydrogenases on Transformation of Hydrogen Carbonate to Formate. Protein Eng. Des. Sel. 2018, 31, 327–335. [Google Scholar] [CrossRef]
  14. Phintha, A.; Chaiyen, P. Rational and Mechanistic Approaches for Improving Biocatalyst Performance. Chem Catal. 2022, 2, 2614–2643. [Google Scholar] [CrossRef]
  15. Han, R.; Li, J.; Shin, H.D.; Chen, R.R.; Du, G.; Liu, L.; Chen, J. Recent Advances in Discovery, Heterologous Expression, and Molecular Engineering of Cyclodextrin Glycosyltransferase for Versatile Applications. Biotechnol. Adv. 2014, 32, 415–428. [Google Scholar] [CrossRef]
  16. Mascarenhas, N.M.; Gosavi, S. Understanding Protein Domain-Swapping Using Structure-Based Models of Protein Folding. Prog. Biophys. Mol. Biol. 2016, 128, 113. [Google Scholar] [CrossRef] [PubMed]
  17. Merz, A.; Yee, M.C.; Szadkowski, H.; Pappenberger, G.; Crameri, A.; Stemmer, W.P.C.; Yanofsky, C.; Kirschner, K. Improving the Catalytic Activity of a Thermophilic Enzyme at Low Temperatures. Biochemistry 2000, 39, 880–889. [Google Scholar] [CrossRef]
  18. Joho, Y.; Vongsouthi, V.; Gomez, C.; Larsen, J.S.; Ardevol, A.; Jackson, C.J. Improving Plastic Degrading Enzymes via Directed Evolution. Protein Eng. Des. Sel. 2024, 37, 9. [Google Scholar] [CrossRef]
  19. Gargiulo, S.; Soumillion, P. Directed Evolution for Enzyme Development in Biocatalysis. Curr. Opin. Chem. Biol. 2021, 61, 107–113. [Google Scholar] [CrossRef]
  20. Arnold, F.H. Innovation by Evolution: Bringing New Chemistry to Life (Nobel Lecture). Angew. Chem. Int. Ed. Engl. 2019, 58, 14420–14426. [Google Scholar] [CrossRef] [PubMed]
  21. Wang, Y.; Xue, P.; Cao, M.; Yu, T.; Lane, S.T.; Zhao, H. Directed Evolution: Methodologies and Applications. Chem. Rev. 2021, 121, 12384–12444. [Google Scholar] [CrossRef]
  22. Ravikumar, A.; Arzumanyan, G.A.; Obadi, M.K.A.; Javanpour, A.A.; Liu, C.C. Scalable, Continuous Evolution of Genes at Mutation Rates above Genomic Error Thresholds. Cell 2018, 175, 1946–1957. [Google Scholar] [CrossRef]
  23. Khakzad, H.; Igashov, I.; Schneuing, A.; Goverde, C.; Bronstein, M.; Correia, B. A New Age in Protein Design Empowered by Deep Learning. Cell Syst. 2023, 14, 925–939. [Google Scholar] [CrossRef]
  24. Dauparas, J.; Anishchenko, I.; Bennett, N.; Bai, H.; Ragotte, R.J.; Milles, L.F.; Wicky, B.I.M.; Courbet, A.; de Haas, R.J.; Bethel, N.; et al. Robust Deep Learning–Based Protein Sequence Design Using ProteinMPNN. Science 2022, 378, 49–56. [Google Scholar] [CrossRef]
  25. Jumper, J.; Evans, R.; Pritzel, A.; Green, T.; Figurnov, M.; Ronneberger, O.; Tunyasuvunakool, K.; Bates, R.; Žídek, A.; Potapenko, A.; et al. Highly Accurate Protein Structure Prediction with AlphaFold. Nature 2021, 596, 583–589. [Google Scholar] [CrossRef] [PubMed]
  26. Gumulya, Y.; Gillam, E.M.J. Exploring the Past and the Future of Protein Evolution with Ancestral Sequence Reconstruction: The ‘Retro’ Approach to Protein Engineering. Biochem. J. 2017, 474, 1–19. [Google Scholar] [CrossRef] [PubMed]
  27. Hochberg, G.K.A.; Thornton, J.W. Reconstructing Ancient Proteins to Understand the Causes of Structure and Function. Annu. Rev. Biophys. 2017, 46, 247. [Google Scholar] [CrossRef]
  28. Genetic Engineering. Available online: https://www.genome.gov/genetics-glossary/Genetic-Engineering (accessed on 21 July 2025).
  29. Knott, G.J.; Doudna, J.A. CRISPR-Cas Guides the Future of Genetic Engineering. Science 2018, 361, 866. [Google Scholar] [CrossRef]
  30. Dong, Q.; Wang, J.; Yang, H.; Wei, C.; Yu, J.; Zhang, Y.; Huang, Y.; Zhang, X.E.; Wei, H. Construction of a Chimeric Lysin Ply187N-V12C with Extended Lytic Activity against Staphylococci and Streptococci. Microb. Biotechnol. 2015, 8, 210–220. [Google Scholar] [CrossRef]
  31. Balabanova, L.; Golotin, V.; Kovalchuk, S.; Bulgakov, A.; Likhatskaya, G.; Son, O.; Rasskazov, V. A Novel Bifunctional Hybrid with Marine Bacterium Alkaline Phosphatase and Far Eastern Holothurian Mannan-Binding Lectin Activities. PLoS ONE 2014, 9, e112729. [Google Scholar] [CrossRef]
  32. Gheybi, E.; Amani, J.; Salmanian, A.H.; Mashayekhi, F.; Khodi, S. Designing a Recombinant Chimeric Construct Contain MUC1 and HER2 Extracellular Domain for Prediagnostic Breast Cancer. Tumor Biol. 2014, 35, 11489–11497. [Google Scholar] [CrossRef] [PubMed]
  33. Otsuka, R.; Imai, S.; Murata, T.; Nomura, Y.; Okamoto, M.; Tsumori, H.; Kakuta, E.; Hanada, N.; Momoi, Y. Application of Chimeric Glucanase Comprising Mutanase and Dextranase for Prevention of Dental Biofilm Formation. Microbiol. Immunol. 2015, 59, 28–36. [Google Scholar] [CrossRef]
  34. Reddy, V.A.; Venu, K.; Rao, D.E.C.S.; Rao, K.V.; Reddy, V.D. Chimeric Gene Construct Coding for Bi-Functional Enzyme Endowed with Endoglucanase and Phytase Activities. Arch. Microbiol. 2009, 191, 171–175. [Google Scholar] [CrossRef]
  35. Nishioka, T.; Yasutake, Y.; Nishiya, Y.; Tamura, T. Structure-Guided Mutagenesis for the Improvement of Substrate Specificity of Bacillus Megaterium Glucose 1-Dehydrogenase IV. FEBS J. 2012, 279, 3264–3275. [Google Scholar] [CrossRef] [PubMed]
  36. Gao, X.; Cui, W.; Ding, N.; Liu, Z.; Tian, Y.; Zhou, Z. Structure-Based Approach to Alter the Substrate Specificity of Bacillus Subtilis Aminopeptidase. Prion 2013, 7, 328–334. [Google Scholar] [CrossRef]
  37. Webster, C.I.; Burrell, M.; Olsson, L.L.; Fowler, S.B.; Digby, S.; Sandercock, A.; Snijder, A.; Tebbe, J.; Haupts, U.; Grudzinska, J.; et al. Engineering Neprilysin Activity and Specificity to Create a Novel Therapeutic for Alzheimer’s Disease. PLoS ONE 2014, 9, e104001. [Google Scholar] [CrossRef]
  38. Umekawa, M.; Li, C.; Higashiyama, T.; Huang, W.; Ashida, H.; Yamamoto, K.; Wang, L.X. Efficient Glycosynthase Mutant Derived from Mucor Hiemalis Endo-β-N-Acetylglucosaminidase Capable of Transferring Oligosaccharide from Both Sugar Oxazoline and Natural N-Glycan. J. Biol. Chem. 2010, 285, 511–521. [Google Scholar] [CrossRef]
  39. Hediger, M.R.; De Vico, L.; Rannes, J.B.; Jäckel, C.; Besenmatter, W.; Svendsen, A.; Jensen, J.H. In Silico Screening of 393 Mutants Facilitates Enzyme Engineering of Amidase Activity in CalB. PeerJ 2013, 2013, e145. [Google Scholar] [CrossRef]
  40. Johnston, C.A.; Whitneye, D.S.; Volkmane, B.F.; Doe, C.Q.; Prehoda, K.E. Conversion of the Enzyme Guanylate Kinase into a Mitotic-Spindle Orienting Protein by a Single Mutation That Inhibits GMP-Induced Closing. Proc. Natl. Acad. Sci. USA 2011, 108, E973–E978. [Google Scholar] [CrossRef]
  41. Drennan, C.L.; Jarrett, J.T. From Single Molecules to Whole Organisms: The Evolving Field of Mechanistic Enzymology. Curr. Opin. Chem. Biol. 2009, 13, 433. [Google Scholar] [CrossRef] [PubMed]
  42. Holdgate, G.A.; Meek, T.D.; Grimley, R.L. Mechanistic Enzymology in Drug Discovery: A Fresh Perspective. Nat. Rev. Drug Discov. 2018, 17, 115–132. [Google Scholar] [CrossRef]
  43. Ding, Y.; Perez-Ortiz, G.; Peate, J.; Barry, S.M. Redesigning Enzymes for Biocatalysis: Exploiting Structural Understanding for Improved Selectivity. Front. Mol. Biosci. 2022, 9, 908285. [Google Scholar] [CrossRef]
  44. Winkler, C.K.; Schrittwieser, J.H.; Kroutil, W. Power of Biocatalysis for Organic Synthesis. ACS Cent. Sci. 2021, 7, 55–71. [Google Scholar] [CrossRef] [PubMed]
  45. Pyser, J.B.; Chakrabarty, S.; Romero, E.O.; Narayan, A.R.H. State-of-the-Art Biocatalysis. ACS Cent. Sci. 2021, 7, 1105–1116. [Google Scholar] [CrossRef] [PubMed]
  46. Bruice, T.C.; Benkovic, S.J. Chemical Basis for Enzyme Catalysis. Biochemistry 2000, 39, 6267–6274. [Google Scholar] [CrossRef] [PubMed]
  47. Warshel, A.; Sharma, P.K.; Kato, M.; Xiang, Y.; Liu, H.; Olsson, M.H.M. Electrostatic Basis for Enzyme Catalysis. Chem. Rev. 2006, 106, 3210–3235. [Google Scholar] [CrossRef]
  48. Gadda, G.; Sobrado, P. Kinetic Solvent Viscosity Effects as Probes for Studying the Mechanisms of Enzyme Action. Biochemistry 2018, 57, 3445–3453. [Google Scholar] [CrossRef]
  49. Robinson, R.; Franceschini, S.; Fedkenheuer, M.; Rodriguez, P.J.; Ellerbrock, J.; Romero, E.; Echandi, M.P.; Martin Del Campo, J.S.; Sobrado, P. Arg279 Is the Key Regulator of Coenzyme Selectivity in the Flavin-Dependent Ornithine Monooxygenase SidA. Biochim. Et Biophys. Acta (BBA)—Proteins Proteom. 2014, 1844, 778–784. [Google Scholar] [CrossRef]
  50. Da Fonseca, I.; Qureshi, I.A.; Mehra-Chaudhary, R.; Kizjakina, K.; Tanner, J.J.; Sobrado, P. Contributions of Unique Active Site Residues of Eukaryotic UDP-Galactopyranose Mutases to Substrate Recognition and Active Site Dynamics. Biochemistry 2014, 53, 7794. [Google Scholar] [CrossRef]
  51. Moni, B.M.; Quaye, J.A.; Gadda, G. Mutation of a Distal Gating Residue Modulates NADH Binding in NADH:Quinone Oxidoreductase from Pseudomonas Aeruginosa PAO1. J. Biol. Chem. 2023, 299, 103044. [Google Scholar] [CrossRef]
  52. Wagner, C.R.; Benkovic, S.J. Site Directed Mutagenesis: A Tool for Enzyme Mechanism Dissection. Trends Biotechnol. 1990, 8, 263–270. [Google Scholar] [CrossRef]
  53. Sunden, F.; Peck, A.; Salzman, J.; Ressl, S.; Herschlag, D. Extensive Site-Directed Mutagenesis Reveals Interconnected Functional Units in the Alkaline Phosphatase Active Site. Elife 2015, 4, e06181. [Google Scholar] [CrossRef] [PubMed]
  54. Kramers, H.A. Brownian Motion in a Field of Force and the Diffusion Model of Chemical Reactions. Physica 1940, 7, 284–304. [Google Scholar] [CrossRef]
  55. Quaye, J.A.; Gadda, G. The Pseudomonas Aeruginosa PAO1 Metallo Flavoprotein D-2-Hydroxyglutarate Dehydrogenase Requires Zn2+ for Substrate Orientation and Activation. J. Biol. Chem. 2023, 299, 103008. [Google Scholar] [CrossRef]
  56. Quaye, J.A.; Ball, J.; Gadda, G. Kinetic Solvent Viscosity Effects Uncover an Internal Isomerization of the Enzyme-Substrate Complex in Pseudomonas Aeruginosa PAO1 NADH:Quinone Oxidoreductase. Arch. Biochem. Biophys. 2022, 727, 109342. [Google Scholar] [CrossRef]
  57. Vodovoz, M.; Gadda, G. Kinetic Solvent Viscosity Effects Reveal a Protein Isomerization in the Reductive Half-Reaction of Neurospora Crassa Class II Nitronate Monooxygenase. Arch. Biochem. Biophys. 2020, 695, 108625. [Google Scholar] [CrossRef]
  58. Achuthan, S.; Chung, B.J.; Ghosh, P.; Rangachari, V.; Vaidya, A. A Modified Stokes-Einstein Equation for Aβ Aggregation. BMC Bioinform. 2011, 12, S13. [Google Scholar] [CrossRef]
  59. Uribe, S.; Sampedro, J.G. Measuring Solution Viscosity and Its Effect on Enzyme Activity. Biol. Proced. Online 2003, 5, 108. [Google Scholar] [CrossRef]
  60. Cleland, W.W. Partition Analysis and the Concept of Net Rate Constants as Tools in Enzyme Kinetics. Biochemistry 1975, 14, 3220–3224. [Google Scholar] [CrossRef]
  61. Fitzpatrick, P.F.; Kurtz, K.A.; Denu, J.M.; Emanuele, J.F. Contrasting Values of Commitment Factors Measured from Viscosity, PH, and Kinetic Isotope Effects: Evidence for Slow Conformational Changes in TheD-Amino Acid Oxidase Reaction. Bioorg. Chem. 1997, 25, 100–109. [Google Scholar] [CrossRef]
  62. Hull, A.W. A New Method of X-Ray Crystal Analysis. Phys. Rev. 1917, 10, 661. [Google Scholar] [CrossRef]
  63. Kovaleva, E.G.; Lipscomb, J.D. Crystal Structures of Fe2+ Dioxygenase Superoxo, Alkylperoxo, and Bound Product Intermediates. Science 2007, 316, 453–457. [Google Scholar] [CrossRef]
  64. Mehra-Chaudhary, R.; Dai, Y.; Sobrado, P.; Tanner, J.J. In Crystallo Capture of a Covalent Intermediate in the UDP-Galactopyranose Mutase Reaction. Biochemistry 2016, 55, 833–836. [Google Scholar] [CrossRef]
  65. Weis, D.D. Hydrogen Exchange Mass Spectrometry of Proteins; Wiley: Hoboken, NJ, USA, 2016. [Google Scholar]
  66. Masson, G.R.; Burke, J.E.; Ahn, N.G.; Anand, G.S.; Borchers, C.; Brier, S.; Bou-Assaf, G.M.; Engen, J.R.; Englander, S.W.; Faber, J.; et al. Recommendations for Performing, Interpreting and Reporting Hydrogen Deuterium Exchange Mass Spectrometry (HDX-MS) Experiments. Nat. Methods 2019, 16, 595–602. [Google Scholar] [CrossRef]
  67. Hodge, E.A.; Benhaim, M.A.; Lee, K.K. Bridging Protein Structure, Dynamics, and Function Using Hydrogen/Deuterium-Exchange Mass Spectrometry. Protein Sci. 2020, 29, 843–855. [Google Scholar] [CrossRef]
  68. Harrison, R.A.; Engen, J.R. Conformational Insight into Multi-Protein Signaling Assemblies by Hydrogen–Deuterium Exchange Mass Spectrometry. Curr. Opin. Struct. Biol. 2016, 41, 187–193. [Google Scholar] [CrossRef] [PubMed]
  69. Vadas, O.; Jenkins, M.L.; Dornan, G.L.; Burke, J.E. Using Hydrogen–Deuterium Exchange Mass Spectrometry to Examine Protein–Membrane Interactions. Methods Enzym. 2017, 583, 143–172. [Google Scholar] [CrossRef]
  70. Englander, S.W.; Mayne, L. The Nature of Protein Folding Pathways. Proc. Natl. Acad. Sci. USA 2014, 111, 15873–15880. [Google Scholar] [CrossRef]
  71. Jiang, Y.; Kalodimos, C.G. NMR Studies of Large Proteins. J. Mol. Biol. 2017, 429, 2667–2676. [Google Scholar] [CrossRef] [PubMed]
  72. Rosenzweig, R.; Kay, L.E. Bringing Dynamic Molecular Machines into Focus by Methyl-TROSY NMR. Annu. Rev. Biochem. 2014, 83, 291–315. [Google Scholar] [CrossRef]
  73. Wang, Y.; Bugge, K.; Kragelund, B.B.; Lindorff-Larsen, K. Role of Protein Dynamics in Transmembrane Receptor Signalling. Curr. Opin. Struct. Biol. 2018, 48, 74–82. [Google Scholar] [CrossRef] [PubMed]
  74. Schütz, A.K. Solid-State NMR Approaches to Investigate Large Enzymes in Complex with Substrates and Inhibitors. Biochem. Soc. Trans. 2021, 49, 131–144. [Google Scholar] [CrossRef]
  75. Mackenzie, N.E.; Paul, J.; Malthouse, G.; Scott, A.I. Studying Enzyme Mechanism by 13C Nuclear Magnetic Resonance. Science 1984, 225, 883–889. [Google Scholar] [CrossRef]
  76. Johnson, K.A. Advances in Transient-State Kinetics. Curr. Opin. Biotechnol. 1998, 9, 87–89. [Google Scholar] [CrossRef] [PubMed]
  77. Anderson, K.S.; Sammons, R.D.; Leo, G.C.; Sikorski, J.A.; Benesi, A.J.; Johnson, K.A. Observation by 13C NMR of the EPSP Synthase Tetrahedral Intermediate Bound to the Enzyme Active Site. Biochemistry 1990, 29, 1460–1465. [Google Scholar] [CrossRef] [PubMed]
  78. Anderson, K.S.; Johnson, K.A. “Kinetic Competence” of the 5-Enolpyruvoylshikimate-3-Phosphate Synthase Tetrahedral Intermediate*. J. Biol. Chem. 1990, 265, 5567–5572. [Google Scholar] [CrossRef]
  79. Appleyard, R. Direct Observation of the Enzyme-Intermediate Complex of 5-Enolpyruvylshikimate-3-Phosphate Synthase by Carbon-13 NMR Spectroscopy. Biochemistry 1989, 28, 7985–7991. [Google Scholar]
  80. Studelska, D.R.; McDowell, L.M.; Espe, M.P.; Klug, C.A.; Schaefer, J. Slowed Enzymatic Turnover Allows Characterization of Intermediates by Solid-State NMR. Biochemistry 1997, 36, 15555–15560. [Google Scholar] [CrossRef] [PubMed]
  81. Johnson, K.A. Transient State Enzyme Kinetics. In Wiley Encyclopedia of Chemical Biology; John Wiley & Sons: Hoboken, NJ, USA, 2008; pp. 1–8. [Google Scholar] [CrossRef]
  82. Fisher, H.F. Transient-State Kinetic Approach to Mechanisms of Enzymatic Catalysis. Acc. Chem. Res. 2005, 38, 157–166. [Google Scholar] [CrossRef]
  83. Mao, Z.; Campbell, C.T. Kinetic Isotope Effects: Interpretation and Prediction Using Degrees of Rate Control. ACS Catal. 2020, 10, 4181–4192. [Google Scholar] [CrossRef]
  84. Kunkes, E.L.; Studt, F.; Abild-Pedersen, F.; Schlögl, R.; Behrens, M. Hydrogenation of CO2 to Methanol and CO on Cu/ZnO/Al2O3: Is There a Common Intermediate or Not? J. Catal. 2015, 328, 43–48. [Google Scholar] [CrossRef]
  85. Wei, J.; Iglesia, E. Isotopic and Kinetic Assessment of the Mechanism of Reactions of CH4 with CO2 or H2O to Form Synthesis Gas and Carbon on Nickel Catalysts. J. Catal. 2004, 224, 370–383. [Google Scholar] [CrossRef]
  86. Ojeda, M.; Li, A.; Nabar, R.; Nilekar, A.U.; Mavrikakis, M.; Iglesia, E. Kinetically Relevant Steps and H2/D2 Isotope Effects in Fischer-Tropsch Synthesis on Fe and Co Catalysts. J. Phys. Chem. C 2010, 114, 19761–19770. [Google Scholar] [CrossRef]
  87. Cleland, W.W. The Use of Isotope Effects to Determine Enzyme Mechanisms. Arch. Biochem. Biophys. 2005, 433, 2–12. [Google Scholar] [CrossRef]
  88. Salvi, F.; Rodriguez, I.; Hamelberg, D.; Gadda, G. Role of F357 as an Oxygen Gate in the Oxidative Half-Reaction of Choline Oxidase. Biochemistry 2016, 55, 1473–1484. [Google Scholar] [CrossRef]
  89. Stepankova, V.; Bidmanova, S.; Koudelakova, T.; Prokop, Z.; Chaloupkova, R.; Damborsky, J. Strategies for Stabilization of Enzymes in Organic Solvents. ACS Catal. 2013, 3, 2823–2836. [Google Scholar] [CrossRef]
  90. Pätzold, M.; Siebenhaller, S.; Kara, S.; Liese, A.; Syldatk, C.; Holtmann, D. Deep Eutectic Solvents as Efficient Solvents in Biocatalysis. Trends Biotechnol. 2019, 37, 943–959. [Google Scholar] [CrossRef]
  91. Carrea, G.; Riva, S. Properties and Synthetic Applications of Enzymes in Organic Solvents. Angew. Chem. Int. Ed. 2000, 39, 2226–2254. [Google Scholar] [CrossRef]
  92. Fox, D.M.; Gilman, J.W.; Morgan, A.B.; Shields, J.R.; Maupin, P.H.; Lyon, R.E.; De Long, H.C.; Trulove, P.C. Flammability and Thermal Analysis Characterization of Imidazolium-Based Ionic Liquids. Ind. Eng. Chem. Res. 2008, 47, 6327–6332. [Google Scholar] [CrossRef]
  93. Emel’Yanenko, V.N.; Boeck, G.; Verevkin, S.P.; Ludwig, R. Volatile Times for the Very First Ionic Liquid: Understanding the Vapor Pressures and Enthalpies of Vaporization of Ethylammonium Nitrate. Chem.–A Eur. J. 2014, 20, 11640–11645. [Google Scholar] [CrossRef]
  94. Maton, C.; De Vos, N.; Stevens, C.V. Ionic liquid thermal stabilities: Decomposition mechanisms and analysis tools. Chem. Soc. Rev. 2013, 42, 5963–5977. [Google Scholar] [CrossRef] [PubMed]
  95. Meine, N.; Benedito, F.; Chemistry, R.R.-G. Thermal stability of ionic liquids assessed by potentiometric titration. Green Chem. 2010, 12, 1711–1714. [Google Scholar] [CrossRef]
  96. Chícharo, B.; Cuesta, M.; Trapasso, G.; Porcar, R.; Martín, N.; Villa, R.; Lozano, P.; Fadlallah, S.; Allais, F.; García-Verdugo, E.; et al. Chemical and Enzymatic Synthetic Routes to the Diglycidyl Ester of 2,5-Furandicarboxylic Acid. Sustain. Chem. Pharm. 2025, 46, 102101. [Google Scholar] [CrossRef]
  97. Abbott, A.P.; Capper, G.; Davies, D.L.; Rasheed, R.K.; Tambyrajah, V. Novel Solvent Properties of Choline Chloride/Urea Mixtures. Chem. Commun. 2003, 70–71. [Google Scholar] [CrossRef]
  98. Wysokowski, M.; Luu, R.K.; Arevalo, S.; Khare, E.; Stachowiak, W.; Niemczak, M.; Jesionowski, T.; Buehler, M.J. Untapped Potential of Deep Eutectic Solvents for the Synthesis of Bioinspired Inorganic–Organic Materials. Chem. Mater. 2023, 35, 7878. [Google Scholar] [CrossRef] [PubMed]
  99. Smith, E.L.; Abbott, A.P.; Ryder, K.S. Deep Eutectic Solvents (DESs) and Their Applications. Chem. Rev. 2014, 114, 11060–11082. [Google Scholar] [CrossRef] [PubMed]
  100. Hansen, B.B.; Spittle, S.; Chen, B.; Poe, D.; Zhang, Y.; Klein, J.M.; Horton, A.; Adhikari, L.; Zelovich, T.; Doherty, B.W.; et al. Deep Eutectic Solvents: A Review of Fundamentals and Applications. Chem. Rev. 2021, 121, 1232–1285. [Google Scholar] [CrossRef]
  101. Liu, Y.; Li, L.; Ma, C.; He, Y.C. Chemobiocatalytic Transfromation of Biomass into Furfurylamine with Mixed Amine Donor in an Eco-Friendly Medium. Bioresour. Technol. 2023, 387, 129638. [Google Scholar] [CrossRef]
  102. Peito, S.; Peixoto, D.; Ferreira-Faria, I.; Margarida Martins, A.; Margarida Ribeiro, H.; Veiga, F.; Marto, J.; Cláudia Paiva-Santos, A. Nano- and Microparticle-Stabilized Pickering Emulsions Designed for Topical Therapeutics and Cosmetic Applications. Int. J. Pharm. 2022, 615, 121455. [Google Scholar] [CrossRef]
  103. Ali, D.C.; Pan, T.; Wu, Q.; Wang, Z. Oil–Water Interfaces of Pickering Emulsions: Microhabitats for Living Cell Biocatalysis. Trends Biotechnol. 2025, 43, 790–801. [Google Scholar] [CrossRef]
  104. Omar, K.A.; Sadeghi, R. Database of Deep Eutectic Solvents and Their Physical Properties: A Review. J. Mol. Liq. 2023, 384, 121899. [Google Scholar] [CrossRef]
  105. Ruß, C.; König, B. Low Melting Mixtures in Organic Synthesis—An Alternative to Ionic Liquids? Green Chem. 2012, 14, 2969–2982. [Google Scholar] [CrossRef]
  106. Sarmad, S.; Xie, Y.; Mikkola, J.P.; Ji, X. Screening of Deep Eutectic Solvents (DESs) as Green CO2 Sorbents: From Solubility to Viscosity. New J. Chem. 2017, 41, 290–301. [Google Scholar] [CrossRef]
  107. Shahbaz, K.; Mjalli, F.S.; Hashim, M.A.; AlNashef, I.M. Prediction of the Surface Tension of Deep Eutectic Solvents. Fluid Phase Equilib. 2012, 319, 48–54. [Google Scholar] [CrossRef]
  108. Bagh, F.S.G.; Shahbaz, K.; Mjalli, F.S.; AlNashef, I.M.; Hashim, M.A. Electrical Conductivity of Ammonium and Phosphonium Based Deep Eutectic Solvents: Measurements and Artificial Intelligence-Based Prediction. Fluid Phase Equilib. 2013, 356, 30–37. [Google Scholar] [CrossRef]
  109. Mjalli, F.S.; Naser, J.; Jibril, B.; Alizadeh, V.; Gano, Z. Tetrabutylammonium Chloride Based Ionic Liquid Analogues and Their Physical Properties. J. Chem. Eng. Data 2014, 59, 2242–2251. [Google Scholar] [CrossRef]
  110. Naser, J.; Mjalli, F.S.; Gano, Z.S. Molar Heat Capacity of Selected Type III Deep Eutectic Solvents. J. Chem. Eng. Data 2016, 61, 1608–1615. [Google Scholar] [CrossRef]
  111. Mjalli, F.S.; Naser, J. Viscosity Model for Choline Chloride-Based Deep Eutectic Solvents. Asia-Pac. J. Chem. Eng. 2015, 10, 273–281. [Google Scholar] [CrossRef]
  112. Jibril, B.; Mjalli, F.; Naser, J.; Gano, Z. New Tetrapropylammonium Bromide-Based Deep Eutectic Solvents: Synthesis and Characterizations. J. Mol. Liq. 2014, 199, 462–469. [Google Scholar] [CrossRef]
  113. Chen, Z.; Ludwig, M.; Warr, G.G.; Atkin, R. Effect of Cation Alkyl Chain Length on Surface Forces and Physical Properties in Deep Eutectic Solvents. J. Colloid Interface Sci. 2017, 494, 373–379. [Google Scholar] [CrossRef]
  114. Zdarta, J.; Jesionowski, T.; Pinelo, M.; Meyer, A.S.; Iqbal, H.M.N.; Bilal, M.; Nguyen, L.N.; Nghiem, L.D. Free and Immobilized Biocatalysts for Removing Micropollutants from Water and Wastewater: Recent Progress and Challenges. Bioresour. Technol. 2022, 344, 126201. [Google Scholar] [CrossRef] [PubMed]
  115. Zdarta, J.; Meyer, A.S.; Jesionowski, T.; Pinelo, M. A General Overview of Support Materials for Enzyme Immobilization: Characteristics, Properties, Practical Utility. Catalysts 2018, 8, 92. [Google Scholar] [CrossRef]
  116. Sathishkumar, P.; Kamala-Kannan, S.; Cho, M.; Kim, J.S.; Hadibarata, T.; Salim, M.R.; Oh, B.-T. Laccase Immobilization on Cellulose Nanofiber: The Catalytic Efficiency and Recyclic Application for Simulated Dye Effluent Treatment. J. Mol. Catal. B Enzymatic 2014, 100, 111–120. [Google Scholar] [CrossRef]
  117. Taheran, M.; Naghdi, M.; Brar, S.; Knystautas, E.; Verma, M.; Surampalli, R. Degradation of Chlortetracycline Using Immobilized Laccase on Polyacrylonitrile-Biochar Composite Nanofibrous Membrane. Sci. Total. Environ. 2017, 605-606, 315–321. [Google Scholar] [CrossRef]
  118. Briggs, G.E.; Haldane, J.B.S. A Note on the Kinetics of Enzyme Action. Biochem. J. 1925, 19, 338–339. [Google Scholar] [CrossRef]
  119. Cleland, W.W. 3 Steady-State Kinetics. Enzymes 1990, 19, 99–158. [Google Scholar] [CrossRef]
  120. Lorsch, J.R. Practical Steady-State Enzyme Kinetics. Methods Enzym. 2014, 536, 3–15. [Google Scholar] [CrossRef]
  121. Elgharbawy, A.A.; Riyadi, F.A.; Alam, M.Z.; Moniruzzaman, M. Ionic Liquids as a Potential Solvent for Lipase-Catalysed Reactions: A Review. J. Mol. Liq. 2018, 251, 150–166. [Google Scholar] [CrossRef]
  122. Johnson, K.A. Kinetic Analysis for the New Enzymology: Using Computer Simulation to Learn Kinetics and Solve Mechanisms; KinTek Corp.: Snow Shoe, PA, USA, 2019; Volume 480. [Google Scholar]
  123. Boyle, J. Lehninger Principles of Biochemistry, 4th ed.; Nelson, D., Cox, M., Eds.; John Wiley & Sons, Ltd.: Bognor Regis, UK, 2005; Volume 33, ISBN 0716743396. [Google Scholar]
  124. Hermes, J.D.; Roeske, C.A.; O’Leary, M.H.; Cleland, W.W. Use of Multiple Isotope Effects to Determine Enzyme Mechanisms and Intrinsic Isotope Effects. Malic Enzyme and Glucose-6-Phosphate Dehydrogenase. Biochemistry 1982, 21, 5106–5114. [Google Scholar] [CrossRef]
  125. Watt, E.D.; Shimada, H.; Kovrigin, E.L.; Loria, J.P. The Mechanism of Rate-Limiting Motions in Enzyme Function. Proc. Natl. Acad. Sci. USA 2007, 104, 11981. [Google Scholar] [CrossRef] [PubMed]
  126. Wolfenden, R.; Snider, M.J. The Depth of Chemical Time and the Power of Enzymes as Catalysts. Acc. Chem. Res. 2001, 34, 938–945. [Google Scholar] [CrossRef]
  127. Brooks, K.M.; Hampel, K.J. A Rate-Limiting Conformational Step in the Catalytic Pathway of the GlmS Ribozyme. Biochemistry 2009, 48, 5669–5678. [Google Scholar] [CrossRef] [PubMed]
  128. Park, C. Visual Interpretation of the Meaning of Kcat/KMin Enzyme Kinetics. J. Chem. Educ. 2022, 99, 2556–2562. [Google Scholar] [CrossRef]
  129. Johnson, K.A. New Standards for Collecting and Fitting Steady State Kinetic Data. Beilstein J. Org. Chem. 2019, 15, 16–29. [Google Scholar] [CrossRef]
  130. Krooshof, G.H.; Kwant, E.M.; Damborský, J.; Koča, J.; Janssen, D.B. Repositioning the Catalytic Triad Aspartic Acid of Haloalkane Dehalogenase: Effects on Stability, Kinetics, and Structure. Biochemistry 1997, 36, 9571–9580. [Google Scholar] [CrossRef]
  131. Khafaga, D.S.R.; Muteeb, G.; Elgarawany, A.; Aatif, M.; Farhan, M.; Allam, S.; Almatar, B.A.; Radwan, M.G. Green Nanobiocatalysts: Enhancing Enzyme Immobilization for Industrial and Biomedical Applications. PeerJ 2024, 12, e17589. [Google Scholar] [CrossRef]
  132. Ge, J.; Lei, J.; Zare, R.N. Protein–Inorganic Hybrid Nanoflowers. Nat. Nanotechnol. 2012, 7, 428–432. [Google Scholar] [CrossRef]
  133. Guo, S.; Li, H.; Liu, J.; Yang, Y.; Kong, W.; Qiao, S.; Huang, H.; Liu, Y.; Kang, Z. Visible-Light-Induced Effects of Au Nanoparticle on Laccase Catalytic Activity. ACS Appl. Mater. Interfaces 2015, 7, 20937–20944. [Google Scholar] [CrossRef] [PubMed]
  134. An, J.; Li, G.; Zhang, Y.; Zhang, T.; Liu, X.; Gao, F.; Peng, M.; He, Y.; Fan, H. Recent Advances in Enzyme-Nanostructure Biocatalysts with Enhanced Activity. Catalysts 2020, 10, 338. [Google Scholar] [CrossRef]
  135. Han, L.; Liu, A. Novel Cell–Inorganic Hybrid Catalytic Interfaces with Enhanced Enzymatic Activity and Stability for Sensitive Biosensing of Paraoxon. ACS Appl. Mater. Interfaces 2017, 9, 6894–6901. [Google Scholar] [CrossRef]
  136. Holyavka, M.G.; Goncharova, S.S.; Redko, Y.A.; Lavlinskaya, M.S.; Sorokin, A.V.; Artyukhov, V.G. Novel Biocatalysts Based on Enzymes in Complexes with Nano- and Micromaterials. Biophys. Rev. 2023, 15, 1127. [Google Scholar] [CrossRef]
  137. Brezovsky, J.; Babkova, P.; Degtjarik, O.; Fortova, A.; Gora, A.; Iermak, I.; Rezacova, P.; Dvorak, P.; Smatanova, I.K.; Prokop, Z.; et al. Engineering a de Novo Transport Tunnel. ACS Catal. 2016, 6, 7597–7610. [Google Scholar] [CrossRef]
  138. Biedermannová, L.; Prokop, Z.; Gora, A.; Chovancová, E.; Kovács, M.; Damborský, J.; Wade, R.C. A Single Mutation in a Tunnel to the Active Site Changes the Mechanism and Kinetics of Product Release in Haloalkane Dehalogenase LinB. J. Biol. Chem. 2012, 287, 29062–29074. [Google Scholar] [CrossRef]
  139. Klvana, M.; Pavlova, M.; Koudelakova, T.; Chaloupkova, R.; Dvorak, P.; Prokop, Z.; Stsiapanava, A.; Kuty, M.; Kuta-Smatanova, I.; Dohnalek, J.; et al. Pathways and Mechanisms for Product Release in the Engineered Haloalkane Dehalogenases Explored Using Classical and Random Acceleration Molecular Dynamics Simulations. J. Mol. Biol. 2009, 392, 1339–1356. [Google Scholar] [CrossRef]
  140. Vardar, G.; Wood, T.K. Alpha-Subunit Positions Methionine 180 and Glutamate 214 of Pseudomonas Stutzeri OX1 Toluene-o-Xylene Monooxygenase Influence Catalysis. J. Bacteriol. 2005, 187, 1511–1514. [Google Scholar] [CrossRef]
  141. Zheng, F.; Tu, T.; Wang, X.; Wang, Y.; Ma, R.; Su, X.; Xie, X.; Yao, B.; Luo, H. Enhancing the Catalytic Activity of a Novel GH5 Cellulase GtCel5 from Gloeophyllum Trabeum CBS 900.73 by Site-Directed Mutagenesis on Loop 6. Biotechnol. Biofuels 2018, 11, 76. [Google Scholar] [CrossRef]
  142. Shannon, A.E.; Pedroso, M.M.; Chappell, K.J.; Watterson, D.; Liebscher, S.; Kok, W.M.; Fairlie, D.P.; Schenk, G.; Young, P.R. Product Release Is Rate-Limiting for Catalytic Processing by the Dengue Virus Protease. Sci. Rep. 2016, 6, 37539. [Google Scholar] [CrossRef]
  143. Oyen, D.; Fenwick, R.B.; Stanfield, R.L.; Dyson, H.J.; Wright, P.E. Cofactor-Mediated Conformational Dynamics Promote Product Release from Escherichia Coli Dihydrofolate Reductase via an Allosteric Pathway. J. Am. Chem. Soc. 2015, 137, 9459–9468. [Google Scholar] [CrossRef]
  144. Oyen, D.; Bryn Fenwick, R.; Aoto, P.C.; Stanfield, R.L.; Wilson, I.A.; Jane Dyson, H.; Wright, P.E. Defining the Structural Basis for Allosteric Product Release from E. Coli Dihydrofolate Reductase Using NMR Relaxation Dispersion. J. Am. Chem. Soc. 2017, 139, 11233–11240. [Google Scholar] [CrossRef] [PubMed]
  145. Thoma, N.H.; Iakovenko, A.; Kalinin, A.; Waldmann, H.; Goody, R.S.; Alexandrov, K. Allosteric Regulation of Substrate Binding and Product Release in Geranylgeranyltransferase Type II. Biochemistry 2001, 40, 268–274. [Google Scholar] [CrossRef] [PubMed]
  146. Brown, C.W.; Oh, E.; Hastman, D.A.; Walper, S.A.; Susumu, K.; Stewart, M.H.; Deschamps, J.R.; Medintz, I.L. Kinetic Enhancement of the Diffusion-Limited Enzyme Beta-Galactosidase When Displayed with Quantum Dots. RSC Adv. 2015, 5, 93089–93094. [Google Scholar] [CrossRef]
  147. Breger, J.C.; Ancona, M.G.; Walper, S.A.; Oh, E.; Susumu, K.; Stewart, M.H.; Deschamps, J.R.; Medintz, I.L. Understanding How Nanoparticle Attachment Enhances Phosphotriesterase Kinetic Efficiency. ACS Nano 2015, 9, 8491–8503. [Google Scholar] [CrossRef]
  148. Malanoski, A.P.; Breger, J.C.; Brown, C.W.; Deschamps, J.R.; Susumu, K.; Oh, E.; Anderson, G.P.; Walper, S.A.; Medintz, I.L. Kinetic Enhancement in High-Activity Enzyme Complexes Attached to Nanoparticles. Nanoscale Horiz. 2017, 2, 241–252. [Google Scholar] [CrossRef]
  149. Wang, L.; Cao, K.; Pedroso, M.M.; Wu, B.; Gao, Z.; He, B.; Schenk, G. Sequence- And Structure-Guided Improvement of the Catalytic Performance of a GH11 Family Xylanase from Bacillus Subtilis. J. Biol. Chem. 2021, 297, 101262. [Google Scholar] [CrossRef]
  150. Goryanova, B.; Goldman, L.M.; Ming, S.; Amyes, T.L.; Gerlt, J.A.; Richard, J.P. Rate and Equilibrium Constants for an Enzyme Conformational Change during Catalysis by Orotidine 5′-Monophosphate Decarboxylase. Biochemistry 2015, 54, 4555–4564. [Google Scholar] [CrossRef] [PubMed]
  151. Duff, M.R.; Redzic, J.S.; Ryan, L.P.; Paukovich, N.; Zhao, R.; Nix, J.C.; Pitts, T.M.; Agarwal, P.; Eisenmesser, E.Z. Structure, Dynamics and Function of the Evolutionarily Changing Biliverdin Reductase B Family. J. Biochem. 2020, 168, 191–202. [Google Scholar] [CrossRef] [PubMed]
  152. Hamnevik, E.; Enugala, T.R.; Maurer, D.; Ntuku, S.; Oliveira, A.; Dobritzsch, D.; Widersten, M. Relaxation of Nonproductive Binding and Increased Rate of Coenzyme Release in an Alcohol Dehydrogenase Increases Turnover with a Nonpreferred Alcohol Enantiomer. FEBS J. 2017, 284, 3895–3914. [Google Scholar] [CrossRef]
  153. Blikstad, C.; Dahlström, K.M.; Salminen, T.A.; Widersten, M. Substrate Scope and Selectivity in Offspring to an Enzyme Subjected to Directed Evolution. FEBS J. 2014, 281, 2387–2398. [Google Scholar] [CrossRef]
  154. Ugwumba, I.N.; Ozawa, K.; Xu, Z.Q.; Ely, F.; Foo, J.L.; Herlt, A.J.; Coppin, C.; Brown, S.; Taylor, M.C.; Ollis, D.L.; et al. Improving a Natural Enzyme Activity through Incorporation of Unnatural Amino Acids. J. Am. Chem. Soc. 2011, 133, 326–333. [Google Scholar] [CrossRef]
  155. Ryckelynck, M.; Baudrey, S.; Rick, C.; Marin, A.; Coldren, F.; Westhof, E.; Griffiths, A.D. Using Droplet-Based Microfluidics to Improve the Catalytic Properties of RNA under Multiple-Turnover Conditions. RNA 2015, 21, 458–469. [Google Scholar] [CrossRef]
  156. An, X.; Chen, C.; Wang, T.; Huang, A.; Zhang, D.; Han, M.J.; Wang, J. Genetic Incorporation of Selenotyrosine Significantly Improves Enzymatic Activity of Agrobacterium Radiobacter Phosphotriesterase. Chembiochem 2021, 22, 2535–2539. [Google Scholar] [CrossRef]
  157. Quaye, J.A.; Ouedraogo, D.; Gadda, G. Targeted Mutation of a Non-Catalytic Gating Residue Increases the Rate of Pseudomonas Aeruginosa d-Arginine Dehydrogenase Catalytic Turnover. J. Agric. Food Chem. 2023, 71, 17343–17352. [Google Scholar] [CrossRef]
  158. Wu, Q.; Zhang, C.; Dong, W.; Lu, H.; Yang, Y.; Li, W.; Xu, Y.; Li, X. Simultaneously Enhanced Thermostability and Catalytic Activity of Xylanase from Streptomyces Rameus L2001 by Rigidifying Flexible Regions in Loop Regions of the N-Terminus. J. Agric. Food Chem. 2023, 71, 12785–12796. [Google Scholar] [CrossRef]
  159. Han, S.S.; Kyeong, H.H.; Choi, J.M.; Sohn, Y.K.; Lee, J.H.; Kim, H.S. Engineering of the Conformational Dynamics of an Enzyme for Relieving the Product Inhibition. ACS Catal. 2016, 6, 8440–8445. [Google Scholar] [CrossRef]
  160. Pavlova, M.; Klvana, M.; Prokop, Z.; Chaloupkova, R.; Banas, P.; Otyepka, M.; Wade, R.C.; Tsuda, M.; Nagata, Y.; Damborsky, J. Redesigning Dehalogenase Access Tunnels as a Strategy for Degrading an Anthropogenic Substrate. Nat. Chem. Biol. 2009, 5, 727–733. [Google Scholar] [CrossRef]
  161. Ying, H.; Wang, J.; Shi, T.; Zhao, Y.; Ouyang, P.; Chen, K. Engineering of Lysine Cyclodeaminase Conformational Dynamics for Relieving Substrate and Product Inhibitions in the Biosynthesis of L-Pipecolic Acid. Catal. Sci. Technol. 2019, 9, 398–405. [Google Scholar] [CrossRef]
  162. Barlow, J.N.; Conrath, K.; Steyaert, J. Substrate-Dependent Modulation of Enzyme Activity by Allosteric Effector Antibodies. Biochim. Et Biophys. Acta (BBA)-Proteins Proteom. 2009, 1794, 1259–1268. [Google Scholar] [CrossRef] [PubMed]
  163. Lima, A.P.C.A.; Almeida, P.C.; Tersariol, I.L.S.; Schmitz, V.; Schmaier, A.H.; Juliano, L.; Hirata, I.Y.; Müller-Esterl, W.; Chagas, J.R.; Scharfstein, J. Heparan Sulfate Modulates Kinin Release by Trypanosoma Cruzi through the Activity of Cruzipain. J. Biol. Chem. 2002, 277, 5875–5881. [Google Scholar] [CrossRef] [PubMed]
  164. Röthlisberger, D.; Khersonsky, O.; Wollacott, A.M.; Jiang, L.; DeChancie, J.; Betker, J.; Gallaher, J.L.; Althoff, E.A.; Zanghellini, A.; Dym, O.; et al. Kemp Elimination Catalysts by Computational Enzyme Design. Nature 2008, 453, 190–195. [Google Scholar] [CrossRef]
  165. Jiang, L.; Althoff, E.A.; Clemente, F.R.; Doyle, L.; Röthlisberger, D.; Zanghellini, A.; Gallaher, J.L.; Betker, J.L.; Tanaka, F.; Barbas, C.F.; et al. De Novo Computational Design of Retro-Aldol Enzymes. Science 2008, 319, 1387–1391. [Google Scholar] [CrossRef]
  166. Burke, A.J.; Lovelock, S.L.; Frese, A.; Crawshaw, R.; Ortmayer, M.; Dunstan, M.; Levy, C.; Green, A.P. Design and Evolution of an Enzyme with a Non-Canonical Organocatalytic Mechanism. Nature 2019, 570, 219–223. [Google Scholar] [CrossRef]
  167. Li, A.; Ye, L.; Yang, X.; Wang, B.; Yang, C.; Gu, J.; Yu, H. Reconstruction of the Catalytic Pocket and Enzyme–Substrate Interactions to Enhance the Catalytic Efficiency of a Short-Chain Dehydrogenase/Reductase. ChemCatChem 2016, 8, 3229–3233. [Google Scholar] [CrossRef]
  168. Siegel, J.B.; Zanghellini, A.; Lovick, H.M.; Kiss, G.; Lambert, A.R.; St.Clair, J.L.; Gallaher, J.L.; Hilvert, D.; Gelb, M.H.; Stoddard, B.L.; et al. Computational Design of an Enzyme Catalyst for a Stereoselective Bimolecular Diels-Alder Reaction. Science 2010, 329, 309–313. [Google Scholar] [CrossRef]
  169. Obexer, R.; Godina, A.; Garrabou, X.; Mittl, P.R.E.; Baker, D.; Griffiths, A.D.; Hilvert, D. Emergence of a Catalytic Tetrad during Evolution of a Highly Active Artificial Aldolase. Nat. Chem. 2016, 9, 50–56. [Google Scholar] [CrossRef]
  170. Wilkinson, H.C.; Dalby, P.A. Fine-Tuning the Activity and Stability of an Evolved Enzyme Active-Site through Noncanonical Amino-Acids. FEBS J. 2021, 288, 1935–1955. [Google Scholar] [CrossRef] [PubMed]
  171. Akbarzadeh, A.; Pourzardosht, N.; Dehnavi, E.; Ranaei Siadat, S.O.; Zamani, M.R.; Motallebi, M.; Nikzad Jamnani, F.; Aghaeepoor, M.; Barshan Tashnizi, M. Disulfide Bonds Elimination of Endoglucanase II from Trichoderma Reesei by Site-Directed Mutagenesis to Improve Enzyme Activity and Thermal Stability: An Experimental and Theoretical Approach. Int. J. Biol. Macromol 2018, 120, 1572–1580. [Google Scholar] [CrossRef]
  172. Jiang, W.; Wang, Y. Improving Catalytic Efficiency and Changing Substrate Spectrum for Asymmetric Biocatalytic Reductive Amination. J. Microbiol. Biotechnol. 2019, 30, 146. [Google Scholar] [CrossRef] [PubMed]
  173. Nam, H.; Hwang, B.J.; Choi, D.Y.; Shin, S.; Choi, M. Tobacco Etch Virus (TEV) Protease with Multiple Mutations to Improve Solubility and Reduce Self-Cleavage Exhibits Enhanced Enzymatic Activity. FEBS Open Bio 2020, 10, 619–626. [Google Scholar] [CrossRef]
  174. Ilić Đurđić, K.; Ostafe, R.; Prodanović, O.; Đurđević Đelmaš, A.; Popović, N.; Fischer, R.; Schillberg, S.; Prodanović, R. Improved Degradation of Azo Dyes by Lignin Peroxidase Following Mutagenesis at Two Sites near the Catalytic Pocket and the Application of Peroxidase-Coated Yeast Cell Walls. Front. Environ. Sci. Eng. 2021, 15, 19. [Google Scholar] [CrossRef]
  175. Dai, S.; Yao, Q.; Yu, G.; Liu, S.; Yun, J.; Xiao, X.; Deng, Z.; Li, H. Biochemical Characterization of a Novel Bacterial Laccase and Improvement of Its Efficiency by Directed Evolution on Dye Degradation. Front. Microbiol. 2021, 12, 633004. [Google Scholar] [CrossRef] [PubMed]
  176. Zhu, L.; Cheng, F.; Piatkowski, V.; Schwaneberg, U. Protein Engineering of the Antitumor Enzyme PpADI for Improved Thermal Resistance. ChemBioChem 2014, 15, 276–283. [Google Scholar] [CrossRef]
  177. Sun, S.; Tan, Y.; Wang, L.; Wu, Z.; Zhou, J.; Wu, G.; Shao, Y.; Wang, M.; Song, Z.; Xin, Z. Improving the Activity and Expression Level of a Phthalate-Degrading Enzyme by a Combination of Mutagenesis Strategies and Strong Promoter Replacement. Environ. Sci. Pollut. Res. 2023, 30, 41107–41119. [Google Scholar] [CrossRef]
  178. Zhang, X.G.; Xue, Y.; Lu, Z.P.; Xu, H.J.; Hu, Y. Chemical Modification for Improving Catalytic Performance of Lipase B from Candida Antarctica with Hydrophobic Proline Ionic Liquid. Bioprocess Biosyst. Eng. 2022, 45, 749–759. [Google Scholar] [CrossRef]
  179. Liu, Z.Q.; Yin, H.H.; Zhang, X.J.; Zhou, R.; Wang, Y.M.; Zheng, Y.G. Improvement of Carbonyl Reductase Activity for the Bioproduction of Tert-Butyl (3R, 5S)-6-Chloro-3, 5-Dihydroxyhexanoate. Bioorg. Chem. 2018, 80, 733–740. [Google Scholar] [CrossRef]
  180. Cheng, Y.S.; Chen, C.C.; Huang, J.W.; Ko, T.P.; Huang, Z.; Guo, R.T. Improving the Catalytic Performance of a GH11 Xylanase by Rational Protein Engineering. Appl. Microbiol. Biotechnol. 2015, 99, 9503–9510. [Google Scholar] [CrossRef]
  181. Voet, D.; Voet, J.G.; Pratt, C.W. Principles of Biochemistry, 3rd ed.; Wiley: Hoboken, NJ, USA, 2008; ISBN 0470233966. [Google Scholar]
  182. Northrop, D.B.; Scovell, W.M. On the Meaning of Km and V/K in Enzyme Kinetics. J. Chem. Educ. 1998, 75, 1153–1157. [Google Scholar] [CrossRef]
  183. Wang, X.; Sheng, Y.; Cui, H.; Qiao, J.; Song, Y.; Li, X.; Huang, H. Corner Engineering: Tailoring Enzymes for Enhanced Resistance and Thermostability in Deep Eutectic Solvents. Angew. Chem. 2024, 136, e202315125. [Google Scholar] [CrossRef]
  184. Ranaei Siadat, S.O.; Mollasalehi, H.; Heydarzadeh, N. Substrate Affinity and Catalytic Efficiency Are Improved by Decreasing Glycosylation Sites in Trichoderma Reesei Cellobiohydrolase I Expressed in Pichia Pastoris. Biotechnol. Lett. 2016, 38, 483–488. [Google Scholar] [CrossRef]
  185. Gao, H.; Zhu, R.; Li, Z.; Wang, W.; Liu, Z.; Hu, N. Improving the Catalytic Efficiency and Substrate Affinity of a Novel Esterase from Marine Klebsiella Aerogenes by Random and Site-Directed Mutation. World J. Microbiol. Biotechnol. 2021, 37, 106. [Google Scholar] [CrossRef]
  186. Fang, Z.; Zhang, J.; Liu, B.; Du, G.; Chen, J. Enhancement of the Catalytic Efficiency and Thermostability of Stenotrophomonas Sp. Keratinase KerSMD by Domain Exchange with KerSMF. Microb. Biotechnol. 2016, 9, 35–46. [Google Scholar] [CrossRef] [PubMed]
  187. Zhang, Y.; Liu, M.; Wang, H.; Lin, J.; Chen, F. Focused Mutagenesis in Non-Catalytic Cavity for Improving Catalytic Efficiency of 3-Ketosteroid-Δ1-Dehydrogenase. Mol. Catal. 2022, 531, 112661. [Google Scholar] [CrossRef]
  188. Yang, J.; Ma, T.; Shang-guan, F.; Han, Z. Improving the Catalytic Activity of Thermostable Xylanase from Thermotoga Maritima via Mutagenesis of Non-Catalytic Residues at Glycone Subsites. Enzyme. Microb. Technol. 2020, 139, 109579. [Google Scholar] [CrossRef]
  189. Yuan, Y.; Song, W.; Liu, J.; Chen, X.; Luo, Q.; Liu, L. Production of α-Ketoisocaproate and α-Keto-β-Methylvalerate by Engineered L-Amino Acid Deaminase. ChemCatChem 2019, 11, 2464–2472. [Google Scholar] [CrossRef]
  190. Mesbah, N.M. Industrial Biotechnology Based on Enzymes from Extreme Environments. Front. Bioeng. Biotechnol. 2022, 10, 870083. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Enzymes as transforming agents in biocatalysis.
Figure 1. Enzymes as transforming agents in biocatalysis.
Catalysts 15 00736 g001
Figure 2. Protein engineering as a tool for generating enhanced enzymes.
Figure 2. Protein engineering as a tool for generating enhanced enzymes.
Catalysts 15 00736 g002
Figure 3. Summarized workflow of the directed evolution and rational design techniques in protein engineering.
Figure 3. Summarized workflow of the directed evolution and rational design techniques in protein engineering.
Catalysts 15 00736 g003
Figure 4. Three critical areas of industrial biocatalysis. Biocatalyst engineering, solvent engineering, and mechanistic enzymology are three interconnected fields that have unique roles in advancing industrial biocatalysis.
Figure 4. Three critical areas of industrial biocatalysis. Biocatalyst engineering, solvent engineering, and mechanistic enzymology are three interconnected fields that have unique roles in advancing industrial biocatalysis.
Catalysts 15 00736 g004
Figure 5. Difference between micro- and macroviscosity in enzyme kinetics.
Figure 5. Difference between micro- and macroviscosity in enzyme kinetics.
Catalysts 15 00736 g005
Figure 6. A plot of (kcat/Km)o/(kcat/Km)η vs. ηrel, illustrating the different outcomes of trends for a KSVE technique. The red line, characterized by a slope of 1, denotes reactions in which the capture of the substrate in enzyme−substrate complexes is fully committed to catalysis in the forward direction. The blue line, characterized by a slope of 0, denotes reactions in which the chemical step of catalysis fully limits enzyme turnover. The green line, characterized by a slope of between 0 and 1, denotes reactions in which the capture of the substrate in enzyme−substrate complexes is partially committed to catalysis in the forward direction.
Figure 6. A plot of (kcat/Km)o/(kcat/Km)η vs. ηrel, illustrating the different outcomes of trends for a KSVE technique. The red line, characterized by a slope of 1, denotes reactions in which the capture of the substrate in enzyme−substrate complexes is fully committed to catalysis in the forward direction. The blue line, characterized by a slope of 0, denotes reactions in which the chemical step of catalysis fully limits enzyme turnover. The green line, characterized by a slope of between 0 and 1, denotes reactions in which the capture of the substrate in enzyme−substrate complexes is partially committed to catalysis in the forward direction.
Catalysts 15 00736 g006
Figure 7. A plot of (kcat)o/(kcat)η vs. ηrel, showing the different trend outcomes for a KSVE technique. The red line, characterized by a slope of 1, denotes reactions in which the overall turnover is fully limited by the rate of product release from the active site. The blue line, characterized by a slope of 0, denotes reactions in which the chemical step of catalysis fully limits enzyme turnover. The green line, characterized by a slope between 0 and 1, denotes reactions in which the overall turnover is partially limited by the chemical step of catalysis and the rate of product release from the active site.
Figure 7. A plot of (kcat)o/(kcat)η vs. ηrel, showing the different trend outcomes for a KSVE technique. The red line, characterized by a slope of 1, denotes reactions in which the overall turnover is fully limited by the rate of product release from the active site. The blue line, characterized by a slope of 0, denotes reactions in which the chemical step of catalysis fully limits enzyme turnover. The green line, characterized by a slope between 0 and 1, denotes reactions in which the overall turnover is partially limited by the chemical step of catalysis and the rate of product release from the active site.
Catalysts 15 00736 g007
Scheme 1. The simplest kinetic mechanism for an enzyme (E) with a single substrate (S) and product (P), involving substrate binding (k1) and dissociation (k2), reversible catalysis (k3 and k4), and product release (k5) [48].
Scheme 1. The simplest kinetic mechanism for an enzyme (E) with a single substrate (S) and product (P), involving substrate binding (k1) and dissociation (k2), reversible catalysis (k3 and k4), and product release (k5) [48].
Catalysts 15 00736 sch001
Figure 8. Synergy between mechanistic enzymology and enzyme engineering is key for industrial biocatalysis.
Figure 8. Synergy between mechanistic enzymology and enzyme engineering is key for industrial biocatalysis.
Catalysts 15 00736 g008
Table 1. Genetically engineered proteins with their modified structural properties and industrial applications.
Table 1. Genetically engineered proteins with their modified structural properties and industrial applications.
ProteinMajor Structural CharacteristicApplicationRef.
Bacteriophage lysine,
Ply187N-V12C
Catalytic domain of Ply187 lysin; bacterial cell-binding domain of PlyV12 lysinEnhanced antimicrobial therapy due to an increase in lytic activity against human pathogens[30]
Tumor-associated antigens MUC1 and HER2MUC1subunit; HER2 subunitEnhanced breast cancer detection in ELISA against MUC1 or HER2[32]
α-Glucanohydrolase, mut-dexAN-terminal mutanase domain; C-terminal dextranase domainEnhanced α-glucan hydrolysis in fighting dental biofilm formation[33]
Bacillus subtilis phytaseN-terminal β-1,4 endoglucanase domain; C-terminal phytase domainEnhanced nutrition uptake as a potential feed additive for monogastric animals[34]
Glucose 1-dehydrogenase IV, BmGlcDH-IVMutation G259AAltered substrate specificity to exclusively D-glucose for clinical testing of blood glucose levels[35]
Bacillus subtilis aminopeptidase, BSAPMutations I387A, I387C and I387SModified substrate specificity to enable hydrolysis of phenylalanine derivatives with bulky side chains[36]
Transmembrane zinc metallopeptidase Neprilysin, NEPMutations G399V/G714KAltered protein cleavage site specificity to target the Phe20-Ala21 bond in amyloid β 1-40 for Alzheimer’s disease treatment[37]
Mucor hiemalis endo-b-N-acetylglucosaminidase, Endo-MMutation N175QEnhanced glycosynthase activity with oxazoline and natural N glycan for the synthesis of glycoproteins[38]
Candida antarctica lipase B, CalBMutations G39A, T103G, W104F, A141Q, I189Y, L278AIncreasing amidase activity to three times that of the wild type for the production of optically active amines[39]
Guanylate kinase (GK)Mutation S68PAltering GK function from enzymatic activity to protein-binding to regulate mitotic spindle orientation and cell adhesion[40]
Table 2. Microviscosigens and macroviscosigens, with their corresponding molecular weight (MW) ranges.
Table 2. Microviscosigens and macroviscosigens, with their corresponding molecular weight (MW) ranges.
MicroviscosigenMacroviscosigen
MW (Da) Typical MW Range (kDa)
Glucose180.2Polyethylene glycol 200–20,000
Sucrose342.3PolyacrylamideVaries, reaching several millions
Glycerol92.1Ficoll300–700
Ethylene glycol62.1Dextran40–500
Trehalose342.3Polyethylene oxide100–1000
Erythritol122.1Polyvinylpyrrolidone10–360
Fructose180.2Methylcellulose30–150
Maltose342.3Carboxymethylcellulose90–700
Mannitol182.2Agarose~120,000
Sorbitol182.2Xanthan gum~2000
Table 3. DESs with their respective HBAs, HBDs, mole ratios, and melting temperatures.
Table 3. DESs with their respective HBAs, HBDs, mole ratios, and melting temperatures.
Hydrogen Bond Acceptor (HBA)Hydrogen Bond Donor (HBD)Mole RatioMelting Temperature (°C)Ref.
Choline chloride1-Methyl urea 1:229[105]
Choline chlorideBenzamide1:292[105]
Choline chloride1,1-Dimethyl urea1:2149[105]
Choline chloride4-Hydroxybenzoic acid1:297[106]
Choline chloride1,3-Dimethyl urea1:270[105]
Choline tetrafluoroborateUrea1:267[105]
Diethylethanolammonium chlorideGlycerol1:2−1.2[107]
Diethylethanolammonium chlorideTrifluoroacetamide1:20.05[108]
N,N-Diethylenethanolammonium chlorideGlycerol1:2−1.33[109]
Tetrabutylammonium chlorideEthylene glycol 1:4−16.84[110]
Tetrabutylammonium chlorideUrea4:127.14[111]
Tetrabutylammonium chlorideDecanoic acid 1:2−11.95[112]
Tetrapropylammonium bromideGlycerol1:3−16.1[109]
Tetrapropylammonium bromideEthylene glycol1:3−13.3[113]
Propylammonium bromideGlycerol1:2−4[114]
Table 4. Strategies for improving the product release step of an enzyme-catalyzed reaction.
Table 4. Strategies for improving the product release step of an enzyme-catalyzed reaction.
CategoryMethodologyKey FindingsRef.
Nanoparticle technologyβ-Galactosidase was coupled with CdSe/ZnS core/shell quantum dots of different sizes and wavelengths. Quantum dots were stabilized by DHLA-CL4 and DHLA-PEG-CL4 ligands.β-Galactosidase with quantum dots exhibited a 3-fold increase in the kcat value, resulting from an increased product dissociation rate. The unique microenvironment surrounding the hydration layer of the enzyme–QD complex was attributed to the observed effect.[147]
CdSe/ZnS core/shell quantum dots (QDs) of different sizes were conjugated with phosphotriesterase (PTE), followed by extensive characterization.When conjugated to QDs, a 4-fold increase in the kcat value and a 2-fold increase in the kcat/Km value were shown by PTE. The enhanced activity was attributed to an increase in the enzyme–product dissociation rate, likely due to the varying microenvironment of the hydration layer of the PTE-QD complex.[148]
A model system was used to test a proposed detailed reaction scheme involving various possible interactions between a substrate and an enzyme–nanoparticle bioconjugate.The enzyme attached to nanoparticles showed a 3.2-fold increase in the kcat value and a 1.7-fold increase in the kcat/Km value. The proposed detailed reaction scheme suggested that product release becomes rate-limiting when the enzyme is attached to the nanoparticles.[149]
MutagenesisSite saturation and iterative mutagenesis were performed on the identified N-terminal residues 5-YWQN-8 in GH11 xylanase XynLC9 from Bacillus subtilis to produce mutants W6F/Q7H and N8Y.The W6F/Q7H and N8Y mutants exhibited increased kcat values by 2.6-fold and 1.8-fold, respectively, and kcat/Km values by 1.5-fold and 1.2-fold, respectively. Both mutants were more thermostable and had enlarged active site clefts.[150]
DNA shuffling was employed to mutagenize the trpC gene, which encodes indole glycerol phosphate synthase, from the hyperthermophilic archaeon Sulfolobus solfataricus.Single and double amino acid mutants had improved activity and increased product release rates. The rate-determining step changed from product dissociation to the chemical step of catalysis.[17]
Site-directed mutagenesis was performed on specific residues of orotidine 5′-monophosphate decarboxylase from Saccharomyces cerevisiae to give the mutants Q215A and Y217F.The Q215A and Y217F mutants exhibited increased kcat and decreased kcat/Km, resulting from increased product release rate and a weakening of protein–phosphodianion interactions.[151]
The structural and functional effects of the evolutionarily changing active site in the biliverdin reductase B (BLVRB) family were investigated by creating multiple human BLVRB mutants through site-directed mutagenesis.A novel mechanism was proposed where coenzyme “clamps” formed by arginine side chains at positions 14 and 78 slowed coenzyme release and enzyme turnover in wild-type enzymes. However, an inverse association was observed between coenzyme release and enzyme turnover for different BLVRB mutants.[152]
Enantioselectivity was modified in alcohol dehydrogenase A (ADH-A) from Rhodococcus ruber DSM 44541 by performing saturation mutagenesis on three targeted sites in the catalytic cavity of ADH-A.The stereoselectivity of the enzyme was altered from (S)- to (R)-1-phenylethanol by the W295A mutant, resulting in a 4-fold increase in the kcat value. W295A showed an enhanced NADH release rate, shifting the rate-determining step from coenzyme release to hydride transfer. Other mutants also showed an increased kcat value relative to the wild type.[153]
Directed evolution was applied to Escherichia coli S-1,2-propanediol oxidoreductase (FucO) in the study, followed by site-directed mutagenesis of specific residues, including F254 and N151.A dual role in enzyme function was found to be played by the residue F254, contributing to increased coenzyme dissociation from the active site and a 4-fold increase in the kcat value with S-1,2-propanediol.[154]
Both natural and non-canonical amino acids were incorporated into bacterial phosphotriesterase (PTE) at position T309 to create different mutants.Replacing Tyr309 with unnatural amino acids improved the phosphotriesterase turnover rate up to 11-fold and the kcat/Km value up to 4-fold. The rate-limiting product release rate was observed to be increased by the deprotonated 7-hydroxyl group of L-(7-hydroxycoumarin-4-yl)ethylglycine through the electrostatic repulsion of 4-nitrophenolate.[155]
A novel approach to in vitro evolution of RNA using droplet-based microfluidics was presented, and cycles of random mutagenesis and selection were conducted to evolve the X-motif.Evolved X-motif ribozyme variants showed an ~28-fold higher kcat value relative to the wild type. This increase in enzyme turnover was attributed to an improved product release rate, which was identified as the rate-limiting step.[156]
The catalytic activity of Agrobacterium radiobacter phosphotriesterase was enhanced by the genetic incorporation of the unnatural amino acid selenotyrosine (SeHF) at Tyr309.The SeHF309 mutant resulted in a 12-fold increase in the kcat value and a 3.2-fold increase in the kcat/Km value. The product release pocket was observed to be opened by the SeHF309 mutant, resulting in an increased rate of product release.[157]
Loop dynamicsSite-directed mutagenesis was performed on E246 in loop L2 located at the entrance of the active site of Pseudomonas aeruginosa d-arginine dehydrogenase to produce the mutants E246Q, E246G, and E246L.The E246G mutant exhibited a 4-fold increase in the kcat value, while the E246G and E246L variants also showed increased kcat values. The increase in the turnover rate was attributed to an increase in the product release rate resulting from altered loop dynamics.[158]
Site-directed mutagenesis was performed on residues at the interface of the mobile loop of the NS2B-NS3pro complex, a two-component viral protease.Product release was primarily affected by mutations in the NS2B mobile loop, which triggered a conformational change that activated the catalytic center. A 3.4-fold increase in the kcat value and a 1.5-fold increase in the kcat/Km value were observed in the glycine-linked S83W mutant.[143]
The thermal stability and catalytic activity of GH11 xylanase XynA from Streptomyces rameus L2001 were improved by rational design when the combined mutations 11YHDGYF16, 23AP24/23SP24, and 32GP33 were introduced to give Mut1 and Mut2 variants.A 4-fold increase in the kcat value was observed for both Mut1 and Mut2, resulting from a broader catalytic cleft and increased “thumb” flexibility, directly affecting product release. Improved thermal stability was observed in both mutants, with over 85% residual activity after 12 h at 80–90 °C.[159]
Product inhibition was relieved in chorismate–pyruvate lyase (CPL) by the modulation of its conformational dynamics. Residues at the enzyme’s product-binding site were mutated to increase flap dynamics and facilitate product release.Mutants exhibited enhanced flap dynamics in chorismate–pyruvate lyase (CPL), characterized by an 8-fold reduction in product inhibition and a 3-fold increase in the kcat value. The product release rate was increased through flap opening without significantly changing the kcat/Km value.[160]
Modifying transport tunnelsThe role of tunnels in DhaA haloalkane dehalogenase was investigated by creating eight mutants via site-directed mutagenesis.Five pathways were discovered for product release and water exchange, with chloride ions released solely through p1 and alcohol released through all five pathways. All mutants exhibited increased kcat values, ranging from ~3- to 30-fold.[140]
A new transport tunnel was designed in haloalkane dehalogenase via computational design and directed evolution. Mutants were designed to block the native tunnel, allowing for the observation of the properties of the newly opened auxiliary tunnel.The LinBW mutant, with a blocked native tunnel, showed an increase in both product release rate and a 4-fold increase in the kcat value when the new auxiliary tunnel was opened. The LinBCC mutant, on the other hand, had a 2-fold decrease in enzyme turnover.[138]
The activity of Rhodococcus rhodochrous haloalkane dehalogenase was enhanced by creating mutants using a novel engineering strategy involving rational design and directed evolution to identify key residues in access tunnels.Higher enzyme activities were observed by all mutants, with up to a 32-fold increase in the kcat value and a 26-fold increase in the kcat/Km value. There was an increased rate of carbon–halogen breakage and a change in the rate-determining step to product release.[161]
The conformational dynamics of lysine cyclodeaminase from Streptomyces pristinaespiralis were modulated to address substrate and product inhibitions. The Val61-Val94-SpLCD variant was designed by mutating Ile61 and Ile94 residues associated with substrate and product delivery processes.The Val61-Val94-SpLCD variant showed improvements in the kcat, kcat/Km, Ki-LYS, and Ki-LPA values by 20, 4, 19, and 9 times, respectively. The expanded substrate and product delivery tunnels of the mutant reduced substrate and product inhibition.[162]
Using allosteric effectorsVHH domain antibodies as allosteric effectors on the nucleoside hydrolase from Trypanosoma vivax (TvNH) were studied. VHH 1589 inhibited N-glycosidic bond cleavage while increasing the rates of product release. The data indicated a substrate-dependent effect on kcat and kcat/Km, with a net increase in the kcat value observed for the natural nucleoside 7-methylguanosine, where product release was rate-limiting.[163]
The allosteric effects of heparan sulfate (HS) on cruzipain-mediated kinin release from kininogen (HK) were studied in Trypanosoma cruzi.HS reduced the inhibitory activity of HK on cruzipain by 10-fold and enhanced the kcat and kcat/Km values by 2.5-fold and 6-fold, respectively. The rate of cruzipain-mediated kinin release was also increased by HS up to 35-fold.[164]
The structural basis of allosteric product release in Escherichia coli dihydrofolate reductase (DHFR) was investigated, with a focus on the rate-determining step of tetrahydrofolate (THF) product release.The transient entry of the NADPH nicotinamide ring into the active site was observed to involve allosteric product release in DHFR, which resulted in the displacement of the THF’s pterin ring. A remodeling of the enzyme structure and conformation of the THF was attributed to the entry of NADPH, which facilitated a rapid product release.[145]
The mechanism of substrate binding and product release in the geranylgeranylation of Rab proteins by geranylgeranyltransferase type II (GGTase-II) was studied.GGpp was observed to be an allosteric activator of GGTase-II by causing an increase in enzyme affinity for the Rab7:REP-1 complex. GGpp was also found to facilitate product release by increasing the dissociation rate of prenylated Rab7:REP-1 complex.[146]
Table 5. Strategies for improving the chemical step of catalysis of an enzyme-catalyzed reaction.
Table 5. Strategies for improving the chemical step of catalysis of an enzyme-catalyzed reaction.
CategoryMethodologyKey FindingsRef.
Computational design and directed evolutionComputational design was used to create eight enzymes with two distinct catalytic motifs for the Kemp elimination reaction. Directed in vitro evolution was subsequently applied to refine the computational designs.The designed enzymes achieved rate enhancements of up to 105 and demonstrated multiple turnovers. Catalytic efficiency was further improved by directed evolution, yielding over 200-fold increase in the kcat/Km value and 106-fold increase in the kcat value relative to the uncatalyzed reaction (kuncat).[165]
Active sites in retro-aldolases were constructed using computational algorithms, which consisted of four distinct catalytic motifs for catalyzing carbon–carbon bond breakage in a nonnatural substrate. The use of explicit water molecules for proton shuffling and charged side-chain networks was included in the design.Of the 72 experimentally characterized designs, 32 exhibited detectable retro-aldolase activity across various protein folds. Designs that incorporated explicit water molecules were more successful, achieving significant increments in reaction rates.[166]
Computational design was used to create enzymes capable of catalyzing the bimolecular Diels–Alder reaction. A transaminase enzyme was modified through directed evolution to recognize complex ketone substrates.High stereoselectivity and substrate specificity were exhibited by the computationally designed Diels–Alder enzymes. The evolved transaminase successfully catalyzed the reaction with a complex ketone substrate, with an 87-fold increase in the kcat value.[169]
A droplet-based microfluidic screening tool was used to enhance an artificial aldolase by combining computational design and directed evolution to achieve properties comparable to those of the natural enzyme.The improved enzyme achieved a 109-fold increase in the enzyme turnover rate, with high stereoselectivity and broad substrate tolerance. It was revealed by further analysis that a Lys-Tyr-Asn-Tyr tetrad was important for enzyme activity.[170]
Incorporating non-canonical amino acidsThe active site of an Escherichia coli transketolase variant (S385Y/D469T/R520Q) was incorporated with non-canonical amino acids via site-specific saturation mutagenesis to produce different variants.The mutants showed a significant increase in the kcat and kcat/Km values. The residue 385 variants exhibited a 43-fold increase in specific activity, a 100% increase in kcat value, a 290% increase in the Km value, and a 240% increase in the kcat/Km value.[171]
Nδ-methylhistidine, a non-canonical amino acid, was incorporated as a catalytic nucleophile into the active site of a novel hydrolytic enzyme, followed by optimization procedures.Over a 9000-fold increase in the ester hydrolysis rate was observed in the engineered enzyme. Nδ-methylhistidine was revealed as a genetically encodable surrogate for dimethylaminopyridine through crystallographic analysis. Histidine methylation was noted to be important for enzyme activity by blocking the production of nonreactive acyl-enzyme intermediates.[167]
MutagenesisRational design was used to reconstruct the catalytic pocket and enzyme–substrate interactions of dehydrogenase/reductase EbSDR8 through targeted mutations at specific residues, namely, G94 and S153.Over a 15-fold improvement in the kcat/Km value was shown by the mutants compared to the wild-type. An increase in the kcat value was also observed for the mutants, and steric repulsion and C−H…π interactions were attributed to this enhanced activity.[168]
Disulfide bonds were eliminated via site-directed mutagenesis in Endoglucanase II (Cel5A) from Trichoderma reesei to create two mutants: C99V and C323H.A 2-fold increase in both the kcat/Km and kcat values was shown by the C99V mutant, whereas a 1.3-fold increase in the kcat/Km value was shown by the C323H mutant. Both mutants had increased flexibility in the substrate-binding cleft.[172]
Rational design involving active-site-targeted, site-specific mutagenesis to construct a broad substrate spectrum version of the phenylalanine dehydrogenase (PheDH) from Bacillus halodurans.Mutant PheDH (E113D-N276L) showed a 6-fold increase in the kcat/Km value for oxidative deamination and a 1.6-fold increase for reductive amination.[173]
Site-directed mutagenesis of Tobacco etch virus (TEV) protease, with mutations at T17S/N68D/I77V, was used to produce two different mutants: S219N and S219V.S219N and S219V showed ~100-fold and ~50-fold increases in their kcat values, respectively. S219N was twofold faster than S219V without a change in its Km value.[174]
Saturation mutagenesis was performed at two positions (165 and 264) near the catalytic Trp171 residue of lignin peroxidase from Phanerochaete chrysosporium.A significant increase in the kcat value (2-fold) and the kcat/Km value (13-fold) was shown by the mutants. A 10-fold increase in affinity for azo dye substrates was also observed.[175]
Site-directed mutagenesis was used to create a lac2-9 mutant from recombinant laccase rlac1338, followed by screening to identify those with improved enzyme activity.Mutant laccase lac2-9 showed a 3.5-fold increase in specific activity, a ~4-fold increase in the kcat/Km value, and a ~2-fold increase in the kcat value.[176]
In the study, various mutants were created from arginine deiminase from Pseudomonas plecoglossicida (PpADI) through directed evolution and site-directed mutagenesis, and improved enzyme activity and thermal stability were tested.The PpADI M9 mutant showed a 64.7-fold increase in the kcat value, a 1.75-fold increase in half-life, and an increase in thermal stability from 47 °C to 54 °C.[177]
Mutagenesis and strong promoter replacement strategies were used to create mutants of the PAE-degrading enzyme EstJ6, with the aim of enhancing its catalytic activity and expression level.The EstJ6M1.1, EstJ6M2, and EstJ6M3.1 mutants exhibited an increased kcat value by up to 2.9-fold, 3.1-fold, and 4.3-fold, respectively, and an increased kcat/Km value by up to 3.7-fold, 4.6-fold, and 6-fold, respectively.[178]
Candida antarctica lipase B (CALB) was modified using proline ionic liquids with varying lengths of hydrophobic alkyl chains on the side.CALB modified with [ProC12][H2PO4] showed a 3.0-fold increase in hydrolysis activity and a 2.8-fold improvement in the kcat/Km value. Thermal stability was increased as well.[179]
Random mutagenesis, site-saturation mutagenesis, and combinatorial mutagenesis were used to enhance the activity of a synthesized stereoselective short-chain carbonyl reductase (SCR), an enzyme that produces tert-butyl(3R,5S)-6-chloro-3,5-dihydroxyhexanoate ((3R,5S)-CDHH) for rosuvastatin synthesis.Mut-Phe145Met/Thr152Ser showed a 1.60-fold increase in the kcat value and a 5.11-fold improvement in the kcat/Km value. A 1.91-fold increase in the kcat value and an 8.07-fold increase in the kcat/Km value were shown by mut-Phe145Tyr/Thr152Ser. A yield of more than 99% in (3R,5S)-CDHH production was also achieved by both mutants.[180]
Site-directed mutagenesis and N-glycosylation analysis were employed to modify the active site residues of the xylanase XynCDBFV from Neocallimastix patriciarum, resulting in the creation of various mutants.The W125F/F163W double mutant showed a ~20% increase in the kcat value, and N-glycosylation at Asn-37 was found to play a dominant role in both catalytic activity and thermal stability.[181]
Table 6. Strategies for improving the catalytic efficiency of an enzyme-catalyzed reaction.
Table 6. Strategies for improving the catalytic efficiency of an enzyme-catalyzed reaction.
CategoryMethodologyKey FindingsRef.
MutagenesisCorner engineering was used to target the transition region in Bacillus subtilis lipase A to create multiple DES-resistant variants with improved activity and thermal stability.The K88E/N89K mutant showed a 10-fold increase in the kcat/Km value in 30% ChCl–acetamide and a 4.1-fold increase in 95% ChCl–ethylene glycol. A 6.7-fold improvement in thermal resistance was observed for the mutants at 50 °C. The approach was validated with two additional industrial enzymes.[184]
Site-directed mutagenesis was used to modify two main N-glycosylation residues (N45, N64) in cellobiohydrolase I (CBH I) from Trichoderma reesei.Increased affinity for carboxymethylcellulose, an improved kcat/Km value, and stability up to 80 °C were shown by both mutants.[185]
The catalytic pocket of dehydrogenase/reductase EbSDR8 was reconstructed via targeted mutations at specific residues: G94 and S153.The mutants showed over 15-fold increase in the kcat/Km value relative to the wild-type. The increased catalytic efficiency was due to steric repulsion and C−H…π interactions.[168]
Two rounds of random mutagenesis were performed on a novel esterase (EstKa) from marine Klebsiella aerogenes to obtain the mutants I6E9 and L7B11.Mutants I6E9 and L7B11 showed 1.56- and 1.65-fold higher kcat/Km values relative to the wild-type EstKa. The Pro96Thr and Gln156Arg mutations increased catalytic efficiency by 1.29- and 1.48-fold, respectively. In contrast, the compound mutant Pro96Thr/Gln156Arg exhibited a 68.9% decrease in the Km value and a 1.41-fold increase in the kcat/Km value.[186]
The N- and C-terminal domains of keratinase KerSMD were replaced with those from a homologous protease, KerSMF, to generate three different mutants.N-terminal domain replacement increased the kcat/Km value by >2-fold, whereas C-terminal domain replacement increased the kcat/Km value by 1.5-fold. The kcat/Km value was enhanced by 1.75-fold when both the N- and C-terminal domains were replaced.[187]
The conformational dynamics of lysine cyclodeaminase from Streptomyces pristinaespiralis were modulated to address substrate and product inhibitions. The Val61-Val94-SpLCD variant was designed by mutating Ile61 and Ile94 residues associated with substrate and product delivery processes.The Val61-Val94-SpLCD variant showed improvements in the kcat, kcat/Km, Ki-LYS, and Ki-LPA values by 20, 4, 19, and 9 times, respectively. The expanded substrate and product delivery tunnels of the mutant reduced substrate and product inhibition.[162]
Saturation mutagenesis was performed on position 233 of loop 6 of an acidic, mesophilic GH5 cellulase from Gloeophyllum trabeum to create variants N233A and N233G.The N233A and N233G mutants showed increased specific activities of 27% and 70%, respectively, and catalytic efficiencies of 45% and 52%, respectively.[142]
Focused site-directed iterative saturation mutagenesis (FSISM) was applied to rationally design mutations at a non-catalytic cavity in 3-ketosteroid-Δ1-dehydrogenase (Δ1-KstD) from Mycobacterium smegmatis.All four mutants, H132M, L113F, V419W, and M51L, exhibited a ~30-fold increase in the kcat/Km value and a nearly 10-fold increase in specific activity.[188]
Site-directed mutagenesis was performed to create four variants based on multiple sequence and 3D structure alignments of endo-β-1,4-xylanase TmxB from Thermotoga maritima.The T74Y mutant showed a 1.3-fold increase in specific activity and a 1.6-fold increase in the kcat/Km value. The DY mutants (two amino acid insertions) showed a 1.2-fold increase in the kcat/Km value.[189]
Rational design was used to engineer a substrate-binding cavity and entrance tunnel in L-amino acid deaminase from Proteus mirabilis.The Q92A mutant exhibited an increased binding cavity volume (994.2 Å3), enzyme activity (191.36 U mg1), and kcat/Km value (1.23 mM1min1). The T436/W438A mutant also showed an increased binding cavity volume and kcat/Km value.[190]
Nanoparticle technologyCdSe/ZnS core/shell quantum dots (QDs) of different sizes were conjugated onto phosphotriesterase (PTE), followed by extensive characterization.There was a 2-fold increase in the kcat/Km value when the enzyme was conjugated with QDs. The results indicated that the increase in catalytic efficiency was due to the different microenvironment of the hydration layer of the PTE-QD bioconjugate.[148]
A proposed model system consisting of a detailed reaction scheme was used to test different possible interactions between a substrate and an enzyme conjugated with nanoparticles.There was a 1.7-fold increase in the kcat/Km value when the enzyme was conjugated to nanoparticles. The data revealed that the product release step becomes rate-limiting when the enzyme is attached to the nanoparticles.[149]
Loop dynamicsThe residues at the interface of the mobile loop of the NS2B-NS3pro complex, a two-component viral protease, were mutated via site-directed mutagenesis.The NS2B mobile loop mutants altered the rate of product release by causing a conformational change that activated the catalytic center. The glycine-linked S83W mutant had a 1.5-fold increase in the kcat/Km value relative to the wild-type.[143]
Incorporating non-canonical amino acidsNon-canonical amino acids were incorporated into the active site of an Escherichia coli transketolase variant (S385Y/D469T/R520Q) through site-specific saturation mutagenesis to produce different variants.A 43-fold increase in specific activity, a 290% increase in the Km value, and a 240% increase in the kcat/Km value were exhibited by the residue 385 variants.[171]
Tyr309 was replaced with two non-canonical amino acids, L-(7-hydroxycoumarin-4-yl)ethylglycine (Hco) and L-(7-methylcoumarin-4-yl)ethylglycine, in bacterial phosphotriesterase.The mutants with unnatural amino acids had increased phosphotriesterase turnover rates, up to 11-fold, and a kcat/Km value up to 4-fold. The deprotonated 7-hydroxyl group of Hco was observed to increase the rate-limiting product release rate via electrostatic repulsion of 4-nitrophenolate.[155]
The non-canonical amino acid, Nδ-methylhistidine, was used as a catalytic nucleophile to design a novel hydrolytic enzyme.The novel enzyme exhibited an over 9000-fold increase in the ester hydrolysis rate and a 15-fold increase in the kcat/Km value. Crystallographic analysis proved Nδ-methylhistidine as a genetically encodable surrogate for dimethylaminopyridine.[167]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Atampugbire, G.A.; Quaye, J.A.; Gadda, G. How Mechanistic Enzymology Helps Industrial Biocatalysis: The Case for Kinetic Solvent Viscosity Effects. Catalysts 2025, 15, 736. https://doi.org/10.3390/catal15080736

AMA Style

Atampugbire GA, Quaye JA, Gadda G. How Mechanistic Enzymology Helps Industrial Biocatalysis: The Case for Kinetic Solvent Viscosity Effects. Catalysts. 2025; 15(8):736. https://doi.org/10.3390/catal15080736

Chicago/Turabian Style

Atampugbire, Gabriel Atampugre, Joanna Afokai Quaye, and Giovanni Gadda. 2025. "How Mechanistic Enzymology Helps Industrial Biocatalysis: The Case for Kinetic Solvent Viscosity Effects" Catalysts 15, no. 8: 736. https://doi.org/10.3390/catal15080736

APA Style

Atampugbire, G. A., Quaye, J. A., & Gadda, G. (2025). How Mechanistic Enzymology Helps Industrial Biocatalysis: The Case for Kinetic Solvent Viscosity Effects. Catalysts, 15(8), 736. https://doi.org/10.3390/catal15080736

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop